mathematical foundations of quantum field theory

461

Upload: others

Post on 11-Sep-2021

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Mathematical Foundations of Quantum Field Theory
Page 2: Mathematical Foundations of Quantum Field Theory

Mathematical Foundations of Quantum Field Theory

11222_9789813278639_TP.indd 111222_9789813278639_TP.indd 1 1/4/20 2:47 PM1/4/20 2:47 PM

Page 3: Mathematical Foundations of Quantum Field Theory

Other Related Titles from World Scientific

Lectures of Sidney Coleman on Quantum Field Theory:Foreword by David Kaiseredited by Bryan Gin-ge Chen, David Derbes, David Griffiths, Brian Hill, Richard Sohn and Yuan-Sen TingISBN: 978-981-4632-53-9ISBN: 978-981-4635-50-9 (pbk)

Quantum Field Theoryby Harald FritzschISBN: 978-981-3141-72-8

Homotopical Quantum Field Theoryby Donald YauISBN: 978-981-121-285-7

From Classical Mechanics to Quantum Field Theory, a Tutorialby Manuel Asorey, Elisa Ercolessi and Valter MorettiISBN: 978-981-121-048-8

KahFee - 11222 - Mathematical Foundations of Quantum Field Theory.indd 1 02-01-20 2:03:10 PM

Page 4: Mathematical Foundations of Quantum Field Theory

NEW JERSEY • LONDON • SINGAPORE • BEIJING • SHANGHAI • HONG KONG • TAIPEI • CHENNAI • TOKYO

World Scientific

Mathematical Foundations of Quantum Field Theory

Albert SchwarzUniversity of California at Davis, USA

11222_9789813278639_TP.indd 211222_9789813278639_TP.indd 2 1/4/20 2:47 PM1/4/20 2:47 PM

Page 5: Mathematical Foundations of Quantum Field Theory

Published by

World Scientific Publishing Co. Pte. Ltd.5 Toh Tuck Link, Singapore 596224USA office: 27 Warren Street, Suite 401-402, Hackensack, NJ 07601UK office: 57 Shelton Street, Covent Garden, London WC2H 9HE

Library of Congress Cataloging-in-Publication DataNames: Shvar ts, A. S. (Alʹbert Solomonovich), author. Title: Mathematical foundations of quantum field theory / Albert Schwarz, University of California at Davis. Description: New Jersey : World Scientific, [2020] | Includes bibliographical references. Identifiers: LCCN 2019034999 | ISBN 9789813278639 (hardcover) Subjects: LCSH: Quantum field theory. Classification: LCC QC174.45 .S3295 2020 | DDC 530.14/3--dc23 LC record available at https://lccn.loc.gov/2019034999

British Library Cataloguing-in-Publication DataA catalogue record for this book is available from the British Library.

Copyright © 2020 by author

All rights reserved.

For any available supplementary material, please visit https://www.worldscientific.com/worldscibooks/10.1142/11222#t=suppl

Desk Editor: Ng Kah Fee

Typeset by Stallion Press Email: [email protected]

Printed in Singapore

KahFee - 11222 - Mathematical Foundations of Quantum Field Theory.indd 2KahFee - 11222 - Mathematical Foundations of Quantum Field Theory.indd 2 13/4/2020 3:02:27 pm13/4/2020 3:02:27 pm

Page 6: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page v

To the memory of my beloved wife Lucy

v

Page 7: Mathematical Foundations of Quantum Field Theory

b2530 International Strategic Relations and China’s National Security: World at the Crossroads

b2530_FM.indd 6 01-Sep-16 11:03:06 AM

This page intentionally left blankThis page intentionally left blankThis page intentionally left blankThis page intentionally left blank

Page 8: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page vii

Preface

This book is addressed to mathematicians and physicists, who are

interested in clear exposition in the foundations of quantum field

theory. I have tried very hard to satisfy both categories of readers.

I wanted the book to be accessible to a mathematician who does

not know quantum mechanics and interesting to a physicist who

specializes in quantum field theory. I aimed to have the rigor of the

proofs to be sufficient for a mathematician, but not so much that

it disturbed the reading for a physicist. I hope that this attempt to

satisfy these criteria is successful at least partly.

In this book, we talk almost exclusively about the results of

quantum field theory that do not depend on the assumption that the

theory is Lorentz-invariant (Lorentz-invariant theories are analyzed

only at the end of the book). This is the most essential difference that

sets this book apart from other books. Another important feature of

the book is the consideration of both the Hamiltonian and axiomatic

approaches to quantum field theory; we also establish the relation

between them.

In some existing books on quantum field theory, one can easily find

examples where the rules of the game (the main definitions) change

in the process of calculation. We can also see formal manipulations

with meaningless expressions, but in the result, we somehow obtain

a meaningful answer. This makes the study of quantum field theory

much more difficult for a mathematically inclined reader. Of course,

the reader understands that in changing the rules of the game,

physicists do not imitate the characters in the books of Lewis Caroll;

vii

Page 9: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page viii

viii Mathematical Foundations of Quantum Field Theory

instead, they are guided by physical intuition, broad use of analogies,

and experiment. However, even if a mathematician recognizes that

physicists are doing the right things, this does not solve his problems.

I have tried to give an exposition of the main notions of quantum

field theory in such a way that without aiming for full mathematical

rigor, we instead obtain maximal clarity. However, in most cases, a

qualified mathematician should be able to fill in the details following

the proof outlines sketched in this book.

The first two chapters of the book and Section 3.1 contain a short

introduction to quantum mechanics, intended for mathematicians. In

Sections 3.1, 6.1, 6.2 and 6.3, we present some basic facts about Fock

space and operators on it; the rest of the book builds on these facts.

Chapter 4 is dedicated to the study of the operator of evolution

in the interaction representation and its adiabatic analog. Chapter 5

presents the theory of potential scattering in quantum mechanics.

Section 8.1 of Chapter 8, is devoted to translation-invariant

Hamiltonians and their operator realizations. The quantization of

classical translation-invariant system with an infinite number of

degrees of freedom is studied in Section 8.3 of Chapter 8.

Chapter 9 contains descriptions of different constructions of the

scattering matrix of a translation-invariant Hamiltonian. The proof

of the equivalence of these constructions is given in Chapter 11.

Chapters 10 and 12 introduce axiomatic scattering theory (in

Chapter 12, we consider Lorentz-invariant theories).

In Chapter 11, we study translation-invariant Hamiltonians in

the framework of perturbation theory; this chapter uses the results

of Chapter 10 on axiomatic scattering theory and the definition of

the canonical Faddeev transformation in Section 9.4 of Chapter 9.

A mathematician should begin to read these sections with Sec-

tion 11.5 of Chapter 11.

Chapter 13, added to the English edition, contains applications

of the methods of the preceding chapters to statistical physics. An

advanced reader can start with this chapter, returning in the case of

necessity to Chapter 10 and to Introduction.

The mathematically inclined reader, after the first five chap-

ters and Sections 6.1 and 6.2 of Chapter 6, can go straight to

Page 10: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page ix

Preface ix

axiomatic scattering theory (Chapter 10 and Sections 12.1 and 12.2

of Chapter 12). After these, one may read Sections 8.1 and 8.3 of

Chapter 8, Section 12.3 of Chapter 12, Section 11.5 of Chapter 11,

and Chapter 13.

A physicist who wants to read the book on a rigorous level can

find the necessary mathematical definitions and theorems in the

appendix. If he is satisfied with a lower level of rigor, he should

begin his reading with the fourth chapter, since the material in the

first three chapters should be familiar. He can neglect the difference

between pre-Hilbert and Hilbert spaces, and the difference between

Hermitian operators and self-adjoint operators. A measurable func-

tion for a physicist is an arbitrary function and a measure space can

be understood as the n-dimensional Euclidean space (more precisely,

if functions on a set X can be integrated, then the set X can be

considered as a measure space). If the physicist encounters unfamiliar

mathematical notion, he can usually keep going without much harm.

The book generally does not contain references to the original

papers. (I have placed references to papers only when results are

formulated but not proved.)

I have received generous help in the production of this book.

I am grateful to Yu. Berezansky, F. Berezin, L. Faddeev,

V. Fateev, E. Fradkin, V. Galitsky, A. Povzner, M. Polivanov,

A. Rosly, V. Sushko, I. Todorov, Yu. Tyupkin, A. Vainshtein,

O. Zavyalov, and other mathematicians and physicists who kindly

devoted their attention to this book. I am also grateful to the

translator of the book, Dmitry Shemetov, for his diligent work and

for his patience.

I am also deeply indebted to my family for their support.

Page 11: Mathematical Foundations of Quantum Field Theory

b2530 International Strategic Relations and China’s National Security: World at the Crossroads

b2530_FM.indd 6 01-Sep-16 11:03:06 AM

This page intentionally left blankThis page intentionally left blankThis page intentionally left blankThis page intentionally left blank

Page 12: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xi

Contents

Preface vii

Introduction xvii

1. Principles of Quantum Theory 1

1.1 Principles of quantum mechanics . . . . . . . . . . . 1

1.2 Evolution of state vectors . . . . . . . . . . . . . . . 1

1.3 Calculating the probabilities . . . . . . . . . . . . . 2

1.4 Heisenberg operators . . . . . . . . . . . . . . . . . 5

1.5 Integrals of motion and stationary states . . . . . . 6

2. Quantum Mechanics of Single-Particle

and Non-Identical Particle Systems 9

2.1 Quantum mechanics of a single scalar particle . . . 9

2.2 Quantum mechanics of particles with spin . . . . . 11

2.3 Quantum description of a system with

non-identical particles . . . . . . . . . . . . . . . . . 14

2.4 A particle in a box with periodic

boundary conditions . . . . . . . . . . . . . . . . . . 16

2.5 One-dimensional harmonic oscillator . . . . . . . . . 17

2.6 Multidimensional harmonic oscillator . . . . . . . . 20

xi

Page 13: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xii

xii Mathematical Foundations of Quantum Field Theory

3. Quantum Mechanics of a System of Identical

Particles 23

3.1 A system of n identical particles . . . . . . . . . . . 23

3.2 Fock space . . . . . . . . . . . . . . . . . . . . . . . 27

4. Operators of Time Evolution S(t, t0)

and Sα(t, t0) 39

4.1 Non-stationary perturbation theory . . . . . . . . . 39

4.2 Stationary states of Hamiltonians depending

on a parameter . . . . . . . . . . . . . . . . . . . . 43

4.3 Adiabatic variation of stationary state . . . . . . . 46

5. The Theory of Potential Scattering 51

5.1 Formal scattering theory . . . . . . . . . . . . . . . 51

5.2 Single-particle scattering . . . . . . . . . . . . . . . 57

5.3 Multi-particle scattering . . . . . . . . . . . . . . . 64

6. Operators on Fock Space 75

6.1 The representations of canonical and

anticommutation relations:

Fock representation . . . . . . . . . . . . . . . . . . 75

6.2 The simplest operators on Fock space . . . . . . . . 82

6.3 The normal form of an operator: Wick’s

theorem . . . . . . . . . . . . . . . . . . . . . . . . 88

6.4 Diagram techniques . . . . . . . . . . . . . . . . . . 98

7. Wightman and Green Functions 107

7.1 Wightman functions . . . . . . . . . . . . . . . . . . 107

7.2 Green functions . . . . . . . . . . . . . . . . . . . . 111

7.3 Kallen–Lehmann representation . . . . . . . . . . . 116

7.4 The equations for Wightman and Green

functions . . . . . . . . . . . . . . . . . . . . . . . . 120

Page 14: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xiii

Contents xiii

8. Translation-Invariant Hamiltonians 123

8.1 Translation-invariant Hamiltonians

in Fock space . . . . . . . . . . . . . . . . . . . . . . 123

8.2 Reconstruction theorem . . . . . . . . . . . . . . . . 131

8.3 Interactions of the form V (φ) . . . . . . . . . . . . 141

9. The Scattering Matrix for Translation-

Invariant Hamiltonians 147

9.1 The scattering matrix for translation-invariant

Hamiltonians in Fock space . . . . . . . . . . . . . . 147

9.2 The definition of scattering matrix by means

of operator realization of a translation-invariant

Hamiltonian . . . . . . . . . . . . . . . . . . . . . . 154

9.3 The adiabatic definition of scattering matrix . . . . 166

9.4 Faddeev’s transformation and equivalence

theorems . . . . . . . . . . . . . . . . . . . . . . . . 170

9.5 Semiclassical approximation . . . . . . . . . . . . . 180

10. Axiomatic Scattering Theory 191

10.1 Main assumptions and the construction

of the scattering matrix . . . . . . . . . . . . . . . . 191

10.2 Proof of lemmas . . . . . . . . . . . . . . . . . . . . 207

10.3 Asymptotic fields (in- and out-operators) . . . . . . 222

10.4 Dressing operators . . . . . . . . . . . . . . . . . . . 229

10.5 Generalizations . . . . . . . . . . . . . . . . . . . . 240

10.6 Adiabatic theorem in axiomatic scattering

theory . . . . . . . . . . . . . . . . . . . . . . . . . 253

11. Translation-Invariant Hamiltonians (Further

Investigations) 275

11.1 Connections between the axiomatic theory and the

Hamiltonian formalism . . . . . . . . . . . . . . . . 275

11.2 Heisenberg equations and canonical

transformations . . . . . . . . . . . . . . . . . . . . 279

11.3 Construction of an operator realization . . . . . . . 286

Page 15: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xiv

xiv Mathematical Foundations of Quantum Field Theory

11.4 Dressing operators for translation-invariant

Hamiltonians . . . . . . . . . . . . . . . . . . . . . . 294

11.5 Perturbation theory via the axiomatic

approach . . . . . . . . . . . . . . . . . . . . . . . . 301

12. Axiomatic Lorentz-Invariant

Quantum Field Theory 313

12.1 Axioms describing Lorentz-invariant scattering

matrices . . . . . . . . . . . . . . . . . . . . . . . . 313

12.2 Axiomatics of local quantum field theory . . . . . . 319

12.3 The problem of constructing a non-trivial

example . . . . . . . . . . . . . . . . . . . . . . . . 324

13. Methods of Quantum Field Theory

in Statistical Physics 331

13.1 Quantum statistical mechanics . . . . . . . . . . . . 331

13.1.1 Examples . . . . . . . . . . . . . . . . . . . 334

13.2 Equilibrium states of translation-invariant

Hamiltonians . . . . . . . . . . . . . . . . . . . . . . 335

13.3 Algebraic approach to quantum theory . . . . . . . 336

13.3.1 Quantum field theory and statistical

physics in Rd . . . . . . . . . . . . . . . . . 338

13.3.2 Particles and quasiparticles . . . . . . . . . 340

13.3.3 Scattering . . . . . . . . . . . . . . . . . . . 341

13.3.4 Asymptotic behavior of 〈Q(x, t)Ψ,Ψ′〉 . . . 346

13.3.5 Scattering theory from asymptotic

commutativity . . . . . . . . . . . . . . . . 348

13.3.6 Green functions and scattering: LSZ . . . . 350

13.3.7 Generalized Green functions; the inclusive

scattering matrix . . . . . . . . . . . . . . . 353

13.4 L-functionals . . . . . . . . . . . . . . . . . . . . . . 355

13.4.1 Translation-invariant Hamiltonians in the

formalism of L-functionals;

one-particle states . . . . . . . . . . . . . . 360

13.4.2 Quadratic Hamiltonians . . . . . . . . . . . 361

13.4.3 Perturbation theory . . . . . . . . . . . . . 362

Page 16: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xv

Contents xv

13.4.4 GGreen functions . . . . . . . . . . . . . . . 365

13.4.5 Adiabatic S-matrix . . . . . . . . . . . . . . 367

13.4.6 Scattering of (quasi-)particles; inclusive

cross-section . . . . . . . . . . . . . . . . . 368

Appendix 375

A.1 Hilbert spaces . . . . . . . . . . . . . . . . . . . . . 375

A.2 Systems of vectors in a pre-Hilbert

vector space . . . . . . . . . . . . . . . . . . . . . . 376

A.3 Examples of function spaces . . . . . . . . . . . . . 377

A.4 Operations with Hilbert spaces . . . . . . . . . . . . 381

A.5 Operators on Hilbert spaces . . . . . . . . . . . . . 383

A.6 Locally convex linear spaces . . . . . . . . . . . . . 392

A.7 Generalized functions (distributions) . . . . . . . . 393

A.8 Eigenvectors and generalized eigenvectors . . . . . . 403

A.9 Group representations . . . . . . . . . . . . . . . . . 404

Bibliography 409

Index 413

Page 17: Mathematical Foundations of Quantum Field Theory

b2530 International Strategic Relations and China’s National Security: World at the Crossroads

b2530_FM.indd 6 01-Sep-16 11:03:06 AM

This page intentionally left blankThis page intentionally left blankThis page intentionally left blankThis page intentionally left blank

Page 18: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xvii

Introduction

The discovery of quantum field theory began with a rudimentary

form of quantum electrodynamics, the basic equations of which

were written at the end of the 1920s (Dirac, Heisenberg, Pauli).

These equations contain a small parameter e2

~c ≈1

137 , so it was

natural to try to find a solution in the framework of perturbation

theory with respect to this parameter. It turned out, however, that

only in the lowest orders did the perturbation theory provide finite

solutions consistent with experimental results. In higher orders, the

perturbation theory resulted in divergent integrals.

Significant progress was made only two decades later, beginning

with the work of Bethe (1947), Schwinger (1958) and Tomonaga

(1946) and finally culminating with Feynman’s (2005) method, which

made it possible to extract finite results from the divergent integrals

of higher orders of the perturbation theory with striking agree-

ment to experimental results (covariant theory of renormalization).

Feynman’s method turned out to be applicable to whole classes of

theories (to the so-called renormalizable theories); it was developed

in many works, among which it is necessary to mention the article

by Dyson (1949). The method was put on a solid mathematical

foundation in the article by Bogolyubov and Parasyuk (1955). These

groundbreaking papers started a stormy period of development of

QFT in many directions. These developments did not come without

delays and disappointments. Moreover, in these years, none of the

fundamental problems of QFT have been solved.

xvii

Page 19: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xviii

xviii Mathematical Foundations of Quantum Field Theory

Important steps forward were the discovery of non-Abelian

gauge fields, quantum chromodynamics, and the Higgs mechanism.

These developments led to the creation of the standard model that

describes electromagnetic, weak and strong interactions in very good

agreement with experiment. (The standard model assumes that

neutrinos are massless; hence one should modify it to take into

account the neutrino mass. This is the only deviation of standard

model from the experiment that has been discovered by now.)

However, physicists believe that the standard model comes from more

fundamental theory. They think that it comes from the theory where

electromagnetic, weak and strong interactions are on equal footing

(grand unification) or from theory that also describes gravity (like

string theory). Furthermore, we now significantly better understand

the mathematical structure of QFT. I would like to first mention the

algebraic approach to quantum field theory that is closely related to

the axiomatic quantum field theory (see, for example, Bogolyubov

et al. (1956), Lehmann et al. (1955), Wightman (1956), Haag (1958),

Araki and Haag (1967), Haag and Kastler (1964), Ruelle (1962),

Hepp (1965)). In particular, one should mention the construction

of a scattering theory in the axiomatic framework. In the case

of renormalizable theories, one can construct objects obeying the

axioms of relativistic QFT in the framework of perturbation theory,

but it is difficult to give a rigorous construction outside of this

framework. Intensive development in this direction received the name

of constructive QFT (see, for example, Jaffe (2000) and Velo and

Wightman (2012)); it was successful in dimensions < 4, but realistic

four-dimensional theories are still out of reach. Remarkable progress

in the study of conformal field theories and supersymmetric theories

(theories having symmetries mixing bosons and fermions) has led to

a much better understanding of QFT. Very important information

about QFT comes from string theory (one can obtain QFT from

string theory in some limit). Today, QFT presents an extensive field

of activity not only in physics but also in mathematics, being a source

of numerous clearly defined mathematical problems.

This book gives an introduction to QFT. It is in many ways

different from other books. In particular, we completely separate

the treatment of renormalization and divergences. It became clear

Page 20: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xix

Introduction xix

long ago that even in the absence of ultraviolet divergences, it is

impossible to use the standard quantum mechanical definition of

scattering matrix. The reason is that in QFT, there exists no natural

representation of full Hamiltonian in the form H0 + gV , where the

first term is a free (quadratic) Hamiltonian and the second term is

the interaction that can be considered as perturbation. This means

that the particles described by H0 (bare particles) are not the same

as the particles described by the full Hamiltonian (dressed particles).

This necessitates the change of the definition of the scattering matrix

and the renormalization of Feynman diagrams. This renormalization

is not related to ultraviolet divergences, a fact that is not emphasized

in many textbooks.

In this book, we do not assume Lorentz invariance in the study of

QFT; we discuss relativistic local theories only at the end. Therefore,

one can apply many of the statements to statistical physics with

minimal changes. This is the reason why I have added a new

chapter devoted to the applications of methods of QFT to statistical

physics. In particular, the new chapter contains the description

of the formalism based on the consideration of states as positive

functionals on Weyl or Clifford algebra (L-functionals) (Schwarz,

1967; Tyupkin, 1973; Schwarz, 2019b). This formalism allows us to

derive the diagram techniques of thermo-field dynamics (TFD) that

coincide in the case at hand with Keldysh diagrams (see Chu and

Umezawa (1994) for the review of TFD and Keldysh formalism).

Some of notions and results of the last chapter are based on recent

papers (Schwarz, 2019b,c). I would like to mention the notion of

inclusive scattering matrix and its expression in terms of generalized

Green functions. (Inclusive scattering matrix is closely related to

the inclusive cross-section that is necessary in the consideration of

scattering in the case when the theory does not have a particle

interpretation, in particular, for the consideration of collisions of

thermal quasiparticles.)

The exposition in the book was influenced by the algebraic

approach to quantum theory. However, the starting point is the

standard Hilbert space formulation. I am including in this edition a

short review of some questions of quantum theory based on algebraic

approach. For some readers, it would be reasonable to start with

Page 21: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xx

xx Mathematical Foundations of Quantum Field Theory

this review, while others could read it after finishing the book or in

parallel with reading the book. An advanced reader can jump to the

last chapter after reading the review.

Review of algebraic approach to quantum theory

First of all some terminological conventions.

Talking about algebra, we always have in mind unital (having unit

element) associative algebra over complex numbers with involution

denoted by ∗. (An involution is an antilinear map A → A∗ obeying

A∗∗=A, (AB)∗=B∗A∗.) We assume that the algebra is a topological

space and all operations are continuous, but for many of our

statements, these requirements are not sufficient (one should assume

that we have a Banach algebra or C∗-algebra or impose some other

conditions). In our terminology, an automorphism preserves not only

operations in the algebra but also the involution.

We say that h is a derivation of the algebra A if h(AB) =

h(A)B + Ah(B); we also assume that the derivation is compatible

with involution.

We say that a derivation is an infinitesimal automorphism if the

equation idAdt = h(A(t)) has a solution for all initial data A(0); then

a map A(0)→ A(t) specifies an automorphism αt. It is easy to check

that αt+s = αtαs (an infinitesimal automorphism generates a one-

parameter group of automorphisms). One can consider infinitesimal

automorphisms as elements of the Lie algebra of the group of

automorphisms. (This Lie algebra should be defined on the vector

space of tangent vectors to one-parameter families of automorphisms

at the unit element of the group. In the infinite-dimensional case, it

is not clear whether such a vector can also be considered as a tangent

vector to a one-parameter group of automorphisms. In what follows,

we disregard these subtleties.)

Hamiltonian formalism

The equations of motion of a three-dimensional non-relativistic

particle in a potential field U(~x) have the form

d~p

dt= −∇U,

Page 22: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xxi

Introduction xxi

where ~p = m~dxdt stands for the momentum of the particle. To solve

these equations (i.e. to find the trajectory of the particle), we should

know the initial data: the coordinates and the momenta at some

moment of time. One says that the coordinates and the momenta

specify the state of our particle at the given moment and that the

equations of motion allow us to find the state of the particle at any

moment if we know it at one of the moments. The equations of motion

can be written in the form

d~p

dt= −∂H

∂~x,

~dx

dt=∂H

∂~p,

where H = p2

m + U(~x) is called Hamiltonian function (one can

say the Hamiltonian function is the energy expressed in terms

of momenta and coordinates). Similar equations are valid for any

mechanical system, but the number of degrees of freedom (the

number of coordinates and momenta) and the Hamiltonian function

can be arbitrary. This gives the so-called Hamiltonian formalism

of mechanics. (In Lagrangian formalism, the state is specified by

coordinates and velocities.)

In Hamiltonian formalism, the (pure) state of a classical mechan-

ical system (at the time t) is characterized by 2n numbers: p =

(p1, . . . , pn) (generalized momenta) and q = (q1, . . . , qn) (gener-

alized coordinates). (Together, these numbers specify a point of

2n-dimensional space called the phase space of the system).

More generally, we can define a state as a probability distribution

on the phase space. The set D of probability distributions is a convex

set, the pure states can be identified with extreme points of this set.

Every state can be considered as a mixture of pure states. (The

mixture of states ω1, . . . , ωn with probabilities p1, . . . , pn is the state

p1ω1+· · ·+pnωn. If states are labeled by continuous parameter λ ∈ Λ,

one defines the mixture of the states as an integral∫ω(λ)ρ(λ)dλ

where ρ(λ) stands for the density of the probability distribution

on Λ. Note that the definition of mixture can be used for any

convex set.)

Page 23: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xxii

xxii Mathematical Foundations of Quantum Field Theory

The evolution of a pure state is governed by Hamiltonian equa-

tions

dq

dt=∂H

∂p,dp

dt= −∂H

∂q, (I.1)

where the function H(p, q, t) is called the Hamiltonian.

To write down the equation of motion for general state, we

introduce the notion of Poisson bracket of two functions on the phase

space by the formula

f, g = −∂f∂p

∂g

∂q+∂f

∂q

∂g

∂p. (I.2)

It is easy to check that the Poisson bracket is antisymmetric and

satisfies Jacobi identity, hence it specifies a structure of Lie algebra

on functions on phase space.1

Denoting by ρ(p, q, t) the density of the probability distribution

on the phase space at the moment t, we obtain the equation

d

dtρ(p, q, t) = H, ρ(p, q, t), (I.3)

governing the evolution of state (Liouville equation). If U(t) denotes

the evolution operator (the operator transforming the state at the

moment 0 into the state at the moment t) we can write (I.3) in the

form

d

dtU(t) = LU(t),

where Lρ = H, ρ. Note that ρ is in general a generalized function

on phase space. To verify (I.3), it is sufficient to check that for pure

states (represented by δ-functions) it is equivalent to (I.1).

1The multiplication and the Poisson bracket specify the structure of Poissonalgebra on the space of functions on the phase space (see the definition of Poissonalgebra in Section 1.3). This means that the phase space is a Poisson manifold.Moreover, it is a symplectic manifold, i.e. the Poisson structure is non-degenerate.We have considered the Hamiltonian formalism on a flat symplectic manifold, butit can be considered on any symplectic manifold.

Page 24: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xxiii

Introduction xxiii

A physical quantity (an observable) can be considered as a real

function f(p, q) on the phase space. It follows from the chain rule

that

d

dtf(p(t), q(t)) = −∂f

∂p

∂H

∂q+∂f

∂q

∂H

∂p. (I.4)

One can rewrite (I.4) in the form

d

dtf(p(t), q(t)) = f,H. (I.5)

It follows that in the case f,H = 0, the expression f(p(t), q(t))

does not depend on time (in other words, f is an integral of motion).

In particular, if H does not depend on time, the function H(p, q)

is an integral of motion. It can be identified with the energy of the

system.

Let us denote by A the set of all complex continuous functions

on the phase space considered as an algebra with respect to the

conventional addition and multiplication of functions. Formula (I.5)

gives an equation for the evolution in the algebra A.

Note that every state ω (considered as a measure on the phase

space) specifies a linear functional on A by the formula ω(f) =∫fω.

This functional obeys the positivity condition: ω(f) ≥ 0 if f ≥ 0. The

evolution of states and the evolution in A are related by the formula

(ω(t))(f) = ω(f(t)).

Quantum mechanics: An algebraic approach

The picture of the preceding section can be modified to describe

quantum mechanics. The main idea is to allow non-commuting

physical quantities.

The starting point is a unital associative algebra A over complex

numbers equipped with an antilinear involution A → A∗ (generaliz-

ing complex conjugation in the algebra A of the preceding section).

States are identified with positive linear functionals on A (linear

functional ω is positive if ω(A∗A) ≥ 0). We assume that states are

normalized, i.e. ω(1) = 1. The set D of normalized states is convex.

Page 25: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xxiv

xxiv Mathematical Foundations of Quantum Field Theory

The extreme points of this set are called pure states. Every state is

a mixture of pure states.

Let us denote by Aut the group of automorphisms of the algebra

A commuting with involution. This group naturally acts on states.

In quantum system, the state depends on time and this dependence

can be described by the evolution operator U(t) transforming ω(0)

into ω(t). This is the so-called Schrodinger picture; it is equivalent

to Heisenberg picture where the elements of A depend on time, but

the states do not: ω(t)(A) = ω(A(t)).

The evolution operator satisfies the equation

idU

dt= H(t)U(t), (I.6)

which is equivalent to the equation of motion idωdt = H(t)ω(t). Here,

H(t) stands for an element of Lie algebra of the group Aut (for

infinitesimal automorphism). It plays the role of the Hamiltonian

of the quantum system. If H does not depend on t, the evolution

operators obey U(t + τ) = U(t)U(τ) (constitute a one-parameter

subgroup).

To specify a quantum system, we should fix an algebra with

involution A and an infinitesimal automorphism H (or a family of

infinitesimal automorphisms H(t)). Then the evolution is governed

by (I.6).

In what follows, we assume that H does not depend on t unless

the dependence on t is explicitly mentioned.

Note that an infinitesimal automorphism can be considered as

a derivation of the algebra A. However, a derivation specifies a

quantum system only if it can be integrated to a one-parameter group

of automorphisms.

The textbook form of quantum mechanics corresponds to the case

when A is the algebra of bounded linear operators in Hilbert space

E with involution defined as Hermitian conjugation. The states are

specified by density matrices (positive operators with unit trace)

by the formula ωK(A) = TrKA. Pure states correspond to vectors

Ψ ∈ E ; proportional vectors specify the same state. If the vector Ψ

is normalized, the corresponding state is the functional 〈AΨ,Ψ〉.

Page 26: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xxv

Introduction xxv

Every unitary operator W determines an automorphism of the

algebra with involution A by the formula A → W−1AW . This

correspondence allows us to identify the group Aut with the group of

unitary operators and infinitesimal automorphisms with self-adjoint

operators. The equation of motion for the density matrix has the form

idKdt = HK, where H is an operator on the space L of trace class

operators in E defined by the formula HK = −HK + KH and the

equation of motion for the vector Ψ ∈ E representing the pure state

is idΨdt = HΨ where H is a self-adjoint operator (Hamiltonian). The

evolution operator in E will be denoted by U(t); it obeys idUdt = HU .

The evolution operator in L will be denoted by U(t); it is easy to

check that U(t)K = U−1(t)KU(t).

Note that in the case at hand pure states have very simple

description, therefore, very often it is convenient to work with pure

states and to consider other states as mixtures of pure states. In

principle, we can work only with pure states for any algebra A, but

in general, this is not convenient because the description of pure

states is complicated.

For any algebra A and any state ω, we can construct a pre-Hilbert

space E and a representation of A by operators in this space such

that for some cyclic vector Φ ∈ E , we have ω(A) = 〈AΦ,Φ〉. (An

element A ∈ A is represented by operator A; we assume that the

map A → A is an algebra homomorphism and is compatible with

involution: A∗ = (A)∗. The vector Φ is cyclic in the following sense:

every other vector can be represented in the form AΦ where A ∈ A.)

This construction (Gelfand–Naimark–Segal (GNS) construction) is

essentially unique (up to equivalence).

Let us sketch the proof of this theorem. Assume that the

representation we need is constructed. Let us introduce in A an inner

product by the formula 〈A,B〉 = ω(B∗A). It is easy to see that the

map ν : A → E sending A to AΦ preserves this inner product. It

follows from cyclicity of Φ that this map is surjective; this allows us

to identify E with the quotient of A with respect to zero vectors.

(Recall that a zero vector is a vector that is orthogonal to all other

vectors.) The obvious relation ν(BA) = Bν(A) allows us to describe

our representation in terms of the algebra A and state ω. Namely,

Page 27: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xxvi

xxvi Mathematical Foundations of Quantum Field Theory

we construct E as A factorized with respect to zero vectors of the

inner product 〈A,B〉 = ω(B∗A), the operation of the multiplication

from the left by A descents to the operator A, the unit element of Acorresponds to the vector Φ.

We worked with pre-Hilbert spaces, but we can take a completion

of E to obtain a representation of A by operators acting in Hilbert

space.

We see that every state of A can be represented by a vector in

Hilbert space. However, we cannot consider all states as elements of

the same Hilbert space.

If ω is a stationary state (a state invariant with respect to time

evolution), then the group U(t) descends to a group U(t) of unitary

transformations of corresponding space E . The generator H of U(t)

plays the role of Hamiltonian.

We say that the stationary state ω is a ground state if the

spectrum of H is non-negative.

This definition agrees with the definition of the ground state in

Hilbert space formulation of quantum mechanics. Let us apply the

GNS construction to the algebra of bounded operators in Hilbert

space E and to a state represented by a vector Φ ∈ E where Φ is

an eigenvector of the Hamiltonian: HΦ = EΦ. Then the space given

by GNS construction can be identified with E and the generator of

U(t) is equal to H − E. The condition that Φ is the eigenstate with

minimal eigenvalue is equivalent to the positivity of H − E.

The representation containing the ground state will be called

ground state representation. This representation is especially impor-

tant in quantum field theory: we will consider particles as elementary

excitations of ground state.

Let us define the correlation functions in a stationary state ω as

functions

wn(t1, . . . , tn) = ω(A1(t1) . . . An(tn)),

where A1, . . . , An ∈ A.

The Green functions in the state ω are defined by the formula

Gn(t1, . . . , tn) = ω(T (A1(t1) . . . An(tn))),

Page 28: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xxvii

Introduction xxvii

where T stands for time ordering. It is easy to express them in terms

of correlation functions.

The correlation functions in the ground state are called Wightman

functions. The properties of particles in quantum field theory can be

expressed in terms of these functions and/or corresponding Green

functions; the same is true for scattering matrix.

Classical and quantum

To relate quantum and classical mechanics, we consider a family

of algebras A~ depending smoothly on the parameter ~ assuming

that for ~ = 0, we have a commutative algebra with the product

that will be denoted A ·B. More precisely, we assume that all these

algebras are defined on the same vector space, in other words, the

addition and multiplication by a number do not depend on ~, but the

multiplication of elements A,B of the algebra (denoted by A ·~ B)

smoothly depends on ~. The commutator [A,B]~ = A ·~ B − B ·~ Avanishes for ~ = 0, therefore, we can introduce a new operation

A,B (Poisson bracket) using the formula

[A,B]~ = iA,B~ +O(~2).

It is easy to verify that A0 (the algebra with commutative

multiplication A ·B that we have for ~ = 0) with the new operation

is Poisson algebra, i.e. the new operation satisfies the axioms of Lie

algebra and

A ·B,C = A,C ·B +A · B,C.

If there exists an involution A→ A∗ compatible with multiplica-

tion in all algebras A~, then it is also compatible with the Poisson

bracket (i.e. the Poisson bracket of two self-adjoint elements is again

a self-adjoint element). This statement is one of the reasons why a

factor i is included in the definition of the Poisson bracket.

Quantization; Weyl algebra

We have found that the classical mechanics can be obtained from

quantum mechanics in the limit ~ → 0. Conversely, quantum

Page 29: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xxviii

xxviii Mathematical Foundations of Quantum Field Theory

mechanics can be obtained as a deformation of classical mechanics.

We can start with Poisson algebra and deform it (this means we

would like to construct a family A~ that gives this Poisson algebra

in the limit ~→ 0).

If a Poisson algebra A is an algebra of polynomial functions on a

vector space with coordinates (u1, . . . , un), then the Poisson bracket

can be written in the form

A,B =1

2σkl(u)

∂A

∂uk∂B

∂ul. (I.7)

It specifies a structure of Poisson manifold on the vector space. (Here,

uk, ul = σkl(u).)

Let us consider the case when the Poisson bracket on polynomial

functions on vector space is defined by the formula (I.7) with constant

coefficients σkl. Moreover, we assume that the matrix σkl is non-

degenerate, then the dimension of the vector space is necessarily

even. (This is the situation in Section 1.1.) Then we can define the

algebra A~ as a unital associative algebra with generators uk and

relations ukul − uluk = iσkl. This algebra is called Weyl algebra.

If the coordinates uk are considered as real numbers, we introduce

an involution in Weyl algebra requiring that the generators uk are

self-adjoint. If the Poisson bracket is written in the form (I.2) (we

can always write it in this form changing coordinates), then the Weyl

algebra is generated by self-adjoint elements (p1, . . . , pn, q1, . . . , qn)

with relations pkpl = plpk, qkql = qlqk, pkq

l − qlpk = ~i δlk. These

relations are called canonical commutation relations (CCR). Instead

of self-adjoint generators (p1, . . . , pn, q1, . . . , qn), one can consider

generators ak, a∗k where ak = 1√

2(qk + ipk), a

∗k = 1√

2(qk− ipk). These

generators satisfy relations akal = alak, a∗ka∗l = a∗l a

∗k, aka

∗l − a∗l ak =

~δkl. These relations are also called CCR. To say that the family of

algebras A~ can be considered as a deformation of the commutative

polynomial algebra, we should realize their elements as polynomials.

This can be done in many different ways. For example, we can note

that using CCR we are able to move all operators qk to the left

and all operators pk to the right; “removing hats” in the expression,

we obtain a polynomial called (q − p) symbol of the element of

Page 30: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xxix

Introduction xxix

Weyl algebra. This representation of elements of Weyl algebra by

polynomials does not agree with involution ((q − p)-symbol of self-

adjoint element is not necessarily real), however, it is easy to modify

the construction to avoid this drawback. For example, one can write

down an element of Weyl algebra in terms of generators ak, a∗k and use

CCR to move a∗k to the left and ak to the right. The expression we get

is called normal form of the element of Weyl algebra. Considering ak,

a∗k in the normal form as complex variables, we obtain a polynomial

called Wick symbol.

We can go in opposite direction and obtain an element of Weyl

algebra from a polynomial. This operation is called quantization.

Quantization allows us, for example, to obtain quantum Hamiltonian

from classical Hamiltonian. It is important to note that the quantiza-

tion depends on the choice of symbol (we have “ordering ambiguity”).

However, for some classical Hamiltonians, there exists a natural

choice of corresponding quantum Hamiltonians. In particular, this

is true for Hamiltonians represented as a sum of kinetic energy

expressed as a function of momenta and potential energy depending

on coordinates (no ordering ambiguity).

Stationary states

A state that does not depend on time is called stationary state. In

what follows, we work in the formalism where the states are described

by density matrices in Hilbert space E . The state represented by

a density matrix K is stationary if K obeys HK = 0, i.e. if

K commutes with the Hamiltonian H. (Recall that H acts as a

commutator with H.) An important particular case of a stationary

state is the Gibbs stateK = Z−1e−βH where Z = Tre−βH . This state

corresponds to the equilibrium state with the temperature T = β−1.

It tends to the ground state as T → 0.

Let us assume that the operator H has discrete spectrum with

orthonormal basis φn of eigenvectors with eigenvalues En. Then it

is convenient to work in representation where H is represented by

a diagonal matrix with entries En (it is called H-representation).

In this representation, the eigenvectors of the operator H in the

space L are matrices ψmn having only one non-zero entry equal

Page 31: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xxx

xxx Mathematical Foundations of Quantum Field Theory

to 1 in the position (m,n). Alternatively, one can define ψmn as

an operator acting by the formula ψmnx = 〈x, φn〉φm where x ∈ E .

Corresponding eigenvalues are Em −En. We see that the vectors φnand the corresponding density matrices ψnn are stationary states.

It follows that all diagonal density matrices in H-representation are

stationary states. Moreover, if the spectrum is simple, all stationary

states are represented by diagonal matrices.

A density matrix K0 represents a ground state if Tr HK0 ≤Tr HK for any density matrix K. If the ground state is unique, it

is necessarily pure and corresponds to the eigenfunction of H with

lowest eigenvalue.

Adiabatic approximation

Let us consider the case of slowly varying Hamiltonian H(t). We

will assume that the energy levels En(t) are distinct and vary

continuously with t, corresponding eigenvectors will be denoted by

φn(t). We assume that these eigenvectors constitute an orthonormal

system. Then it is easy to prove that in the first approxima-

tion the evolution operator U(t) transforms eigenvector to eigen-

vector

U(t)φn(0) = e−iαn(t)φn(t),dαn(t)

dt= En(t). (I.8)

To verify (I.8), we check that the RHS satisfies the equation of motion

up to terms that are small for slow varying H(t) (see Section 4.3 for

more details).

Let us introduce the operators ψmn(t) by the formula ψmn(t)x =

〈x, φn(t)〉φm(t). (These operators are eigenvectors of the operator

H(t) in the space L.) Applying (I.8) or analyzing directly the

evolution operator U(t) in L, we obtain that U(t) transforms

eigenvector into eigenvector

U(t)ψmn(0) = e−iβmn(t)ψmn(t),dβmn(t)

dt= Em(t)− En(t) (I.9)

(this equation is true up to terms that can be neglected for slowly

varying Hamiltonian). Note that βmm does not depend on t.

Page 32: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xxxi

Introduction xxxi

Decoherence

Let us consider a quantum system (atom, molecule, etc.) described

by a Hamiltonian H with simple discrete spectrum. We assume that

the system is “almost closed” in the following sense: the interaction

with the environment can be described as an adiabatic change of

the Hamiltonian H. Let us consider the evolution operator U(T )

assuming that H(0) = H(T ). If the Hamiltonian is time-independent,

then for the eigenvector φn with the eigenvalue En, we can say that

U(T )φn = Cn(T )φn where Cn(T ) = e−iEnT . If the Hamiltonian is

slowly changing, we have the same formula with Cn(T ) calculated

from (I.8). Hence, if we started with pure stationary state, we remain

in the same state. Similarly, U(T )ψmn = Cmn(T )ψmn and the phase

factor is constant for m = n.

It is natural to assume that the environment is random (the

time-dependent Hamiltonian depends on some parameters λ ∈ Λ

with some probability distribution on Λ), then for m 6= n, we have

a random phase factor Cmn(λ, T ). If we start with density matrix

K =∑kmnψmn (with matrix entries kmn in H-representation), then

the density matrix Kλ(T ) is equal to∑Cmn(λ, T )kmnψmn, i.e. the

matrix entries acquire phase factor Cmn(λ, T ). Now, we should take

the mixture K(T ) of states Kλ(T ) (this means that we should take

the average of phase factors). It is obvious that non-diagonal entries

of K(T ) are smaller by absolute value than corresponding entries of

K. Imposing some mild conditions on the probability distribution on

Λ, one can prove that the non-diagonal entries of K(T ) tend to zero

as T → ∞. In other words, the matrix K(T ) tends to a diagonal

matrix K having the same diagonal entries as K. (See Schwarz and

Tyupkin (1987) and Schwarz (2019a) for more details.)

The matrix K can be considered as a mixture of pure states

corresponding to the vectors φn with probabilities knn.

This phenomenon is known as decoherence.

Observables and probabilities

Until now, we did not relate the formalism of quantum mechanics

to experiment. We know that (at least in some cases) quantum

Page 33: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xxxii

xxxii Mathematical Foundations of Quantum Field Theory

mechanics can be considered as a deformation of classical mechanics,

therefore, one can conjecture that classical observables correspond

to quantum observables. In particular, remembering that the energy,

which is represented by Hamiltonian function, is an integral of motion

if the system is invariant with respect to time shift should have

quantum analog with similar properties. This is not quite true: a

quantum system does not have a definite energy and measuring its

energy by means of macroscopic device we obtain different values

of energy with some probabilities. (A similar thing happens when

we are trying to use a thermometer to measure the temperature of

a system that is out of equilibrium. Such a system does not have

a temperature, but still we can measure it; thermometer readings

will be different, but we will obtain some probability distribution

of these readings.) The interaction with macroscopic device leads to

decoherence (the non-diagonal matrix entries of the density matrix

K in H-representation die) and we obtain a mixture of pure states

with probabilities knn. If the density matrix K corresponds to a

pure state described by a vector φ in E , then in general, decoherence

leads to mixture of pure states corresponding to the vectors φn with

probabilities pn = |〈φ, φn〉|2. However, if φ coincides with one of

eigenfunction φk, all probabilities pn vanish except pk = 1. Therefore,

we say that an eigenfunction of H with eigenvalue E has definite

value of energy equal to E. For any other normalized vector φ, we

can speak only about a probability to get some value E measuring

the energy. This probability is non-zero only for eigenvalues En of

H; it is equal to pn = |〈φ, φn〉|2. For a state represented by density

matrix K, the probability is given by the formula pn = 〈Kφn, φn〉.Other observables are represented by self-adjoint operators; the

formulas we have written for energy remain valid for any observable.

Again, if the physical quantity A is represented by self-adjoint

operator A in E , the quantity A has definite value a if a state is

represented by an eigenvector of A with eigenvalue a. Otherwise, we

can talk only about probabilities. It is convenient to write the density

matrix K representing a state in A-representation; if the operator

A has simple discrete spectrum, the probabilities are equal to the

diagonal entries of K in A-representation.

Page 34: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xxxiii

Introduction xxxiii

In the general case, we can specify the probability distribution of

the observable A by the formula

f(A) = TrKf(A), (I.10)

where f stands for any piecewise continuous function and f(A) =∫f(a)ρ(a)da denotes the mean value of f(A) with respect to the

probability distribution ρ(a)da. (It is also called the expectation

value of f(A).) If the operator A has simple discrete spectrum, (I.10)

is equivalent to the formulas for probabilities we gave in this case.

If A1, . . . , An are commuting self-adjoint operators, one can define

the joint probability distribution of corresponding physical quantities

A1, . . . , An using the formula

f(A1, . . . , An) = TrKf(A1, . . . , An). (I.11)

More generally, if quantum mechanical system is specified by

C∗-algebra A and a state by a positive functional ω, then the joint

probability distribution of physical quantities A1, . . . , An is given by

the formula

f(A1, . . . , An) = ω(f(A1, . . . , An)). (I.12)

(Physical quantities Ai correspond to self-adjoint elements Ai of A.

The assumption that A is a C∗-algebra guarantees that the notion

of a function of a family of commuting self-adjoint elements makes

sense.)

Integrals of motion

One can work either in the Schrodinger picture where states are

time-dependent, but observables do not depend on time, or in the

Heisenberg picture, where states do not depend on time, but the

observables do. These pictures are equivalent:

TrK(t)f(A) = TrKf(A(t))

if A(t) obeys the Heisenberg equation

idA

dt= [H, A].

Page 35: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xxxiv

xxxiv Mathematical Foundations of Quantum Field Theory

This implies that the observable A commuting with the Hamiltonian

H is an integral of motion (corresponding probabilities do not depend

on time).

Let us suppose that we have a one-parameter family U(t)

of symmetries of the system. (A symmetry is an automorphism

preserving the equations of motion. In our case, we consider it as

a unitary operator, commuting with the Hamiltonian.) The operator

idUdt is an integral of motion — a self-adjoint operator commuting

with the Hamiltonian.

The Hamiltonian itself is an integral of motion corresponding to

the time translation. We have already noted that the corresponding

observable is energy.

The integral of motion corresponding to invariance with respect

to spatial translations is a component of momentum.

The integral of motion corresponding to the invariance with

respect to rotation around some axis is a component of angular

momentum.

Weyl and Clifford algebras

In quantum theory in algebraic approach, we are starting with a

unital associative algebra with involution. We have seen already that

a natural candidate for this algebra is Weyl algebra. In this section,

we will study this algebra and its cousin, Clifford algebra. Recall

that the Weyl algebra is generated by self-adjoint elements pi, qi

with relations

pkpl = plpk, qkql = qlqk, pkql − qlpk =

1

iδlk. (I.13)

These relations are called canonical commutation relations (CCR).

Instead of self-adjoint generators pi, . . . , qi, one can consider gen-

erators ak, a∗k where ak = 1√

2(qk + ipk), a

∗k = 1√

2(qk − ipk). These

generators satisfy relations

akal = alak, a∗ka∗l = a∗l a

∗k, aka

∗l − a∗l ak = δkl. (I.14)

Both (I.13) and (I.14) are called CCR.

For finite number of degrees of freedom, there exists only one

irreducible representation of CCR. The relations (I.13) can be

Page 36: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xxxv

Introduction xxxv

represented in the space L2(Rn) by operators of multiplication and

differentiation: qkψ(q1, . . . , qn) = qkψ(q1, . . . , qn), pkψ(q1, . . . , qn) =1i∂∂qk

ψ(q1, . . . , qn). In non-relativistic quantum mechanics, this rep-

resentation is very convenient. If the classical Hamiltonian has the

form T (p) + V (q) where the kinetic energy T (p) is a quadratic

function of momenta and the potential energy V (q) has a non-

degenerate minimum, then in a neighborhood of this minimum, the

classical system can be approximated by multidimensional harmonic

oscillator. Similarly, for the quantized Hamiltonian H = T (p)+V (q),

we can approximate low energy levels as energy levels of quantum

oscillator. The Hamiltonian of one-dimensional quantum oscillator

can be written in the form

H =p2

2+q2

2= a∗a+

1

2.

Noting that Ha∗ = a∗(H + 1), Ha = a(H − 1), we obtain that the

operator a∗ transforms an eigenfunction φ with eigenvalue E in an

eigenfunction with eigenvalue E + 1. Similarly, the operator a either

sends φ to zero (if φ is the ground state) or to an eigenfunction with

eigenvalue E−1. Using this fact, we can check that the ground state

θ has energy 12 and the states 1√

n!(a∗)nθ constitute an orthonormal

basis consisting of eigenfunctions of H with eigenvalues n+ 12 .

In appropriate coordinates, the Hamiltonian of multidimensional

quantum oscillator can be considered as a sum of non-interacting

one-dimensional oscillators:

H =∑(

ωka∗kak +

ωk2

);

again applying many times operators a∗k to the ground state θ, we

obtain a basis of eigenfunctions.

We see that it is convenient to use the operators a∗k, ak in the

analysis of excitations of the ground state. In quantum field theory,

we are interested first of all in the excitations of ground state; this

is one of the many reasons why these operators are so useful.

Let us consider now the Weyl algebra with infinite number of

generators. This notion can be made precise in different ways. We

Page 37: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xxxvi

xxxvi Mathematical Foundations of Quantum Field Theory

can consider simply an algebra with infinite number of generators

obeying (I.14), but it is more convenient to start with some (pre)-

Hilbert space B and consider an algebra with generators a(f), a∗(f)

and relations

a(λf + µg) = λa(f) + µa(g), a∗(f) = (a(f))∗,

a(f)a(g) = a(g)a(f), a∗(f)a∗(g) = a∗(g)a∗(f),

a(f)a∗(g)− a∗(g)a(f) = 〈f, g〉.(I.15)

A representation of Weyl algebra (= representation of CCR) is

a family of operators in a (pre)-Hilbert space obeying (I.15). (Note

that we use the same notation for elements of Weil algebra and for

operators.)

If B = l2, then a(f) =∑fkak, a

∗(f) =∑fka∗k where ak, a

∗k

obey (I.14). If B is some space of test functions on Rn (for example,

the Schwartz space), we can say that the formal expression a(f) =∫f(x)a(x)dx specifies a(x) as a generalized operator function on Rn.

We can also use this notation and terminology in cases when B is a

space of functions on some set.

The theory of Weyl algebra is very similar to the theory of

Clifford algebra (an algebra that is defined by canonical anti-

commutation relations (CAR) where commutators are replaced by

anticommutators). The representations of Clifford algebra are also

called representations of CAR.

In the simplest form, Clifford algebra can be defined as an algebra

with generators obeying

akal = −alak, a∗ka∗l = −a∗l a∗k, aka∗l + a∗l ak = δkl. (I.16)

More generally, to define the Clifford algebra, we start with some

(pre)-Hilbert space B and consider an algebra with generators

a(f), a∗(f) and relations

a(λf + µg) = λa(f) + µa(g), a∗(f) = (a(f))∗,

a(f)a(g) = −a(g)a(f), a∗(f)a∗(g) = −a∗(g)a∗(f),

a(f)a∗(g) + a∗(g)a(f) = 〈f, g〉.

(I.17)

Page 38: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xxxvii

Introduction xxxvii

Note that both Weyl and Clifford algebras can be extended in

various ways (for example, one can introduce some norm and consider

a completion with respect to this norm).

Weyl algebra or Clifford algebra (or tensor product of these

algebras) can play the role of the algebra with involution A in the

algebraic description of quantum system.

In this description, we also need an infinitesimal automorphism

H of the algebra. The simplest way to construct it is to take a

self-adjoint element h of the algebra and to define a derivation

by the formula HA = [h,A]. (This construction works for any

algebra; derivations obtained this way are called inner derivations. In

C∗-algebra, one can prove that an inner derivation is an infinitesimal

automorphism considering eith.) If the algebra is generated by ak, a∗k

obeying (I.14) or (I.16), the element h can be represented in normal

form (where ak are from the right) by means of finite sum:

h =∑mn

∑k1,...,km,l1,...,ln

Γk1,...,km,l1,...,lna∗k1 . . . a

∗kmal1 . . . aln . (I.18)

Here, Γk1,...,km,l1,...,ln = Γln,...,l1,km,...,k1 (this condition guarantees that

h is self-adjoint).

In the case of infinite number of generators, one can modify this

construction to obtain other infinitesimal automorphisms. Namely,

we can consider (I.18) as a formal expression; if the expressions

[ak, h], [a∗k, h] can be regarded as elements ofA, these formulas specify

a derivation of algebra. This happens if for every k, there exists

only a finite number of summands in (I.18) where one of the indices

is equal to k. (Then only a finite number of terms survives in the

commutator.) To check that the derivation specifies a one-parameter

family of automorphism, we should verify that the equations of

motion idakdt = [ak, h], ia∗kdt = [a∗k, h] have a solution. If h is a quadratic

hamiltonian, the equations of motion are linear and the solution is

the same as in classical theory. If h is a sum of quadratic Hamiltonian

and a summand multiplied by a parameter g, then it is easy to

prove that the equations of motion can be solved in the framework

of perturbation theory with respect to g.

Page 39: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xxxviii

xxxviii Mathematical Foundations of Quantum Field Theory

For example, we can take h =∑εka∗kak. Then we obtain

an infinitesimal automorphism leading to well-defined equations of

motion:

idakdt

= −εkak, ida∗kdt

= εka∗k.

If εk ≥ 0, the ground state representation is in this case the

Fock representation (the representation containing a cyclic vector

θ obeying akθ = 0). To prove this fact, we note that in the space

of Fock representation (Fock space), we have an orthogonal basis

of eigenvectors (a∗1)n1 . . . (a∗k)nk . . . θ with eigenvalues

∑εknk. Here,

nk = 0, 1, 2, . . . in the case of CCR and nk = 0, 1 in the case of CAR,

only finite number of nk does not vanish.

If εk < 0, for some k, then in the case of CCR, the ground state

does not exist. Let us consider the case of CAR. Let us denote by N

the set of indices k where εk < 0. If N is finite, one can find ground

state Φ in the same Fock space: it obeys the conditions akΦ = 0

for k /∈ N , and a∗kΦ = 0 for k ∈ N. If N is infinite, we should

introduce a new system of generators satisfying the same relations

(make canonical transformation): bk = ak, b∗k = a∗k for k /∈ N , bk =

a∗k, b∗k = ak for k ∈ N . Then the ground state θ′ lies in the Fock

space constructed by means of bk, b∗k and obeys bkθ

′ = 0.

The above statements can be easily generalized in various

directions. In particular, one can work with algebras defined by

relations (I.15) or (I.17). Then the Fock representation can be defined

as a representation in a (pre)-Hilbert space F that contains a cyclic

vector θ obeying a(f)θ = 0. If B consists of functions on M, we

can write formally a(f) =∫f(k)a(k)dk (integration over M) and

consider a(k), a∗(k) as generalized functions on M.

The equations of motion

ida(k)

dt= −ε(k)a(k), i

da∗(k)

dt= ε(k)a∗(k) (I.19)

are coming from the formal Hamiltonian∫ε(k)a∗(k)a(k)dk.

Again in the case ε(k) ≥ 0, the Fock representation is the ground

state representation and the vector θ is the ground state. The formal

Hamiltonian becomes a self-adjoint operator in Fock space.

Page 40: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xxxix

Introduction xxxix

More generally, any quadratic Hamiltonian leads to linear classical

equations of motion that can also be considered as quantum equa-

tions of motion. This is true, for example, for Klein–Gordon equation

and for Dirac equation. To construct the ground state representation,

one should change variables to write the equations in the form (I.19)

where a, a∗ obey CCR or CAR.

If a formal Hamiltonian is represented as H0 + V where H0

is a quadratic Hamiltonian, then in good situations, not only H0

but also the full Hamiltonian is a well-defined self-adjoint operator

in the ground state representation of H0. If we started with a

translation-invariant Hamiltonian without ultraviolet divergences,

then usually we obtain a “good” Hamiltonian after the volume cutoff

(infrared cutoff). We can calculate Wightman functions of “good”

Hamiltonian (at least in the framework of perturbation theory).

Then taking the limit of Wightman functions as the volume tends to

infinity, we obtain Wightman functions of the original Hamiltonian

(these construction can be considered as a definition of Wightman

functions). Finally, using Wightman functions and an analog of GNS

construction, we can construct the ground state representation of

the original formal Hamiltonian and of Heisenberg operators in this

space. The Heisenberg operators satisfy the equations of motion

coming from the formal Hamiltonian. We say that we have obtained

an operator realization of formal Hamiltonian (see Section 11.3 for

more details).

Quantum field theory via the algebraic approach

Quantum field theory can be considered as a particular case of

quantum mechanics.

To define quantum field theory on d-dimensional space (on (d+ 1)-

dimensional space-time), we assume that the group of space-time

translations acts on the algebra of observables. We can say that quan-

tum field theory is quantum mechanics with the action of commutative

Lie group on the algebra of observables A. In other words, we assume

that operators α(x, t) where x ∈ Rd, t ∈ R are automorphisms of Apreserving the involution and that α(x, t)α(x′, t′) = α(x + x′, t + t′).

We will use the notation A(x, t) for α(x, t)A where A ∈ A.

Page 41: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xl

xl Mathematical Foundations of Quantum Field Theory

The action of translation group on A induces the action of this

group on the space of states.

Let us now consider a state ω that is invariant with respect to

translation group. We will define (quasi-)particles as “elementary

excitations” of ω. (If ω is the ground state, one uses the word

“particles”; if ω is an equilibrium state, we are talking about thermal

quasiparticles.) To consider collisions of (quasi-)particles, we require

that ω satisfies the cluster property in some sense. Instead of cluster

property, one can impose a condition of asymptotic commutativity

of the algebra A(ω) that consists of operators A corresponding to

the elements A ∈ A in GNS construction. In other words, one can

require that the commutator [A(x, t), B] where A,B ∈ A is small for

x→∞.

The weakest form of cluster property is the following condition:

ω(A(x, t)B) = ω(A)ω(B) + ρ(x, t), (I.20)

where A,B ∈ A and ρ in some sense is small for x→∞. For example,

we can impose the condition that∫|ρ(x, t)|dx < c(t) where c(t)

has at most polynomial growth. Note that (I.20) implies asymptotic

commutativity in some sense: ω([A(x, t), B]) is small for x→∞.

To formulate more general cluster property, we introduce the

notion of correlation functions in the state ω :

wn(x1, t1, . . .xn, tn) = ω(A1(x1, t1) · · ·An(xn, tn)),

where Ai ∈ A. They generalize Wightman functions of relativistic

quantum field theory. We consider corresponding truncated correla-

tion functions wTn (x1, t1, . . .xn, tn) (see Section 10.1 for definition).

We have assumed that the state ω is translation-invariant; it

follows that both correlation functions and truncated correlation

functions depend on differences xi − xj , ti − tj . We say that the

state ω has cluster property if the truncated correlation functions

are small for xi − xj → ∞. A strong version of cluster property is

the assumption that the truncated correlation function tends to zero

faster than any power of ||xi − xj ||. Then its Fourier transform with

respect to variables xi has the form νn(p2, . . . ,pn, t1, . . . , tn)δ(p1 +

· · ·+ pn) where the function νn is smooth.

Page 42: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xli

Introduction xli

Let us show how one can define one-particle excitations of the

state ω and the scattering of (quasi-)particles.

The action of translation group on A generates unitary represen-

tation of this group on the (pre-)Hilbert space H constructed from

ω. Generators of this representation P and −H are identified with

momentum operator and Hamiltonian. The vector in the space Hthat corresponds to ω will be denoted by Φ. If Φ is a ground state,

we say that it is the physical vacuum. If ω obeys KMS-condition

ω(A(t)B) = ω(BA(t+ iβ)), (I.21)

we say that ω is an equilibrium state with the temperature T = 1β .

(It is assumed that A(t) can be analytically continued to the strip

0 ≤ Im t ≤ β.)

We say that a state σ is an excitation of ω if it coincides with ω

at infinity. More precisely, we should require that σ(A(x, t))→ ω(A)

as x → ∞ for every A ∈ A. Note that the state corresponding to

any vector AΦ where A ∈ A is an excitation of ω; this follows from

cluster property.

One can define one-particle state (one-particle excitation of the

state ω) as a generalized H-valued function Φ(p) obeying PΦ(p) =

pΦ(p), HΦ(p) = ε(p)Φ(p). (More precisely, for some class of test

functions f(p), we should have a linear map f → Φ(f) of this

class into H obeying PΦ(f) = Φ(pf), HΦ(f) = Φ(ε(p)f) where

ε(p) is a real-valued function. For definiteness, we can assume that

test functions belong to the Schwartz space S(Rd).) We require that

there exist an element B ∈ A such that BΦ = Φ(φ) where φ is a

non-vanishing function. (The symbol B denotes the operator in Hcorresponding to the element B ∈ A.) We also assume that there

exists an element A ∈ A and a function g(x, t) ∈ S(Rd+1) such that

B =∫g(x, t)A(x, t)dxdt (it follows from this assumption that B(x, t)

is a smooth function of x, t).

We assume that Φ(f) is normalized (i.e. 〈Φ(f),Φ(f ′)〉 = 〈f, f ′〉).Of course, it is possible that there are several one-particle states;

to simplify the notations, we assume that there exists only one kind

of particles.

Page 43: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xlii

xlii Mathematical Foundations of Quantum Field Theory

Note that in relativistic theory, one-particle states can be iden-

tified with irreducible representations of the Poincare group; our

assumption means that we consider a scalar particle.

Let us fix a function f(p). Define a function f(x, t) as the

Fourier transform of the function f(p)e−iε(p)t with respect to p.

Let us introduce the notation B(f, t) =∫f(x, t)B(x, t)dx. (We

assume that this expression specifies an element of A such that

B(f, t) =∫f(x, t)B(x, t)dx.) It is easy to check that

B(f, t)Φ =

∫f(p)φ(p)Φ(p)dp (I.22)

does not depend on t.

Let us consider the vectors

Ψ(f1, . . . , fn|t) = B(f1, t) · · ·B(fn, t)Φ, (I.23)

where f1, . . . , fn satisfy the following condition: if fi(p) 6= 0, fj(p′) 6=

0, i 6= j, then ∇ε(p) 6= ∇ε(p′). (More precisely, we should assume

that the distance between ∇ε(p) and ∇ε(p′) is bounded from below

by a positive number.) We impose the additional requirement that

ε(p) is a smooth strictly convex function. Then one can derive from

cluster property or from asymptotic commutativity of A(ω) that

these vectors have limits in H as t→ ±∞; these limits will be denoted

by Ψ(f1, . . . , fn| ±∞).

Let us assume the vectors Ψ(f1, . . . , fn|−∞) span a dense subset

of H. (One can hope that this condition is satisfied when ω is the

ground state. In other cases, one should consider inclusive cross-

sections and inclusive scattering matrix; see Chapter 13 and Schwarz

(2019c) for more detail.) Then we can define in-operators a+in, ain on

H. In particular, to define a+in(f) =

∫(f(p)a+

in(p)dp, we can use the

formula

a+in(f φ)Ψ(f1, . . . , fn| −∞) = Ψ(f, f1, . . . , fn| −∞),

The definition of out-operators is similar. Using these notions, we

can define the scattering amplitudes by the formula

Smn(p1, . . . ,pm|q1, . . . ,qn)

= 〈a+in(q1) . . . a+

in(qn)Φ, a+out(p1) . . . a+

out(pm)Φ〉.

Page 44: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xliii

Introduction xliii

Concluding remarks

In conclusion, a couple of general words about quantum theory.

A quantum mechanical system is specified by a Hamiltonian

(infinitesimal generator of time evolution) and a QFT is specified

by a Hamiltonian and a momentum operator (generator of spatial

translations). Knowing these data, we can define a notion of particle

in QFT; a particle can be interpreted as an elementary excitation of

ground state. Under certain conditions, we can define the scattering

of particles.

Note that string field theory can be regarded as a quantum field

theory in the sense of this definition.

Quantum mechanics is a deterministic theory. This means that

knowing the state at some moment and the Hamiltonian governing

the evolution of the state, we can in principle predict the state at

any moment — precisely as in classical mechanics. The probabilities

in quantum mechanics can be explained in the same way as in

classical statistical physics — they come from a random environment

(see Sections “Decoherence” and “Observables and probabilities” on

page xxxi for an explanation of how probabilities can be obtained

from decoherence).2

The notion of particle in quantum field theory is a secondary

notion. There exists no physical difference between elementary and

composite particles. (There are theories that can be represented in

two different ways: elementary particles of one approach or com-

posite particles of another approach.) Probably, the most revealing

illustration of properties of quantum particles is given by analogy

with nonlinear scattering in classical field theory. Such a theory can

have particle-like solutions, for example, solitons. (A soliton is a finite

energy solution of the form s(x−vt). It can be visualized as a bump

moving with constant speed that does not change its form. If we have

a bump moving with constant average speed, but the form of the

bump changes with time, we can talk about generalized soliton or, in

different terminology, about particle-like solution.) It seems that for

2We do not consider subtle questions of measurement theory. The abovestatement means only that the standard formulas for probabilities can be obtainedfrom decoherence.

Page 45: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xliv

xliv Mathematical Foundations of Quantum Field Theory

large class of theories, the asymptotic behavior of any finite energy

solution as t → ±∞ can be described as superposition of particle-

like solutions and almost linear “tail” (see, for example, Tiupkin

et al., 1975; Soffer, 2006; Tao, 2009). (For integrable theories with one

spatial dimension, this is a rigorous theorem; in higher dimensions,

we only have experimental confirmation.) This means, in particular,

that we can talk about scattering of solitons. Two solitons collide

and we see a field configuration that does not resemble any particle-

like solution. After that, solitons miraculously reappear (with the

same velocities in integrable theories, with different velocities in non-

integrable ones.) This is precisely the picture we can see considering

a collision of quantum particles. It seems that for a given asymptotic

behavior, one can rigorously construct a solution of equations of

motion having this behavior. (In quantum field theory, this was

done in Haag–Ruelle–Araki scattering theory in the relativistic case.

Non-relativistic generalization of this theory was given in Fateev

and Shvarts (1973); see also Chapters 10 and 13.) Both in classical

and quantum integrable theories in two-dimensional space-time,

one can analyze asymptotic behavior of any solution of equations

of motion and rigorously justify the above picture. This can be

done also in non-relativistic quantum mechanics: one can prove

that asymptotically every solution behaves as a superposition of

elementary and composite particles (bound states); see, for example,

Hunziker and Sigal (2000). It is not known whether this statement is

correct for more general theories (and it is definitely wrong for many

relativistic conformal field theories).

We see that scattering of solitons eerily resembles quantum scat-

tering. This similarity becomes even more complete if the classical

theory has topological integrals of motion (i.e. the space of finite

energy solutions is disconnected). Magnetic charge is an example of

topological integral of motion, the solitons with minimal (by absolute

value) magnetic charge are magnetic monopoles (if the charge is

positive) or antimonopoles (if the charge is negative). Monopoles can

annihilate with antimonopoles.

Quantum mechanics is a consistent theory that does not need an

interpretation in the framework of classical field theory. I do not think

that quantum scattering can be interpreted as classical scattering

Page 46: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xlv

Introduction xlv

of solitons in random environment (although this possibility cannot

be excluded). My point is that nonlinear classical scattering is very

similar to quantum scattering. This idea is supported by remark that

in the limit ~ → 0, the scattering of solitons can be obtained from

the scattering of quantum particles.

Notation

A ∪B the union of sets A and B;

A ∩B the intersection of sets A and B;

A×B the cross product of sets A and B (the set of all (a, b)

with a ∈ A and b ∈ B);

En the n-dimensional Euclidean space;

L2(X) the Hilbert space of square-integrable complex

functions on the measure space X;

〈x, y〉 the scalar multiplication of elements of (pre-)Hilbert

space (for Euclidean space elements x,y ∈ En, this

quantity will sometimes be denoted xy);

S(En) the space of smooth functions of n variables that

rapidly decay; by smooth, we mean infinitely

differentiable and by rapidly decaying, we mean

faster than any power;

Ω the cube with length L edges in E3, satisfying the

equations 0 ≤ x ≤ L, 0 ≤ y ≤ L, 0 ≤ z ≤ L;

TΩ a set of vectors with k = 2πnL , where n is an integer

vector;

φk the orthonormal basis in the space L2(Ω), formed by

the functions φk(x) = eixk;

H1 +H2 a direct sum of Hilbert spaces;

H1 + · · ·+Hn + · · ·

a direct sum of an infinite sequence of Hilbert spaces;

F (H) the Fock space built with the Hilbert space H (if

H = L2(X), then F (L2(X)) = F0 + F1 + · · ·+Fn + · · · , where Fn consists of square-integrable

symmetric functions in n variables in the case of

bosons; in the case of fermions, the functions are

antisymmetric);

Page 47: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-00-fm page xlvi

xlvi Mathematical Foundations of Quantum Field Theory

[A,B] =

AB −BAthe commutator of A and B;

[A,B]+ =

AB +BA

the anticommutator of A and B;

Two operators A and B on a Hilbert space are (Hermitian)

conjugate to each other if 〈Ax, y〉 = 〈x,By〉 for all x, y;

A∗ (the adjoint operator) the conjugate operator to A

with the maximal domain of definition;

A+ the conjugate operator to A where the domain of

definition coincides with A (for operators defined

on the whole Hilbert space A∗ = A+; in this case,

we will use A∗);

slimAn the strong limit of a sequence of operators An (it is

required that all the operators have the same

domain). One says that A = slimAn if

Ax = limAnx for all x in the domain D;

wlimAn the weak limit of a sequence of operators An. One

says that A = wlimAn if 〈Ax, y〉 = lim〈Anx, y〉 for

all x, y in the domain D;

δk,l = δkl the Kronecker symbol (δk,l = 0 if k 6= l, else it is 1);

δ(ξ, η) delta function (here, ξ and η belong to a measure

space); if ξ, η ∈ En, then δ(ξ, η) = δ(ξ − η);

θ(t) = 0 the Heaviside step function; θ(t) = 0 if t < 0, and

θ(t) = 1 with t ≥ 0;1

w+i0 the formula limε→0+1

w+iε = 1i

∫θ(t) exp(iwt)dt;

the d’Alembert operator∂2

∂t2− ∂2

∂x2− ∂2

∂y2− ∂2

∂z2= ∂2

∂t2−∆;

supp f the support of the function f defined as the closure

of the set of arguments where f is non-zero.

For the momentum variables, we use the notations k, p, l, q; for

the energy variable, we usually use ω. For coordinates, we use x, ξ,

y, η, and time is denoted by t or τ .

We use the system of units where the Planck constant ~ and the

speed of light c are equal to 1.

Page 48: Mathematical Foundations of Quantum Field Theory

March 26, 2020 11:53 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch01 page 1

Chapter 1

Principles of Quantum Theory

1.1 Principles of quantum mechanics

Almost all modern physics theories begin with the following scheme.

First, one specifies a mathematical object ψt that describes the

state of an observed physical system at time t. Second, one obtains

equations that govern the object’s change with time (usually taking

the form ∂ψt

∂t = A(ψt)). Finally, one defines ways to calculate

experimentally observed quantities from the object ψt.

Accordingly, we will formulate the postulates of quantum mechan-

ics in this manner.

The state of a quantum system at a fixed moment in time is

specified by a non-zero vector ψ taken from a complex Hilbert space

R; two vectors ψ and ψ′ = Cψ, differing only by a non-zero scalar

factor C, represent the same state. By choosing C = ‖ψ‖−1, we can

normalize the vector ψ′ and hence a state can be represented as a

normalized vector (a vector with norm equal to 1). If we don’t say

otherwise, we consider quantum states to be normalized vectors in

the rest of the book.

1.2 Evolution of state vectors

Knowing the state of a quantum system at time t0 (denoted by ψt0),

one can calculate the state ψt for any moment in time t, that is,

there exists a linear operator U(t, t0), the evolution operator, that

transforms the state vector ψt0 into the vector ψt.

The evolution operator U(t, t0) satisfies the group relations

U(t2, t1)·U(t1, t0) = U(t2, t0), U(t0, t0) = 1. It obeys the Schrodinger

1

Page 49: Mathematical Foundations of Quantum Field Theory

March 26, 2020 11:53 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch01 page 2

2 Mathematical Foundations of Quantum Field Theory

equation:

i∂U(t, t0)

∂t= H(t)U(t, t0), (1.1)

with the initial condition U(t0, t0) = 1. Here, H(t) is a self-adjoint

operator called the Hamiltonian operator of the quantum system.

The Schrodinger equation can also be written in the form

idψtdt

= H(t)ψt, (1.2)

though less precisely, since the operator H(t) may not be defined on

the full Hilbert space R, while U(t, t0) is defined everywhere.

It is easy to check that U∗(t, t0) = U(t0, t); using the group

property, we can see that the operator is unitary (to be rigorous, we

should assume that the evolution operator is unitary and prove that

the Hamiltonian is self-adjoint using the unitarity of the evolution

operator).

If the Hamiltonian does not depend on time explicitly, we can

write U(t, t0) in the form U(t, t0) = exp(−iH(t− t0)).For a non-relativistic one-dimensional particle, the state space R

can be taken to be L2(E1), the space of square-integrable functions

ψ(x) of one variable x, where −∞ < x < ∞. The Hamiltonian

operator in this case can be written as H(t) = − 12m

d2

dx2+ V(x, t)

where V(x, t) is the multiplication operator by the function V(x, t)

(t plays the role of a parameter here).1 This Hamiltonian describes

the non-relativistic particle with mass m, moving in the field with

time-dependent potential V(x, t).

1.3 Calculating the probabilities

An observable in quantum mechanics corresponds to a self-adjoint

operator A on the space R. We will consider the correspondence

between classical and quantum observables later in the book.

1It is easy to check that V(x, t) = V(x, t), where x is the operator ofmultiplication by x.

Page 50: Mathematical Foundations of Quantum Field Theory

March 26, 2020 11:53 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch01 page 3

Principles of Quantum Theory 3

An important general remark: the energy corresponds to the

Hamiltonian operator H.

If we know the operator A corresponding to the observable a and

a state of the quantum system, written as a normalized vector ψ ∈ R,

we can obtain the probability distribution of the values obtained by

measuring the observable a in the state ψ. Let us denote by µ(α)

the probability that after measurement we obtain the value a to

be less than or equal to α. In quantum mechanics, we postulate that

µ(α) = 〈eα(A)ψ,ψ〉, where eα(λ) is a function equal to 1 when λ ≤ αand 0 otherwise.

This postulate can be formulated another way: for any function

f , the mean value f(a) =∫f(α)dµ(α) of f(a) in the state ψ is given

by the formula f(a) = 〈f(A)ψ,ψ〉. (The probability µ(a) is equal to

the mean value of eα(a), hence the first form of the postulate follows

from the second form. To derive the second form from the first one,

we note that any function f can be represented as a limit of linear

combinations of the functions eα.)

If the observables a1, . . . , an correspond to commuting operators

A1, . . . , An, then a1, . . . , an are simultaneously measurable. This

means that for any state ψ we can find the joint probability

distribution for the values obtained when measuring a1, . . . , ansimultaneously. Namely, let us denote by µ(α1, . . . , αn) the prob-

ability that in the state ψ we obtain a1 ≤ α1, . . . , an ≤ αn. We

postulate this probability to be 〈eα1,...,αn(A1, . . . , An)ψ,ψ〉 where

eα1,...,αn(λ1, . . . , λn) is a function equal to 1 when λ1 ≤ α1, . . . , λn ≤αn and 0 otherwise.

As in the case of a single observable, the above postulate can

be reformulated by demanding that for any function f(a1, . . . , an)

the mean value f(a1, . . . , an) of f(a1, . . . , an) in the state ψ can be

obtained by the formula f(a1, . . . , an) = 〈f(A1, . . . , An)ψ,ψ〉.For a single one-dimensional particle, the position observable x

(the x-coordinate of the particle) corresponds to the multiplication

operator xψ = xψ. The momentum of the particle is represented by

the operator p = 1iddx , with the operator p2

2m = − 12m

d2

dx2corresponding

to the kinetic energy. The operator corresponding to the potential

energy is the multiplication operator by V(x, t). The momentum

Page 51: Mathematical Foundations of Quantum Field Theory

March 26, 2020 11:53 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch01 page 4

4 Mathematical Foundations of Quantum Field Theory

and the kinetic energy are simultaneously measurable and so are

the position and the potential energy. However, it is impossible to

measure the position and the momentum simultaneously.

If A is a self-adjoint operator with discrete spectrum, then there

exists an orthonormal basis of eigenvectors ri in R for the operator

A with corresponding eigenvalues ai.

In the basis ri, the matrices for the operators A and eλ(A) are

diagonal, with the diagonal elements equal to ai for the operator A

and equal to eλ(ai) for the operator eλ(A). Let us suppose that a

normalized state vector is represented in the basis ri as ψ =∑

i ciri,

then

〈eλ(A)ψ,ψ〉 =∑ij

cic∗jeλ(ai)〈ri, rj〉 =

∑i

eλ(ai)|ci|2.

It follows that when measuring the physical quantity correspond-

ing to the operator A, we obtain only the eigenvalues ai with non-zero

probability. The probability of obtaining the value a in the state ψ

is equal to∑

i |〈ψ, ri〉|2, where the sum is taken over those i where

Ari = ari.

To each state vector ψ, we can assign the sequence (c1, . . . , ci, . . .)

of coefficients that arise from the decomposition of the vector ψ with

respect to the basis ri, consisting of the eigenvectors of operator A.

This construction specifies an isomorphism of the space R and the

space l2. We call this isomorphism an A-representation and the

sequence (c1, . . . , ci, . . .), we call the A-representation of the vector ψ.

The A-representation of the vector ψ allows us to easily calculate the

probabilities of values of the observable a in the state ψ.

The notion of an A-representation can also be defined when the

operator A has continuous spectrum. Moreover, it can be generalized

to the case when we are dealing with a family of commutative self-

adjoint operators. Namely, if A1, . . . , Ak are self-adjoint, pairwise

commuting operators, then an (A1, . . . , Ak)-representation is an

isomorphism of Hilbert spaces R and L2(M) in which the operators

A1, . . . , Ak are transformed to the multiplication operators by the

functions a1(m), . . . , ak(m) (hereM is a measure space, withm∈M).

In the (A1, . . . , Ak)-representation of the vector ψ ∈ R, one can easily

Page 52: Mathematical Foundations of Quantum Field Theory

March 26, 2020 11:53 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch01 page 5

Principles of Quantum Theory 5

calculate the probability distribution of the observables A1, . . . , Akin the state ψ. It is also possible to prove that the (A1, . . . , Ak)-

presentation exists for any system of commuting self-adjoint opera-

tors A1, . . . , Ak. (If the operators A1, . . . , Ak have discrete spectrum,

then there exists an orthonormal basis of the space R consisting of

the common eigenfunctions of the operators A1, . . . , Ak).

The (A1, . . . , Ak)-presentation is closely tied to the generalized

eigenfunctions of the operators A1, . . . , Ak. That is, let rm, where m

belongs to the space with measure M , be a δ-normalized, generalized

basis for the space R (for details, see A.8). If the basis is an

eigenbasis for the operators A1, . . . , Ak (i.e. Airm = ai(m)rm), then

the isomorphism of the spaces R and L2(M), given by the formula

a(m) = 〈x, rm〉, is a (A1, . . . , Ak)-presentation.

In what follows, we will identify physical quantities with their

corresponding self-adjoint operators.

1.4 Heisenberg operators

Up to this point, we have considered our state vectors to be

dependent on time, while the operators for the physical quantities

were constant (Schrodinger’s picture). However, it is possible to view

this from a different but equivalent angle, where the operators depend

on time, but the vectors stay constant (Heisenberg’s picture).

Let us define for every operator A a time-dependent operator At(a Heisenberg operator) by the formula At = U∗(t, 0)AU(t, 0).

We note the following relations:

(AB)t = U∗(t, 0)ABU(t, 0)

= U∗(t, 0)AU(t, )U∗(t, 0)BU(t, 0)

= AtBt,

(A∗)t = U∗(t, 0)A∗U(t, 0)

= (U∗(t, 0)AU(t, 0))∗

= (At)∗,

f(At) = f(A)t,

〈Atψ,ψ〉 = 〈Aψt, ψt〉.

Page 53: Mathematical Foundations of Quantum Field Theory

March 26, 2020 11:53 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch01 page 6

6 Mathematical Foundations of Quantum Field Theory

Combining the last two relations, we can verify that calculating

the probability distributions for the operator A and the state

ψt = U(t, 0)ψ, we obtain the same answer as when we calculate the

probability distributions for the operator At and the state ψ. Hence,

the Heisenberg and Schrodinger pictures are equivalent.

From (1.1) and the Hermitian conjugate equation i∂U∗(t,0)∂t =

−U∗(t, 0)H(t), we obtain the following equation for the operator At(the Heisenberg equation of motion):

dAtdt

= i[Ht, At],

where Ht = U∗(t, 0)H(t)U(t, 0). When the Hamiltonian H does

not depend on time, Ht = H,At = exp(itH)A exp(−itH), and the

Heisenberg equation takes the form

dAtdt

= i[H,At].

Let us calculate the Heisenberg equation of motion in a concrete

example. Let R be the space of square-integrable functions of one

variable x and H = p2

2m + V(x, t), where p = 1iddx and x is the

operator of multiplication by x. Using the fact that time evolution

commutes with summation and multiplication of operators, as well

as the operation of taking a function of an operator, we obtain

Ht =p2t2m

+ V(xt, t).

(In the case when H does not depend on time, then H = Ht.)

From the relation [p,V(x)] = 1iV′(x), it follows that [pt,

p2t2m +

V(xt, t)] = 1iV′(xt, t) and therefore dp

dt = −V ′(xt, t). (We are using the

fact that evolution preserves commutation relations.) Analogously,

we have that dxtdt = 1

m pt. Similar reasoning can be applied whenever it

is necessary to write the Heisenberg equations for a specific operator.

1.5 Integrals of motion and stationary states

A physical quantity (an observable) A is called an integral of motion

if for any state vector ψ, the probability distribution of the quantity

A in the state ψt = U(t, 0)ψ does not depend on the time t.

Page 54: Mathematical Foundations of Quantum Field Theory

March 26, 2020 11:53 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch01 page 7

Principles of Quantum Theory 7

In the Heisenberg picture, the quantity A is an integral of motion

when the Heisenberg operator At does not depend on t. We will

assume for the rest of this section that the Hamiltonian H does not

depend on time. It is clear from the Heisenberg equations that the

quantity A will be an integral of motion if and only if the operator

A commutes with the Hamiltonian H. In particular, the physical

quantity that corresponds to a time-independent operator H is an

integral of motion; as we have noted, an observable of this form has

the physical meaning of energy.

The state vectors corresponding to the eigenvectors of the oper-

ator H (states with definite energy E) are called stationary states.

Indeed, if Hψ = Eψ, then ψt = exp(−iEt)ψ satisfies the Schrodinger

equation. This means that ψt differs from ψ only by a numerical

factor and consequently describes the same state; hence, the solution

for the equation Hψ = Eψ (the stationary Schrodinger equation)

is “stationary”. Conversely, when the vector ψt = U(t, 0)ψ for any

t differs only by a factor C(t) from the vector ψ, it follows from

the Schrodinger equation that iC ′(t)ψ = C(t)Hψ, hence C(t) =

exp(−iEt) and Hψ = Eψ. Therefore, any state that doesn’t change

with time is stationary in the sense of the above definition.

A stationary state corresponding to the smallest energy value is

called the ground state. It is easy to check that ground states can

also be characterized as states with the smallest average energy; in

other words, if Φ0 is a ground state and Ψ is any other state, then

〈HΨ,Ψ〉 ≥ 〈HΦ0,Φ0〉 (we assume that Φ0 and Ψ are normalized

vectors). Indeed, let us consider for simplicity the case when the

Hamiltonian H only has discrete spectrum. Decomposing the vector

Ψ into the eigenvectors Φn of the operator H, we have

〈HΨ,Ψ〉 =⟨H

∑cnΦn,

∑cnΦn

⟩=

∑En|cn|2

≥ E0

∑|cn|2

≥ E0 = 〈HΦ0,Φ0〉

(here, En are the eigenvalues of the operator H corresponding to the

eigenvectors Φn and E0 = minEn is the energy of the ground state).

Page 55: Mathematical Foundations of Quantum Field Theory

b2530 International Strategic Relations and China’s National Security: World at the Crossroads

b2530_FM.indd 6 01-Sep-16 11:03:06 AM

This page intentionally left blankThis page intentionally left blankThis page intentionally left blankThis page intentionally left blank

Page 56: Mathematical Foundations of Quantum Field Theory

March 26, 2020 12:16 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch02 page 9

Chapter 2

Quantum Mechanics of Single-Particleand Non-Identical Particle Systems

2.1 Quantum mechanics of a single scalar particle

To describe a quantum mechanical system, we specify the Hilbert

space R of quantum states and the Hamiltonian operator H defining

time evolution. Similarly, we need to specify the operators for the

major physical quantities. In the following, we describe how to write

simple quantum systems.

The state for a single scalar particle in three-dimensional space

is described by a wave function — a square-integrable function ψ(r)

(i.e. the Hilbert space R in this case is L2(E3)). The Hamiltonian

operator for a non-relativistic particle in a potential field with energy

V(r) is written in the form H = − 12m∆ + V(r), where m is the mass

of the particle, ∆ is the Laplace operator, and V(r) is the operator

of multiplication by the function V(r). The coordinate operator x

for the coordinate x is defined as the multiplication operator by x

(i.e. ψ(r) = xψ(r)). Operators y and z are defined analogously. Since

the operators x, y, z commute, we can obtain the joint probability

distribution for the coordinates x, y, z in the state ψ(r). From the

previously formulated postulates, it follows that the density of this

probability distribution is |ψ(r)|2 (as always, we consider the state

vectors ψ(r) to be normalized). The operators for projections of the

momentum p are px = 1i∂∂x , py = 1

i∂∂y , pz = 1

i∂∂z , respectively.

The components of angular momentum are operators defined by the

formulas: Mx = ypz − zpy; My = zpx − xpz; Mz = xpy − ypx.

9

Page 57: Mathematical Foundations of Quantum Field Theory

March 26, 2020 12:16 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch02 page 10

10 Mathematical Foundations of Quantum Field Theory

For brevity, we will often use the vector operators: r = (x, y, z),

p = (px, py, pz), M = r× p = (Mx, My, Mz).

For the Heisenberg operators rt and pt, the equations of motion

are

drtdt

=ptm

;dptdt

= −dVdr

(rt).

Up to this point, the states for one particle were written as ψ(r)

with r ∈ E3. If α is a unitary operator, transforming the space

L2(E3) to the space L2(M), where M is some measure space (an

isomorphism of the space L2(E3) and the space L2(M)), then we

can equivalently represent the state vectors ψ ∈ L2(E3) as vectors

ψ = αψ ∈ L(M). For every operator A acting on the space L2(E3),

there will be a corresponding operator A = αAα−1 acting on L2(M),

and from 〈f(A)ψ, ψ〉 = 〈f(A)ψ,ψ〉, it follows that the probabilities

obtained from the operator A and the state vector ψ will be the same

as those obtained from A and ψ.

Changing the representation of a state vector is useful in many

cases. For example, let α be the Fourier transform

ψ(p) = (2π)−3/2

∫exp(−ipr)ψ(r)dr

(in this case M = E3). The Fourier transform of the operator px =1i∂∂x is the multiplication operator ˜pxψ(p) = pxψ(p) (analogously for

the other variables). It is clear that the joint probability density of

px, py, pz is equal to |ψ(p)|2.

Let us describe the geometric origin of momentum and angular

momentum operators. Note that to every translation T in the space

E3, there corresponds a unitary operator WT on the space L2(E3),

that is, WT transforms wave functions when the underlying space

is transformed by T . WT can be written in the form WTψ(r) =

ψ(T−1r). It is obvious that WT1WT2 = WT1T2 and therefore the

operators WT form a unitary representation of the translation group.

The component operator for the momentum along an axis can

be written as the operator of infinitesimal translation along the

axis, that is, the operator px, for example, can be written as

i lima→0WTa−1

a , where Ta is translation by a along the axis x

Page 58: Mathematical Foundations of Quantum Field Theory

March 26, 2020 12:16 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch02 page 11

Single-Particle and Non-Identical Particle Systems 11

(Ta(x, y, z) = (x+a, y, z)). Accordingly, the component of the angular

momentum can be written as the operator of infinitesimal small

rotation around the corresponding axis, that is, Mz can be written

as i limψ→01ψ (WSψ − 1) where Sψ denotes rotation by the angle

ψ around the z axis (i.e. Sφ(x, y, z) = (x cosφ − y sinφ, x sinφ +

y cosφ, z)).

This geometric interpretation allows us to understand when

components of momentum and angular momentum are integrals of

motion. For example, if the Hamiltonian is invariant with respect to

rotation around the z axis (i.e. if H commutes with WSφ), then the

Hamiltonian will commute with the operator Mz and therefore Mz

will be an integral of motion.

2.2 Quantum mechanics of particles with spin

The state of a particle with spin is written as a column vector ψ of

k square-integrable functions defined on three-dimensional space

ψ =

ψ1(r)

ψ2(r)

...

ψk(r)

,or, in other words, a function ψ(r) with k-column outputs. Instead

of notation ψi(r), we will use ψ(r, i), where i is a parameter, that

takes the values 1, 2, . . . , k. In other words, the space of states R can

be taken to be the space of square-integrable functions ψ(ξ), where

ξ = (r, i) ∈ E3 ×B, r ∈ E3, i ∈ B = 1, . . . , k.Integration over ξ has the same effect as integrating over r and

summing over i, for example, taking two functions φ ∈ R and ψ ∈ R,

we have

〈φ, ψ〉 =

∫φ(ξ)ψ(ξ)dξ =

k∑i=1

∫φ(r, i)ψ(r, i)dr.

The operators of position, momentum, and the Hamiltonian, if

the particle is in a potential field, are the same as in the case of the

Page 59: Mathematical Foundations of Quantum Field Theory

March 26, 2020 12:16 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch02 page 12

12 Mathematical Foundations of Quantum Field Theory

spinless particle. The discrete variable acts like a parameter under

the action of the listed operators on the function ψ(ξ) = ψ(r, i).

However, the operators of angular momentum are different. As we

have seen, these operators are related to the action of rotations on the

wave functions and, in general, the rotations act not only on space

variables but also on discrete variables. Therefore, these operators

are represented as sums of the orbital part and the part coming from

discrete variables (the spin part). More precisely, the components of

operators for the angular momentum Mx, My, Mz can be written

in the form Mx = lx + sx; My = ly + sy; Mz = lz + sz, where

l = (lx, ly, lz) is the operator of the orbital angular momentum and

s = (sx, sy, sz) is the operator for the spin angular momentum. The

operator for the orbital angular momentum obeys the same formulas

as the operator of angular momentum for spinless particles (i.e.

l = r × p); the operator for the spin angular momentum acts only

on the discrete variable (we say that the operator A acts only on the

discrete variable if there exist matrices aij , where 1 ≤ i, j ≤ k, such

that Aψ(r, i) =∑k

j=1 aijψ(r, j)).

Remark. We can verify these relationships by following the con-

siderations in Section 2.1. To every rotation T in three-dimensional

space, there corresponds a unitary transformation WT of the space

R, this transformation changes the argument r of the wave function

(as in scalar case), but we should take into account discrete variables.

For example, if k = 3, we can consider ψ(r) as a function of r, taking

values in three-dimensional vectors. Then the operator WT acts on

the function ψ(r) in the natural way, that is, the function ψ′ = WTψ

is given by the equation ψ′i(r) =∑3

j=1 Tijψj(T−1r), where Tij is the

matrix for the transformation T . In general, the operator WT should

be written in the form

(WTψ)(r, i) =

k∑j=1

DTijψ(T−1r, j).

This form of the operator WT immediately leads to the formulas

previously stated for the components of angular momentum, which

are identified with operators of infinitesimally small rotation around

Page 60: Mathematical Foundations of Quantum Field Theory

March 26, 2020 12:16 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch02 page 13

Single-Particle and Non-Identical Particle Systems 13

an axis. Operators WT must satisfy the relation WT1T2 = WT1WT2 ;

this means that they specify a unitary representation of the rotation

group. It is clear that matrices DT define k-dimensional unitary

representation of the rotation group. Knowing the finite-dimensional

unitary representation of the rotation group, it’s not difficult to

establish the form of spin projection operators. The correspondence

between rotation T with the operator WT and the matrices DT is in

general two-valued (the operator WT and the matrix DT are defined

up to a sign). This does not affect our calculations since state vectors

in quantum mechanics are defined up to multiplication by a number.

It is well known that for every value k ≥ 0, there exists one (up to

equivalence) k-dimensional unitary irreducible representation of the

rotation group. For k odd, the representation is single-valued; for k

even, it is two-valued. The functions

ψ(r) =

ψ1(r)

ψ2(r)

...

ψk(r)

,taking values in the state space, describe a particle with spin s = k−1

2 .

This terminology comes from the remark that the projection of the

spin takes values between −s and s (more precisely, every operator

sx, sy, sz, constructed as described in this paragraph, has eigenvalues

−s,−s+ 1, . . . , s− 1, s).

Especially important is the case of particles with spin 1/2, due to

many elementary particles, such as the electron, proton, and neutron,

having spin 1/2. Components of the spin can be written in this case

as matrices

sx =1

2

[0 1

1 0

]; sy =

1

2

[0 −ii 0

]; sz =

1

2

[1 0

0 −1

].

The above apparatus is also useful when particles have various other

quantum numbers, not just spin (e.g. isospin). The only difference

is in that when there are no quantum numbers besides spin, the

matrices DT specify an irreducible representation; otherwise, they

define a reducible representation.

Page 61: Mathematical Foundations of Quantum Field Theory

March 26, 2020 12:16 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch02 page 14

14 Mathematical Foundations of Quantum Field Theory

2.3 Quantum description of a system with

non-identical particles

Two electrons (or two protons) have the same mass, the same charge;

their other properties are also identical. Therefore, we say that

all electrons are identical particles. If two particles belonging to a

system of particles are identical, then we must essentially change the

quantum mechanical description of the system. In this section, we

will consider only a system of particles where any two particles are

not identical.

If we consider a system of n three-dimensional particles with spin

zero, then the Hilbert space R of the states of such a system can

be considered as a space L2(E3n) of square-integrable functions of n

spatial variables.

Let us consider the case of a system with n particles with spin.

We assume that the jth particle can be described by a function ψ(ξ),

defined on the set Bj (in coordinate representation Bj is the set

of pairs (r, i), where r is a point in three-dimensional space and i

runs over the discrete set 1, 2, . . . , kj). Then the space of states of

the whole system is the space R = L2(B1 × · · · × Bn) of square-

integrable functions ψ(ξ1, . . . , ξn) of variables ξ1, . . . , ξn that run

over the sets B1, . . . , Bn. In the coordinate representation, the wave

function ψ(x1, y1, z1, i1, . . . , xn, yn, zn, in) of the system of particles

depends on the coordinates (xj , yj , zj) and spin variables ij of every

particle.

Every operator corresponding to a physical quantity related to the

jth particle can be considered in a natural way as an operator on the

space R. For example, in the coordinate representation, the operators

xj , yj , zj of the coordinates of the jth particle act in R as operators

of multiplication by xj , yj , zj ; the operators pjx, pjy, pjz of the

momentum components of the jth particle act in R as operators of

coordinate differentiation of the jth particle: pjx = 1i∂∂xj

, pjy = 1i∂∂yj

,

pjz = 1i∂∂zj

(which we can write more compactly as pj = 1i∂∂rj

).

In general, if the operator A (acting on the space L(Bj)) trans-

forms the function ψ(ξ) into the function ψ′(ξ) =∫A(ξ, η)ψ(η)dη,

then the corresponding operator in the space R transforms the

Page 62: Mathematical Foundations of Quantum Field Theory

March 26, 2020 12:16 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch02 page 15

Single-Particle and Non-Identical Particle Systems 15

function ψ(ξ1, . . . , ξn) into the function

ψ′(ξ1, . . . , ξn) =

∫A(ξj , η)ψ(ξ1, . . . , ξj−1, η, ξj+1, . . . , ξn)dη.

The operator of total momentum p is given by the formula

p =n∑j=1

pj =n∑j=1

1

i

∂rj. (2.1)

The Hamiltonian operator of a system of n particles, in the

simplest case, can be written in the form

H =

n∑j=1

p2j

2mj+ V(r1, . . . , rn) =

n∑j=1

(− ∆j

2mj) + V(r1, . . . , rn) (2.2)

(here, the function V typically takes the form V(r1, . . . , rn) =∑nj=1 Vj(rj) +

∑j<l Vjl(|rj − rl|)).

If V(r1, . . . , rn) = V1(r1) + · · · + Vn(rn) (the case of non-

interacting particles in an external field), then the problem of the

motion of n particles can be reduced to n single-particle problems.

Let us calculate, as an example, the stationary states of n non-

interacting particles. Let us denote by ψ(j)1 (ξj), . . . , ψ

(j)r (ξj), . . . the

eigenfunctions of the operator Hj =p2j

2mj+ Vj(rj) in the space

L2(Bj) (the stationary states of the jth particle). The corresponding

eigenvalues will be denoted by E(j)r . It is easy to check that

the function φ(1)r1 (ξ1)φ

(2)r2 (ξ2) . . . φ

(n)rn (ξn) is a stationary state of the

Hamiltonian H = H1+· · ·+Hn in the space R with the corresponding

eigenvalue E(1)r1 + · · · + E

(n)rn . If, for every j, the functions φ

(j)r (ξj)

form a complete orthonormal system in the space L2(Bj), then the

products φ(1)r1 (ξ1)φ

(2)r2 (ξ2) . . . φ

(n)rn (ξn) form a complete orthonormal

system in the space R.

Until now, we have worked with coordinate representations, which

means that we considered wave functions as functions of coordinates

and spin variables. Let us use another representation for every

particle; let us choose an isomorphism αj of the space L2(Bj) and

the space L2(Mj), where Mj is a measure space. Then it is easy to

construct an isomorphism of the space R = L2(B1×· · ·×Bn) and the

Page 63: Mathematical Foundations of Quantum Field Theory

March 26, 2020 12:16 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch02 page 16

16 Mathematical Foundations of Quantum Field Theory

space L2(M1×· · ·×Mn) of square-integrable functions of n variables

(m1, . . . ,mn) that run over the sets M1, . . . ,Mn. In other words,

the state of a system of n particles can be represented by functions

ψ(m1, . . . ,mn), where mj ∈Mj . Denoting the space of states of the

j-th particle by Rj , we can say that the space of states R of an n

particle system is isomorphic to the tensor product R1 ⊗ · · · ⊗ Rn(this follows immediately from the definition of the tensor product

given in A.4).

In particular, it is very often convenient to use the momentum

representation. In the momentum representation, the system of n

particles is described as a function of momentum variables p1, . . . ,pnand spin variables i1, . . . , in, given by the formula

ψ(p1, i1, . . . ,pn, in) = (2π)−3n/2

∫exp

(−i

n∑k=1

pkrk

)×ψ(r1, i1, . . . , rn, in)dnr,

where ψ(r1, i1, . . . , rn, in) is the wave function in the coordinate

representation.

The momentum operator of the ith particle pi in the momentum

representation can be considered as the operator of multiplication by

pi; the operator of total momentum, which is equal to p =∑

pi, can

be considered as an operator of multiplication by p1 + · · ·+ pn.

2.4 A particle in a box with periodic

boundary conditions

Let us consider, as an example, a scalar particle moving freely in

the cube Ω, specified by the inequalities 0 ≤ x ≤ L, 0 ≤ y ≤L, 0 ≤ z ≤ L. To make this problem precise, we will impose periodic

boundary conditions on the wave functions of stationary states. The

space of states R in this case can be identified with the space of all

square-integrable functions on the cube Ω and the Hamiltonian is

the operator H = (− 12m)∆ with periodic boundary conditions (more

precise definition of this Hamiltonian will be given later).

Page 64: Mathematical Foundations of Quantum Field Theory

March 26, 2020 12:16 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch02 page 17

Single-Particle and Non-Identical Particle Systems 17

Let us fix a vector a in three-dimensional space and consider the

operation of parallel translation by the vector a. The corresponding

unitary shift operator Wa on the space R transforms state vectors as

translation (Waψ)(r) = ψ(r− a), where ψ(r) is a periodic extension

of ψ(r), given on the cube Ω, to the whole three-dimensional

space. The momentum operators px, py, pz can be defined as

infinitesimal shift operators along the axes x, y, z (for example,

px = i lima→0W(a,0,0)−1

a ).

It is easy to check that pxψ = 1i∂ψ∂x . The operator px is defined

on the functions ψ ∈ L2(Ω), obeying the condition ∂ψ∂x ∈ L

2(Ω) and

satisfying the periodic boundary conditions ψ(0, y, z) = ψ(L, y, z).

In terms of the operator p, the Hamiltonian can be written in the

form H = 12m p2.

The functions φk = L−3/2 exp(−ikr), where k runs over the

lattice TΩ consisting of vectors 2πL n (where n is a vector of integer

coordinates), form a complete orthonormal system of common

eigenfunctions of the operators H, px, py, pz. Every function ψ ∈ Rcan be represented in the form ψ(r) =

∑ckφk (where the sum is

taken over the lattice TΩ); the function ck on the lattice TΩ is called

the wave function in the momentum representation.

2.5 One-dimensional harmonic oscillator

Let us consider a one-dimensional non-relativistic quantum particle

with mass m and potential energy V (q). We assume that V (q) has

one global minimum at the point q0 and V ′′(q0)) = k > 0. Then we

can study the ground state and low-lying energy levels replacing V (q)

with V (q0)+ k2 (q−q0)2. This means that we should study a quantum

harmonic oscillator. Taking into account higher order terms in the

Taylor expansion of V (q), we obtain the Hamiltonian of anharmonic

oscillator. Without loss of generality, we can assume that m = 1,

k = 1, V (q0) = 0. Then in the variable x = q − q0, the Hamiltonian

of the harmonic oscillator takes the form

H =p2

2+x2

2= −1

2

d

dx2+

1

2x2.

Page 65: Mathematical Foundations of Quantum Field Theory

March 26, 2020 12:16 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch02 page 18

18 Mathematical Foundations of Quantum Field Theory

First, let us find the Heisenberg operators pt and xt from the

Heisenberg equations

dptdt

= −xt;dxtdt

= pt.

This system of operator equations is linear and therefore it can be

solved precisely as the corresponding system of numerical equations.

One of the possible ways to solve it is based on the introduction of

auxiliary operators

a =1√2

(x+ ip); a+ =1√2

(x− ip).

The equations for the Heisenberg operators at and a+t have a very

simple form

datdt

= −iat;da+

t

dt= ia+

t ,

hence a+t = a+ exp(it); at = a exp(−it). Using these relations, we

can immediately obtain the formulae xt = 1√2(at + a+

t ) and pt =1√2(a+t − at).

The operators a+ and a are very convenient for solving problems

related to the harmonic oscillator.

Note, first of all, that the Hamiltonian can be expressed in terms

of these operators by the formula H = a+a + 1/2. Second, note

that the commutation relations with the Hamiltonian have the form

[H, a+] = a+, [H, a] = −a, and their commutator is [a, a+] = 1.

Let us find the stationary states of the Hamiltonian H. We use

the following statement: If Hφ = Eφ, then H(aφ) = (E−1)aφ (hence

if φ is a stationary state and aφ 6= 0, then aφ is also a stationary

state). This statement follows from the relations H(aφ) = a(H −1)φ = (E − 1)aφ.

Similarly, H(a+φ) = a+(H + 1)φ = (E + 1)a+φ.

Note that the ground state φ0 should satisfy the relation aφ0 = 0

(otherwise, aφ0 is a stationary state with lower energy). Solving

the equation aφ0 = 1√2(x + d

dx)φ0 = 0, we obtain that φ0 =

π−1/4 exp(−12x

2) (the constant π−1/4 comes from the normalization

Page 66: Mathematical Foundations of Quantum Field Theory

March 26, 2020 12:16 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch02 page 19

Single-Particle and Non-Identical Particle Systems 19

condition) and that the corresponding eigenvalue is E0 = 1/2.

Starting with φ0, we can get an infinite number of stationary states

φn = cn(a+)nφ0 = cn2−n/2(x− ddx)n(π−1/4 exp(−x2

2 )) (here cn is the

normalization constant). Noting that the operator a+ increases the

energy by 1, we have Hφn = (n + 12)φn. It is easy to see that the

functions φn have the form φn(x) = Hn(x) exp(−12x

2), where Hn(x)

are polynomials of n-th degree; one can check that Hn(x) coincides

with the n-th Hermite polynomial up to a constant factor. Let us

calculate the normalization constant cn assuming that all of them

are taken to be real and positive. It is obvious that φn = γna+φn−1,

where γn = cn/cn−1. Taking the scalar square of this equality, we

obtain

1 = 〈φn, φn〉

= γ2n

⟨a+φn−1, a

+φn−1

⟩= γ2

n

⟨aa+φn−1, φn−1

⟩= γ2

n

⟨(H +

1

2

)φn−1, φn−1

⟩= nγ2

n,

hence, γn = 1/√n and cn = (n!)−1/2.

The orthonormal system of stationary states φn = 1√n!a+nφ0,

with energy levels En = n+ 1/2, exhausts all stationary states. One

can check this by proving the completeness of functions φn in the

space L2(E1). However, one can give a more direct proof. Let φ

denote a stationary state of H. Let us denote by n then, the minimal

number satisfying anφ0 = 0 (such numbers necessarily exist because

the eigenvalues of the Hamiltonian H are bounded from below).

Then, an−1φ = λφ0, where λ 6= 0. Applying the operator (a+)n−1

to this equation, after some easy calculations, we obtain that φ is

proportional to φn−1.

The more general Hamiltonian H = p2

2m + mω2x2

2 can be reduced

to the Hamiltonian we have considered by means of changing the

system of units. One can, however, use the same consideration in

Page 67: Mathematical Foundations of Quantum Field Theory

March 26, 2020 12:16 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch02 page 20

20 Mathematical Foundations of Quantum Field Theory

this case, taking

a =1√2

(√mωx+

ip√mω

); a+ =

1√2

(√mωx− ip√

).

These operators are also related by the formula [a, a+] = 1, however,

the expression of the Hamiltonian in terms of these operators has

the form H = ω(a+a + 1/2), hence [H, a] = −ωa, [H, a+] = ωa+.

The dependence of operators at and a+t on time has the form at =

a exp(−iωt), a+t = a+ exp(iωt). The stationary states can be written

in terms of the ground state φ0 by the formula φn = (n!)−1/2(a+)nφ0,

with the energies equal to En = (n+ 1/2)ω.

Adding to the quadratic potential energy some higher order

terms with respect to x, we obtain the Hamiltonian of anharmonic

oscillators. It is convenient to express the Hamiltonian of anharmonic

oscillators in terms of the operators a+ and a. For example, if

H = p2

2m + mω2x2

2 + αx3 + βx4, then its expression in terms of a+

and a looks as follows:

H =ω

2+ ωa+a+ γ(a+ + a)3 + δ(a+ + a)4,

where γ = α(2mω)−3/2, δ = β(2mω)−2.

2.6 Multidimensional harmonic oscillator

Let us consider a quantum mechanical system with space of states

R = L2(En) and the Hamiltonian

H = −n∑j=1

αj∂2

∂x2j

+∑i,j

ki,j xixj

(here, αj > 0 and the quadratic form∑kijxixj is positively definite).

In this way, we can write down, for example, the Hamiltonian for a

system of oscillators coupled by means of elastic forces. By means

of a linear change of variables ξi =∑

j aijxj (the coordinates ξi are

called normal coordinates), the Hamiltonian can be written in the

Page 68: Mathematical Foundations of Quantum Field Theory

March 26, 2020 12:16 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch02 page 21

Single-Particle and Non-Identical Particle Systems 21

following form:

H = −1

2

n∑i=1

∂2

∂ξ2i

+1

2

n∑i=1

ω2i ξ

2i .

(This statement follows immediately from the theorem that a

symmetric matrix can be diagonalized by means of an orthogonal

transformation, or from the equivalent theorem that for any two

quadratic forms, one of which is positively definite, one can transform

both into a sum of squares by means of a linear transformation.)

The Hamiltonian, written in normal coordinates, represents a

system of n non-interacting one-dimensional oscillators. Using the

results of Section 2.3, we can calculate its eigenvectors and eigen-

values. Note that the operators a+i = 1√

2(√ωiξi − 1√

ωi∂∂ξi

); ai =1√2(√ωiξi + 1√

ωi∂∂ξi

) allow us to write the Hamiltonian in the form

H =∑n

i=1 ωi(a+i ai + 1/2). The operators a+

i , ai(1 ≤ i ≤ n) satisfy

the relations

[a+i , a

+i ] = [ai, aj ] = 0; [ai, a

+j ] = δij

(these relations, called the canonical commutation relations, will

be explored further in a later section). The ground state Φ of

the Hamiltonian H satisfies the relations aiΦ = 0 for all i =

1, . . . , n. We can obtain the rest of the stationary states from Φ by

means of the operators a+i , i.e. the stationary states have the form

a+v11 a+v2

2 . . . a+vnn Φ, where v1, . . . , vn are non-negative integers; the

corresponding energies are (v1 + 12)ω1 + · · ·+ (vn + 1

2)ωn.

The Hamiltonian H ′ for a multidimensional anharmonic oscillator

has the form H ′ = H + V , where H is the Hamiltonian of harmonic

oscillator and V =∑

k1,...,knck1,...,kn x

k11 . . . xknn . Similar to H, the

Hamiltonian H ′ is also conveniently represented in terms of the

operators a+i , ai. We obtain

H ′ = const +

n∑i=1

ωia+i ai

+∑

k1,...,km,l1,...,ln

Γk1,...,km|l1,...,ln a+k11 . . . a+kn

n al11 . . . alnn ,

where the coefficients Γk|l can be expressed in terms of ck1,...,kn .

Page 69: Mathematical Foundations of Quantum Field Theory

March 26, 2020 12:16 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch02 page 22

22 Mathematical Foundations of Quantum Field Theory

In general, one cannot calculate explicitly the stationary states

of the Hamiltonian H ′ and other physical quantities associated with

this Hamiltonian. However, in the case that the anharmonic terms V

can be considered as a small perturbation, there exist techniques that

allow us to calculate the expansions of physical quantities associated

with H ′ as a power series with respect to the small parameter. These

techniques (Feynman diagrams) are described in Section 6.4.

Page 70: Mathematical Foundations of Quantum Field Theory

March 26, 2020 13:43 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch03 page 23

Chapter 3

Quantum Mechanics of a Systemof Identical Particles

3.1 A system of n identical particles

In Section 2.3, we mentioned that two electrons are identical

particles and that a system of two electrons cannot be described

by means of the prescriptions of Section 2.3. If the system of

two electrons were specified by a wave function ψ(ξ1, ξ2), where

ξi = (ri, si) stands for the coordinate variable ri and the spin

variable si, then we could ask about the physical meaning of the

number∑

s1,s2|ψ(r1, s1, r2, s2)|2 = p(r1, r2). It would be natural to

consider this number as the probability, or more precisely as the

probability density, of the first electron occupying the point r1 and

the second electron occupying the point r2. However, the first and

second electrons are identical, therefore we can only talk about the

probability of finding one of these electrons at the point r1 and the

other electron at the point r2. So we should make some changes

and the right way to make these changes is as follows. For the wave

function of two electrons, we should consider only square-integrable

functions ψ(ξ1, ξ2) = ψ(r1, s1, r2, s2) that are anti-symmetric with

respect to the variables ξ1, ξ2 (i.e. ψ(ξ2, ξ1) = −ψ(ξ1, ξ2)). Then, it

is obvious that p(r1, r2) = p(r2, r1) and that this quantity makes

sense as the probability density measuring the probability of finding

one of the electrons at the point r1 and the other at r2. For the

system of two protons, two neutrons, or two µ-mesons, the situation

is precisely the same. However, for the system of two π0-mesons, we

23

Page 71: Mathematical Foundations of Quantum Field Theory

March 26, 2020 13:43 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch03 page 24

24 Mathematical Foundations of Quantum Field Theory

should consider as wave functions only functions ψ(r1, r2) that are

symmetric with respect to the variables r1, r2.

Let us now give precise definitions.

All particles are divided into two classes — bosons and fermions.

The space Fn of the states of the systems of n identical bosons

(fermions) is a Hilbert space of square-integrable functions of n

variables ψ(ξ1, . . . , ξn) that are symmetric (antisymmetric) with

respect to the variables (ξ1, . . . , ξn) (here, ξi ∈ X, where X is a

measure space). To emphasize that we are dealing with bosons, we

use the notation F sn for the space Fn and in the case of fermions,

we use the notation F an . The space L2(Xn) = L2(X × · · · × X)

will be denoted by Bn. The spaces F sn and F an are subspaces of this

space.

Experiments show that particles with half-integer spin (electrons,

protons, and so on) are fermions and particles with integer spin

(π-mesons, photons, and so on) are bosons. In relativistic quantum

theories, this statement can be derived from theoretical considera-

tions.

It is easy to check that the space Fn is completely determined

by the Hilbert space F1 of states of one particle and the number

n (i.e. it doesn’t depend on the representation of the space F1 of

one particle in the form L2(X)). Indeed, as we have mentioned

already in Section 2.3, the isomorphism α of the spaces L2(X) and

L2(Y ) can be naturally extended to an isomorphism αn of the spaces

L2(Xn) = L2(X×· · ·×X) and L2(Y n) = L2(Y ×· · ·×Y ). It is easy to

check that the isomorphism αn transforms symmetric functions into

symmetric functions and antisymmetric functions into antisymmetric

functions.

If B is a Hilbert space of states of one particle, then the space

of states of n identical particles Fn can be described as an nth

symmetric tensor power of the space B in the case of bosons and

an nth antisymmetric tensor power of the space B in the case of

fermions.

Let us define the operators of symmetrization Ps and

antisymmetrization Pa in the space Bn = L2(Xn) by means of the

Page 72: Mathematical Foundations of Quantum Field Theory

March 26, 2020 13:43 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch03 page 25

Quantum Mechanics of a System of Identical Particles 25

formulas

Psψ(ξ) =1

n!

∑π

ψ(π(ξ)),

Paψ(ξ) =1

n!

∑π

(−1)γπψ(π(ξ))

(here, ξ = (ξ1, . . . , ξn) ∈ Xn, π is a permutation (i1, . . . , in) of the

indices 1, . . . , n, with π(ξ) = (ξi1 , . . . , ξin) ∈ Xn, γπ is the parity of

the permutation π, and∑

π is a sum over all permutations π). In

particular, for n = 2,

Psψ(ξ1, ξ2) =1

2(ψ(ξ1, ξ2) + ψ(ξ2, ξ1)),

Paψ(ξ1, ξ2) =1

2(ψ(ξ1, ξ2)− ψ(ξ2, ξ1)).

It is clear that PsBn = F sn; PaBn = F an ; Psψ = ψ if ψ ∈ F sn, and

Paψ = ψ if ψ ∈ F an .

It is easy to check that Ps and Pa are orthogonal projections of

the space Bn onto the subspaces F sn and F an , respectively.

Let us denote by H an operator on Bn that commutes with

operators Ps and Pa. Then, F sn and F an are H-invariant subspaces.

If ψ ∈ Bn is an eigenfunction of the operator H, then Psψ and

Paψ (if they are non-zero) are also eigenfunctions of the operator

H that belong to subspaces F sn and F an , respectively, with the same

eigenvalue.

Let us consider a system of n non-interacting identical parti-

cles, i.e. a system having a Hamiltonian transforming the function

ψ(ξ1, . . . , ξn) into the function

ψ(ξ1, . . . , ξn) =n∑i=1

∫A(ξi, ηi)ψ(ξ1, . . . , ξi−1, ηi, ξi+1, . . . , ξn)dηi,

(3.1)

in particular, a single-particle Hamiltonian H1 transforms the

function ψ(ξ) into the function ψ(ξ) =∫A(ξ, η)ψ(η)dη (i.e.

Page 73: Mathematical Foundations of Quantum Field Theory

March 26, 2020 13:43 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch03 page 26

26 Mathematical Foundations of Quantum Field Theory

the generalized function A(ξ, η) is a kernel of the operator H1).

Formula (3.1) specifies an operator Hn on the space Bn, commuting

with the operators Ps and Pa, hence this operator transforms the

subspaces F sn and F an into themselves.1 The operators on the spaces

F sn and F an , defined by relation (3.1), will be denoted by Hsn and Ha

n,

respectively. These operators are the Hamiltonians of systems of n

non-interacting bosons or fermions.

Let us assume, for definiteness, that the single-particle Hamilto-

nian H1 has discrete spectrum, and let us denote by φn the complete

orthonormal system of its eigenvectors, with corresponding eigen-

values En. Then it is obvious that the functions φk1(ξ1) . . . φkn(ξn)

constitute a complete orthonormal system of eigenvectors of the

operator Hn in the space Bn (see Section 2.3). It follows that

Psφk1(ξ1) . . . φkn(ξn) are eigenfunctions of the Hamiltonian Hsn and

belong to the space F sn. Similarly, Paφk1(ξ1) . . . φkn(ξn) are eigenvec-

tors of the Hamiltonian Han and belong to F an . The corresponding

eigenvalues, in both cases, are equal to Ek1 + · · ·+Ekn . In the case of

fermions, the functions Paφk1(ξ1) . . . φkn(ξn) should only be consid-

ered in the case of distinct indices k1, . . . , kn because in the case when

a pair of indices coincides, the antisymmetrization gives zero.2 It is

easy to check that the functions Psφk1(ξ1) . . . φkn(ξn) form a complete

system of functions in F sn and the functions Paφk1(ξ1) . . . φkn(ξn)

form a complete system in F an . These functions are not normalized,

however, one can check that two functions belonging to one of these

systems are either orthogonal or proportional. In the coordinate

representation, the space of states of one particle is realized as the

space L2(E3 × B), where B is a finite set, therefore the space of

states of n identical particles can be considered as a space of square-

integrable symmetric (antisymmetric) functions ψ(r1, i1, . . . , rn, in)

1More precisely, the definition of the operator Hn can be given as follows: Hn isa self-adjoint operator on Bn that transforms functions φ1(ξ1) . . . φn(ξn), wherethe functions φi belong to the domain of the operator H1, into the function∑ni=1 φ1(ξ1) . . . φi−1(ξi−1)(H1φi)(ξi)φi+1(ξi+1) . . . φn(ξn). Such an operator is

unique.2In other words, in the system of non-interacting identical fermions, two fermions

cannot be found in the same state. This statement is called the Pauli exclusionprinciple.

Page 74: Mathematical Foundations of Quantum Field Theory

March 26, 2020 13:43 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch03 page 27

Quantum Mechanics of a System of Identical Particles 27

depending on the coordinate variables r1, . . . , rn ∈ E3 and discrete

variables i1, . . . , in ∈ B (the functions ψ, in the case of bosons, are

invariant with respect to simultaneous permutation of coordinates rland rm and discrete variables il and im, while in the case of fermions,

this function changes sign under such permutation of variables).

The operator of total momentum Pn and the Hamiltonian Hn

of a system of n non-relativistic identical particles can be written

precisely in the same way as specified in Section 2.3 ((2.1) and (2.2));

the only difference is that the masses of identical particles are

equal and the potential energy Vn(r1, . . . , rn) should be a symmetric

function of the coordinates. We will always assume that

Vn(r1, . . . , rn) =

n∑j=1

V(rj) +∑

1≤j≤l≤nW(|rj − rl|)

(the first sum corresponds to the potential energy of particles in an

external field and the second sum corresponds to the interaction). It

is easy to see that the operators Pn and Hn transform symmetric

functions and antisymmetric functions into antisymmetric functions,

i.e. they can be considered as operators on the spaces F sn and F an .

The operators Pj = 1i∂∂rj

do not have this property; this means that

for the identical particles, we cannot speak of the momentum of the

jth particle.

In the case of non-interacting particles, the operator of potential

energy has the form Vn(r1, . . . , rn) = V(r1) + · · · + V(rn); it was

shown earlier that the stationary states of n particles, in this case,

can be expressed in terms of stationary states of a single particle.

3.2 Fock space

In Section 3.1, we fixed a measure space X. We have considered

Hilbert spaces F sn and F an of square-integrable symmetric and

antisymmetric functions ψ(ξ1, . . . , ξn), depending on n variables

ξ1, . . . , ξn ∈ X. It is often convenient to consider the direct sums

F s = F s0 + F s1 + · · ·+ F sn + · · · ,

F a = F a0 + F a1 + · · ·+ F an + · · · .

Page 75: Mathematical Foundations of Quantum Field Theory

March 26, 2020 13:43 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch03 page 28

28 Mathematical Foundations of Quantum Field Theory

Here, F s0 and F a0 are spaces of constant functions (i.e. one-

dimensional spaces).

The spaces F s and F a are called Fock spaces. The elements of

Fock spaces represent the states of a system of identical bosons

or fermions in the case when the number of particles is not fixed.

Speaking of bosons and fermions simultaneously, we will use the

notation F for one of the spaces F s or F a. Vectors in the space

F s or F a can be considered as sequences (f0, f1, . . . , fn, . . . ), where

fn ∈ F sn (or correspondingly fn ∈ F an ), satisfying the condition∑∞n=0 ‖fn‖2 < ∞. In other words, the elements of the space F can

be considered as column vectors

f =

f0

f1(ξ1)...

fn(ξ1, . . . , ξn)...

of symmetric (antisymmetric) functions fn(ξ1, . . . , ξn) as entries; they

should satisfy the condition3

∞∑n=0

∫|fn(ξ1, . . . , ξn)|2dξ1 . . . dξn <∞.

Linear combination and scalar product of two such column vectors

are defined in the natural way, in particular, the scalar product of f

and g is

〈f, g〉 =

∞∑n=0

∫fn(ξ1, . . . , ξn)gn(ξ1, . . . , ξn)dξ1 . . . dξn.

The spaces Fn are naturally embedded in the Fock space F (to the

vector f ∈ Fn, there corresponds a sequence (f0, . . . , fn, . . . ) ∈ F ,

where fn = f, fk = 0 for k 6= n).

Let us introduce the operator of the number of particles

N = N0 +N1 + · · ·+Nn + · · · ,

3Such column vectors are called Fock states.

Page 76: Mathematical Foundations of Quantum Field Theory

March 26, 2020 13:43 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch03 page 29

Quantum Mechanics of a System of Identical Particles 29

where Nn is an operator on the space Fn, multiplying every vector

by the number of particles n. The spaces Fn are eigenspaces of the

operator N .

The operator of total momentum P and the Hamiltonian of the

system of non-relativistic particles H in the Fock space are defined

as direct sums of operators Pn and Hn introduced in Section 3.1

P = P0 + P1 + · · ·+ Pn + · · · ,

H = H0 +H1 + · · ·+Hn + · · ·

(the operators P and H act on Fock space, constructed starting with

the measure space E3 ×B).

In non-relativistic quantum mechanics, the operators of all observ-

ables leave the spaces Fn ⊂ F invariant (i.e. they commute with

the operator of the number of particles N). However, even in the

non-relativistic case, it is very convenient to express the operators

of physical quantities in terms of auxiliary operators that do not

commute with N . Namely, we will introduce the operators a(f) and

a+(f) that can be considered as the operators of annihilation and

creation of a particle with the wave function f ∈ L2(X).

First of all, let us define the operator an(f), where f ∈ L2(X),

acting on the space Fn into the space Fn−1. Namely, we will assume

that the operator an(f) transforms a function φ(ξ1, . . . , ξn) ∈ Fn into

the function φ(ξ1, . . . , ξn−1) ∈ Fn−1, defined by the formula

φ(ξ1, . . . , ξn−1) =√n

∫φ(ξ1, . . . , ξn−1, ξ)f(ξ)dξ

(the operator a0(f) transforms the space F0 into zero).

The operator a+n (f), defined on the space Fn−1 and taking values

in the space Fn, can be described as the operator adjoint to an(f). It

is easy to calculate that the operator a+n (f) transforms the function

φ(ξ1, . . . , ξn−1) into the function√nP s(φ(ξ1, . . . , ξn−1)f(ξn)) in the

bosonic case and into the function√nP a(φ(ξ1, . . . , ξn−1)f(ξn)) in the

fermionic case.

Let us consider a subset D of the space F , consisting of finite

sequences (φ0, . . . , φn, . . . ) (in other words, D is the smallest linear

subspace that contains all subspaces Fn). The operators a(f) and

Page 77: Mathematical Foundations of Quantum Field Theory

March 26, 2020 13:43 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch03 page 30

30 Mathematical Foundations of Quantum Field Theory

a+(f), where f ∈ L2(X), will be defined on the space D, namely,

a(f) transforms the sequence (φ0, . . . , φn, . . . ) ∈ D into the sequence

(a1(f)φ1, a2(f)φ2, . . . , an+1(f)φn+1, . . . )

and the operator a+(f), transforms the same sequence into

(0, a∗1(f)φ0, . . . , a∗n(f)φn−1, . . . ).

It is obvious that these operators transform the subspace D into

itself and that they are Hermitian conjugate on D.

It is easy to check (see Section 6.1) that in the case of fermions, the

operators a(f), a+(f) are bounded (‖a(f)‖ = ‖f‖) and therefore can

be extended by continuity to the whole Fock space F ; however, in the

bosonic case, the operators a(f), a+(f) are not bounded (‖an(f)‖ =√n‖f‖).By means of straightforward calculations, it is easy to find the

commutation relations for the operators a(f), a+(f). In the bosonic

case, the operators satisfy the conditions

[a(f), a(g)] = [a+(f), a+(g)] = 0; [a(f), a+(g)] = 〈f, g〉 (3.2)

(these relations are called the canonical commutation relations or

CCR). In the fermionic case, we have similar conditions for the

anticommutators

[a(f), a(g)]+ = [a+(f), a+(g)]+ = 0; [a(f), a+(g)]+ = 〈f, g〉 (3.3)

(these relations are called the canonical anticommutation relations

or CAR).

The operators a(f) depend linearly on f ∈ L2(X). This means

that they can be considered as operator generalized functions on X.

In other words, we can introduce the symbols a(x), a+(x), that are

related to the operators a(f), a+(f) by the following formulas:

a(f) =

∫f(x)a(x)dx,

a+(f) =

∫f(x)a+(x)dx.

Page 78: Mathematical Foundations of Quantum Field Theory

March 26, 2020 13:43 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch03 page 31

Quantum Mechanics of a System of Identical Particles 31

The symbols a(x), a+(x) are not operators. The integral∫f(x)

a(x)dx is considered simply as another notation for the opera-

tors a(f).

Note, however, that in principle, the symbol a(x) can be consid-

ered as an operator on some subset of the space F and the symbol

a+(x) can be considered as an operator acting from the space F into

some larger space (see Appendix A.7).

In what follows, we will consider the expression of the form

A =

∫f(x1, . . . , xm|y1, . . . , yn)a+(x1) . . . a

+(xm)a(y1) . . .

× a(yn)dmxdny, (3.4)

where f is a generalized function. Clearly, such expressions define an

operator on D if the function f takes the form

f1(x1) . . . fm(xm)g1(y1) . . . gn(yn), (3.5)

where fi, gi ∈ L2(X) (then one should assume that A = a+(f1) . . .

a+(fm)a(g1) . . . a(gn)). This definition can be extended to the case

when the function f is a finite sum of products of the form (3.5).

One can prove easily that under this condition the operator A

transforms a sequence φ = (φ0, . . . , φk, . . . ) ∈ D, into the sequence

ψ = (ψ0, . . . , ψk, . . . ) ∈ D, where

ψk(ξ1, . . . , ξk) =

√k!

(k −m)!· (k −m+ n)!

(k −m)!

×P∫φn−m+k(ξ1, . . . , ξk−m, x1, . . . , xn)

× f(ξk−m+1, . . . , ξk|x1, . . . , xn)dnx (3.6)

(P denotes the operator of symmetrization in the bosonic case and

antisymmetrization in the fermionic case).

We can use the formula (3.6) to define the operator A for an

arbitrary function f . The domain of the operator A, in this case,

consists of all sequences φ ∈ D, for which the functions ψk(ξ1, . . . , ξk),

obtained by the formula (3.6), are square-integrable. Considering

Page 79: Mathematical Foundations of Quantum Field Theory

March 26, 2020 13:43 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch03 page 32

32 Mathematical Foundations of Quantum Field Theory

expressions of the form

A =

∞∑m,n

∫Am,n(x1, . . . , xm|y1, . . . , yn)a+(x1)

. . . a+(xm)a(y1) . . . a(yn)dmxdny, (3.7)

we define every summand using the consideration above and the

sum of the series we understand in terms of strong convergence.

The representation of the operator A in the form (3.7) is called the

representation in normal form (in the expression (3.7), the creation

operators stand to the left of the operators of annihilation).

Every bounded operator can be represented in the form (3.7) if

the convergence of the series is understood as weak convergence.

The vector θ, specified by the sequence (1, 0, 0, . . . ), plays a special

role in the Fock space F . This vector corresponds to the state that

contains no particles (θ ∈ F0) and is called the vacuum vector ; it

satisfies the condition a(f)θ = 0 for all functions f ∈ L2(X). It is easy

to check that the vector θ ∈ F is a cyclic vector with respect to the

family of operators a+(f) (in other words, linear combinations of the

vectors of the form a+(f1) . . . a+(fs)θ are dense in F ). One can write

down an explicit expression of every vector φ = (φ0, . . . , φn, . . . ) ∈ Fin terms of the generalized operator functions a+(x) and the vector

θ, namely, by applying the formula (3.6), we see that

f =∞∑n=0

1√n!

∫fn(ξ1, . . . , ξn)a+(ξ1) . . . a

+(ξn)θdnξ. (3.8)

Let us consider the simplest operators on the Fock space. Let us

start with operators of the form∫A(x, y)a+(x)a(y)dxdy. (3.9)

The operator (3.9) transforms the sequence (φ0, . . . , φk, . . . ) ∈ D into

the sequence (ψ0, . . . , ψk, . . . ), where

ψk(ξ1, . . . , ξk) = kP

∫φk(ξ1, . . . , ξk−1, x)A(ξk, x)dx. (3.10)

Page 80: Mathematical Foundations of Quantum Field Theory

March 26, 2020 13:43 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch03 page 33

Quantum Mechanics of a System of Identical Particles 33

It is easy to check that in the case when the function A(x, y) is

a kernel of a self-adjoint operator on L2(X), the operator (3.9) is

essentially self-adjoint on F . Let us denote by A the self-adjoint

extension of the operator∫A(x, y)a+(x)a(y)dxdy. The operator A

commutes with the operator N , hence, it transforms the space Fninto itself. The operator induced by the operator A on the space Fnwill be denoted by An. Operators of the form An were considered

earlier, namely, the Hamiltonian of a system of n non-interacting

identical particles can be written in this form. In Section 3.1, we

have found eigenvectors and eigenvalues of the operator An in the

assumption that the operator A1 is a self-adjoint operator with

discrete spectrum in L2(X). Using this result, we can obtain the

description of eigenvectors and eigenvalues of the self-adjoint oper-

ator A. Namely, the vectors∏∞i=1(a

+(φi))niθ constitute a complete

system of eigenvectors of the operator A and the corresponding eigen-

values are equal to∑∞

i=1 niEi (here, φ1, . . . , φi, . . . is the complete

system of eigenfunctions of the operator A1 and E1, . . . , En, . . . are

the corresponding eigenvalues; the numbers ni constitute a finite

sequence. In the fermionic case, the numbers ni are equal to 0 or 1 and

in the bosonic case n = 0, 1, 2, . . . ). The proof can be reduced to the

remark that the vector a+(f1) . . . a+(fn)θ is equal to the vector that

corresponds in Fock space to the function Pf1(ξ1) . . . fn(ξn)θ ∈ Fn,

up to a constant factor (recall that Fn is naturally embedded in F ).

The operators

B =

∫B(x1, x2|y1, y2)a+(x1)a

+(x2)a(y1)a(y2)dx1dx2dy1dy2

(3.11)

also commute with the operator of the number of particles (as well as

the operators of the form (3.4), obeying m = n). This means that the

operator B transforms the subspace Fn into itself and it follows from

formula (3.6) that after restricting the operator B to the subspace

Fn, we obtain an operator Bn transforming the vector φ into function

ψk(ξ1, . . . , ξk) = k(k − 1)P

∫φk(ξ1, . . . , ξk−2, x1, x2)

×B(ξk−1, ξk|x1, x2)dx1dx2. (3.12)

Page 81: Mathematical Foundations of Quantum Field Theory

March 26, 2020 13:43 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch03 page 34

34 Mathematical Foundations of Quantum Field Theory

Now we can write down the operators N,P, and H in the normal

form.

The operator N can be represented as

N =

∫a+(ξ)a(ξ)dξ (3.13)

(or, more precisely, the operator N can be expressed by the for-

mula (3.9), where A(ξ, η) = δ(ξ, η)). This follows immediately from

the relation (3.10), which implies that the operator N transforms the

sequence (φ0, . . . , φk, . . . ) ∈ D into the sequence (ψ0, . . . , ψk, . . . ),

where

ψk(ξ1, . . . , ξk) = kP

∫φk(ξ1, . . . , ξk−1, x)δ(ξk, x)dx

= kPφk(ξ1, . . . , ξk) = kφk(ξ1, . . . , ξk).

The momentum operator P and the Hamiltonian H of a system

of non-relativistic identical particles act on Fock space, constructed

starting from the measure space E3 × B. The operators acting on

this space can be expressed in terms of the operator generalized

functions a+(x, s), a(x, s) in the coordinate representation and in

terms of the operator generalized functions a+(k, s), a(k, s) in the

momentum representation (here, x,k ∈ E3, s ∈ B). In the following,

we will sometimes use the notation a+s (k) = a+(k, s), as(k) = a(k, s).

Again, using the relation (3.10), it is easy to derive that in the

coordinate representation, the operator P is

P =∑s

1

i

∫a+(x, s)

∂xa(x, s)dx, (3.14)

or, more precisely,

P =∑s,s′

∫A(x, s,y, s′)a+(x, s)a(x, s′)dxdy,

where A(x, s,y, s′) = δs′s

1i∂∂xδ(x− y). In the momentum representa-

tion, the operator P takes the form

P =∑s

∫ka+(k, s)a(k, s)dk. (3.15)

Page 82: Mathematical Foundations of Quantum Field Theory

March 26, 2020 13:43 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch03 page 35

Quantum Mechanics of a System of Identical Particles 35

It follows from the relations (3.10) and (3.12) that the Hamiltonian H

of the system of non-relativistic identical particles can be represented

as a sum of operators of the form (3.9) and (3.11). Namely, in the

coordinate representation,

H =∑s

∫a+(x, s)

(− ∆

2m

)a(x, s)dx

+∑s

V(x)a+(x, s)a(x, s)dx +∑s

1

2

∫W(|x1 − x2|)a+(x1, s)

× a+(x2, s)a(x2, s)a(x1, s)dx1dx2, (3.16)

and in the momentum representation,

H =∑s

∫k2

2ma+(k, s)a(k, s)dk

+∑s

V(k1 − k2)a+(k1, s)a(k2, s)dk1dk2

+∑s

1

2

∫W(k1 − k4)δ(k1 + k2 − k3 − k4)

× a+(k1, s)a+(k2, s)a(k3, s)a(k4, s)d

4k, (3.17)

where V(k), W(k) are the Fourier transforms of the functions

V(x),W(x).

It is easy to construct examples of expressions of the form (3.7)

that do not define an operator in Fock space. For example, the

expression

A =

∫α(x1, . . . , xm)a+(x1) . . . a

+(xm)dx1 . . . dxm (3.18)

can define an operator on Fock space only in the case when the

function α is square integrable. Namely, by the general definition,

the operator defined by the expression (3.8) should transform the

sequence (φ0, . . . , φn, . . . ) ∈ D into the sequence (ψ0, . . . , ψn, . . . ),

Page 83: Mathematical Foundations of Quantum Field Theory

March 26, 2020 13:43 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch03 page 36

36 Mathematical Foundations of Quantum Field Theory

where

ψn(ξ1, . . . , ξn) =

√n!

(n−m)!Pφn−m(ξ1, . . . , ξn−m)α(ξn−m+1, . . . , ξn).

If the function α is square integrable, then the functions ψn are also

square integrable, hence the expression (3.18) specifies an operator

A on the whole space D and this operator transforms D into

itself. However, if the function α is not square integrable, then

the function ψn may be square integrable only under the condition

φn−m ≡ 0. This means that the domain of the operator specified by

the expression (3.18) contains only the zero vector; in other words,

the expression (3.18) does not specify any operator because the

domain of the operator should be dense in Fock space.

Let us consider the Fock space F (L2(E3)), constructed starting

with the measure space E3, and operators defined by expressions of

the form

A =

m+n≤s∑m,n

∫Am,n(k1, . . . ,km|l1, . . . , ln)

× δ(k1 + · · ·+ km − l1 − · · · − ln)

× a+(k1) . . . a+(km)a(l1) . . . a(ln)dmkdnl (3.19)

(operators of this form commute with the momentum operator). Let

us suppose that the functions Am,n belong to the space S(E3(m+n))

of smooth, rapidly decreasing functions (faster than any power

function). Let us single out the subspace S∞ ⊂ F that consists

of sequences (φ0, . . . , φk, . . . ) ∈ D, obeying φk ∈ S(E3k). In what

follows, it will be convenient to consider operators on the space

F (L2(E3)) only on the set S∞. By means of the relation (3.6), it is

easy to check that in the case when Am,0 ≡ 0, the operator specified

by the expression (3.19) is well defined on all elements of the set

S∞. This operator transforms every sequence in S∞ into a sequence

belonging to the same set. If one of the functions Am,0 is non-zero,

then the expression (3.19) cannot define an operator on Fock space

because the function Am,0(k1, . . . ,km)δ(k1 + · · ·+km) is not square

integrable.

Page 84: Mathematical Foundations of Quantum Field Theory

March 26, 2020 13:43 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch03 page 37

Quantum Mechanics of a System of Identical Particles 37

Let us consider the fermionic case. In this case, we can introduce

the notion of even vectors as vectors belonging to a direct sum

of subspaces F2n and odd vectors as vectors belonging to a direct

sum of subspaces F2n+1. An operator on Fock space is called

parity-preserving if it transforms an even vector to an even vector

and an odd vector to an odd vector. Introducing an involution τ ,

transforming a sequence (φ0, φ1, . . . , φ2n, φ2n+1, . . . ) into a sequence

(φ0,−φ1, . . . , φ2n,−φ2n+1, . . . ), we can say that an even vector is

invariant with respect to this involution, τx = x, and an odd

vector satisfies the condition τx = −x. A parity-preserving operator

commutes with this involution. An operator represented in the

form (3.7) (normal form) is parity preserving if the functions Am,ndo not vanish only in the case when the numbers m and n have the

same parity.

Operators corresponding to physical quantities should be parity

preserving. In particular, the Hamiltonian is always a parity-

preserving operator.

Page 85: Mathematical Foundations of Quantum Field Theory

b2530 International Strategic Relations and China’s National Security: World at the Crossroads

b2530_FM.indd 6 01-Sep-16 11:03:06 AM

This page intentionally left blankThis page intentionally left blankThis page intentionally left blankThis page intentionally left blank

Page 86: Mathematical Foundations of Quantum Field Theory

March 26, 2020 14:5 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch04 page 39

Chapter 4

Operators of Time EvolutionS(t, t0) and Sα(t, t0)

4.1 Non-stationary perturbation theory

Let us consider a Hamiltonian H(t) having the form H(t) = H0 +

gV (t). Let us assume that we can calculate the operator of evolution

U0(t, t0) = exp(−iH0(t − t0)) corresponding to the Hamiltonian

H0. We will solve the problem of finding the evolution operator

U(t, t0) from the Hamiltonian H(t) as a series with respect to the

parameter g. It is convenient to introduce the operator S(t, t0) =

exp(iH0t)U(t, t0) exp(−iH0t0), expressed in terms of the operator

U(t, t0), and calculate it. We will find out later that the operator

S(t, t0) is very important on its own. In the case when we should

emphasize that the operators U(t, t0) and S(t, t0) depend on the

parameter g, we will use the notation U(t, t0|g) and S(t, t0|g).

The operator U(t, t0) satisfies equation (1.1). Using this equation,

we can easily obtain the following equation for the operator S(t, t0):

i∂S(t, t0)

∂t= gV (t)S(t, t0), (4.1)

where V (t) = exp(iH0t)V (t) exp(−iH0t); the initial condition is

specified in the form S(t0, t0) = 1.

In some physical situations, the operator H0 can be considered as

a free Hamiltonian (i.e. it describes non-interacting particles), and

the operator H − H0 = gV (t) is the interaction Hamiltonian. This

terminology is also commonly used in quantum field theory. However,

39

Page 87: Mathematical Foundations of Quantum Field Theory

March 26, 2020 14:5 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch04 page 40

40 Mathematical Foundations of Quantum Field Theory

in quantum field theory, as we will see later, there exists no natural

partition of the Hamiltonian into the free and interaction parts. We

will not use these terms in order to avoid misleading associations.

We will express S(t, t0) as a power series in the variable g as

follows:

S(t, t0) =∞∑n=0

gnSn(t, t0).

Substituting this series into the equation for the operator S(t, t0), we

obtain the recurrence relation

i∂Sn(t, t0)

∂t= V (t)Sn−1(t, t0).

Using the initial condition S(t0, t0) = 1, we see that S0(t0, t0) = 1

and Sn(t0, t0) = 0 for n ≥ 1, hence

Sn(t, t0) =1

i

∫ t

t0

V (τ)Sn−1(τ, t0)dτ. (4.2)

The differential equation for S with an initial condition is equivalent

to the integral equation

S(t, t0) = 1 + g

∫ t

t0

V (τ)S(τ, t0)dτ,

one can solve the integral equation using the method of iterations.

From (4.2) or from the integral equation, we can conclude that

S1(t, t0) =1

i

∫ t

t0

V (τ)dτ,

S2(t, t0) =

(1

i

)2 ∫ t

t0

dτ1

∫ τ1

t0

dτ2V (τ1)V (τ2),

...

Sn(t, t0) =

(1

i

)n ∫ t

t0

dτ1

∫ τ1

t0

dτ2· · ·∫ τn−1

t0

dτnV (τ1) . . . V (τn).

Page 88: Mathematical Foundations of Quantum Field Theory

March 26, 2020 14:5 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch04 page 41

Operators of Time Evolution S(t, t0) and Sα(t, t0) 41

We can also write

Sn(t, t0) =

(1

i

)n ∫Γn

dτ1 . . . dτnV (τ1) . . . V (τn),

where the domain of integration Γn is defined by t ≥ τ1 ≥ · · · ≥τn≥ t0.

We introduce the following notation:

T (V (τ1) . . . V (τn)) = V (τi1) . . . V (τin),

where i1, . . . , in is a permutation of the indices 1, . . . , n, satisfying

the relation τi1 ≥ · · · ≥ τin . In other words, T (V (τ1) . . . V (τn))

(chronological or a T-product of the operators V (τ1) . . . V (τn)) is

defined as the product of the operators V (τ1), . . . , V (τn) in order

of decreasing time τi. If some of the times τi are identical, then the

permutation for which τi1 ≥ · · · ≥ τin is not unique, however, it is

easy to see that the T -product does not depend on this choice of

permutation. Using the T -product, we can write Sn(t, t0) in the form

Sn(t, t0) =1

n!

(1

i

)n ∫ t

t0

· · ·∫ t

t0

T (V (τ1) . . . V (τn))dτ1 . . . dτn.

Indeed, the domain of integration in this integral can be partitioned

into n! regions GP , with each region GP corresponding to a permu-

tation P = (j1, . . . , jn) of the indices 1, . . . , n. Here, GP is the region

singled out by the inequalities t ≥ τj1 ≥ · · · ≥ τjn ≥ t0. It is clear

that (1

i

)n ∫GP

T (V (τ1) . . . V (τn))dτ1 . . . dτn

=

(1

i

)n ∫GP

V (τj1) . . . V (τjn)dτ1 . . . dτn,

and that this integral, after a change of variables, is reduced to

the integral Sn(t, t0). This means, that the integral 1n!(

1i )n∫ tt0. . .∫ t

t0T (V (τ1) . . . V (τn))dτ1 . . . dτn splits into n! identical integrals,

equal to Sn(t, t0); this gives the needed formula. The full operator

Page 89: Mathematical Foundations of Quantum Field Theory

March 26, 2020 14:5 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch04 page 42

42 Mathematical Foundations of Quantum Field Theory

S(t, t0) can be written as the series

S(t, t0) =∞∑n=0

1

n!

(1

i

)ngn∫ t

t0

· · ·∫ t

t0

T (V (τ1) . . . V (τn))dτ1 . . . dτn.

(4.3)

This series can be compactly written in the form

S(t, t0) = T exp

(1

ig

∫ t

t0

V (τ)dτ

), (4.4)

which is called T-exponential. If the operator V (t) is bounded for

every t and is continuous (in the sense of strong limits) in t, then the

series (4.3) is convergent in norm (hence, in this case, the existence

of solutions of equations (1.1) and (4.1) follows). Indeed, we have

that supt0≤τ≤t ‖V (τ)‖ = supt0≤τ≤t ‖V (τ)‖ < ∞ and ‖Sn(t, t0)‖ ≤1n!(supt0≤τ≤t ‖V (τ)‖)n|t− t0|n. In the case when it is not possible to

prove the convergence of the series (4.3), the corresponding results

are conditional (if a series expansion in g is possible, then the series

takes the form (4.3)). Regardless of the possibility of the series

expansion in g, one can prove the following relation:

∂nS(t, t0|g)

∂gn|g=0 =

(1

i

)n ∫ n

t0

· · ·∫ t

t0

T (V (τ1) . . . V (τn))dτ1 . . . dτn.

If the HamiltonianH can be represented in the formH0+V , where

H0 and V do not depend on time, then it is convenient to consider the

Hamiltonian Hα(t) = H0 + exp(−α|t|)V . We introduce the operator

Sα(t, t0) = exp(iH0t)Uα(t, t0) exp(−iH0t0), where Uα(t, t0) is the

operator of evolution, constructed with Hα(t).

In a similar way, we can define the operator Sα(t, t0) as the

solution of the differential equation

i∂Sα(t, t0)

∂t= exp(−α|t|)V (t)Sα(t, t0)

with the initial condition Sα(t0, t0) = 1 (here, V (t) = exp(iH0t)

V exp(−iH0t)). Following the discussion above, we can write Sα(t, t0)

Page 90: Mathematical Foundations of Quantum Field Theory

March 26, 2020 14:5 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch04 page 43

Operators of Time Evolution S(t, t0) and Sα(t, t0) 43

in the form of the T -exponent:

Sα(t, t0) = T exp

(1

i

∫ t

t0

exp(−α|τ |)V (τ)dτ

).

The operator Sα(∞,−∞) = slim t→∞t0→−∞

Sα(t, t0), denoted by Sα,

plays a special role and is called the adiabatic S-matrix. Two other

important operators Sα(0,±∞) = slimt→±∞ Sα(0, t) will be denoted

Sα+, and Sα− and are called the adiabatic Møller operators.

In the case when H0 can be considered a free Hamiltonian

and V is the interaction, considering the Hamiltonian Hα(t) with

α → 0 describes adiabatic (infinitesimally slow) turning on and

off the interaction; this explains the above-introduced terminology.

In place of the Hamiltonian H = H0 + V , it is often useful

to consider the family of Hamiltonians Hg = H0 + gV , where

g is a constant, which is typically called the coupling constant.

The operators Sα(t, t0), Sα(∞,−∞), Sα(0,±∞), constructed with

the Hamiltonian Hg, will be denoted by Sα(t, t0|g), Sα(∞,−∞|g) =

Sα(g), and Sα(0,±∞|g) = Sα±(g).

4.2 Stationary states of Hamiltonians depending

on a parameter

Let us consider the family of Hamiltonians H(g), where 0 ≤ g ≤ g0.

We will assume that for every g, there exists a normalized eigenvector

φg of the operator H(g), with eigenvalue E(g), which is differentiable

with respect to the parameter g. Without loss of generality, we can

assume that ⟨φg,

dφgdg

⟩= 0.

Indeed, if this condition is not satisfied, one can always find a phase

factor exp(iα(g)), such that the vector φg = exp(−iα(g))φg satisfies

the condition: namely, we can choose α(g) = 1i

∫ g0

⟨φλ,

dφλdλ

⟩dλ (the

function α(g) is real, since from the relation 〈φλ, φλ〉 = 1, one can

Page 91: Mathematical Foundations of Quantum Field Theory

March 26, 2020 14:5 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch04 page 44

44 Mathematical Foundations of Quantum Field Theory

obtain that⟨φλ,

dφλdλ

⟩+

⟨dφλdλ

, φλ

⟩=

⟨φλ,

dφλdλ

⟩+

⟨φλ,

dφλdλ

⟩= 0).

By differentiating in g the relation

H(g)φg = E(g)φg,

we obtain

(H(g)− E(g))dφgdg

= −dH(g)

dgφg +

dE(g)

dgφg. (4.5)

Taking the scalar product of this relation with φg, we obtain

dE(g)

dg=

⟨dH(g)

dgφg, φg

⟩. (4.6)

Let us consider the commonly occurring case of H(g) = H0 + gV ;

then, it is clear that

dE(g)

dg= 〈V φg, φg〉 , (4.7)

(H0 + gV − E(g))dφgdg

= (〈V φg, φg〉 − V )φg. (4.8)

We will sometimes assume that the vector φg is analytic with respect

to g in the neighborhood around g = 0, and we will search for an

expansion of the vector φg and the eigenvalue E(g) in a series in

g (stationary perturbation theory). Formulas (4.7) and (4.8), when

g = 0, provide the linear terms in g (first-order terms), namely,

E(g) = E(0) + g 〈V φ0, φ0〉+ · · · ,

φg = φ0 + gψ + · · · ,

where ψ satisfies the relation

(H0 − E(0))ψ = (〈V φ0, φ0〉 − V )φ0. (4.9)

Given that the vector φg satisfies the equation⟨φg,

dφgdg

⟩= 0, we

obtain the following condition on the vector ψ:

〈ψ, φ0〉 = 0. (4.10)

Page 92: Mathematical Foundations of Quantum Field Theory

April 6, 2020 15:51 Mathematical Foundations of Quantum Field Theory 9in x 6in 11222-ch04 page 45

Operators of Time Evolution S(t, t0) and Sα(t, t0) 45

In the case when E(0) is a simple eigenvalue of the Hamiltonian H0,

then equations (4.9) and (4.10) give an unambiguous definition of the

vector ψ. The derivation of the relations (4.7) and (4.8) is rigorous

if all the operators H0 + gV have the same domain.

We did not consider the more subtle question of the existence of

eigenvectors φg that depend on the parameter g smoothly. With mild

assumptions, the answer to this question is given by the following

theorem [Kato, 2013].

Let H0 denote a self-adjoint operator and V a Hermitian operator

with domain that contains the domain of H0. Let us suppose that

E0 is a simple isolated eigenvalue of the operator H0 and φ0 is

the corresponding eigenvector. Then for sufficiently small g: (1) the

operator H(g) = H0 + gV is a self-adjoint operator with the

same domain as the operator H0; (2) there exists an eigenvector

φg of the operator H(g), which is analytic in g and is equal to

φ0 for g = 0; (3) the corresponding eigenvalue E(g) also depends

analytically on g and is simple; (4) the relations (4.7) and (4.8) are

valid.

We do not wish to give a direct calculation of the higher terms

with respect to g in the series for φg and E(g). Instead, we will prove

the formula that allows us to get the decomposition of eigenvectors

and eigenvalues in a series in g from non-stationary perturbation

theory. Let us suppose that for 0 ≤ g ≤ g0, the self-adjoint operators

H(g) = H0 + gV have the same domain and E(g) is an isolated

eigenvalue of the operator H(g) that depends continuously on the

parameter g in the interval 0 ≤ g ≤ g0, and φg is the eigenvector

corresponding to this eigenvalue, obeying⟨φg,

dφgdg

⟩= 0. Then

φg = lima→0

exp

(iC(g)

α

)Sα(0,−∞|g)φ0,

φg = lima→0

exp

(iC(g)

α

)Sα(0,+∞|g)φ0,

(4.11)

where C(g) =∫ g

0E(λ)−E(0)

λ dλ and Sα(0,∓∞|g) are adiabatic Møller

matrices, constructed with the pair of operators H(g), H0. This

statement is proven in Section 4.3.

Page 93: Mathematical Foundations of Quantum Field Theory

March 26, 2020 14:5 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch04 page 46

46 Mathematical Foundations of Quantum Field Theory

4.3 Adiabatic variation of stationary state

Let us suppose that the Hamiltonian H(t) varies slowly (adiabati-

cally) with time.

Let us consider a solution of the Schrodinger equation (1.2)

with initial condition ψt0 = φ0, where φ0 is a stationary state of

the Hamiltonian H(t0) corresponding to a non-degenerate energy

level. We will show that for every t this solution is very close

to the stationary state of the Hamiltonian H(t). (When we are

talking about the stationary state of Hamiltonian H(t), we always

assume that we consider an eigenvector of H(t) for a fixed t.) For

definiteness, we restrict ourselves to the case when we consider a

family of Hamiltonians Hg that depend on the parameter g, and the

Hamiltonian H(t) is defined by the formula H(t) = Hαt, where α is

a small positive number. Then equation (1.2) can be reduced to the

equation

iαdψ(g)

dg= Hgψ(g), (4.12)

by means of the change of variables g = αt. In what follows, we will

consider equation (4.12).

Let us assume that Hg is a family of self-adjoint operators with

the same domain D, smoothly depending on the parameter g in the

interval g0 ≤ g ≤ g1 (i.e. for any x ∈ D, the vector Hgx smoothly

depends on g). We will assume that for every g in the interval

g0 ≤ g ≤ g1, there exists a stationary state φg of the Hamiltonian

Hg, smoothly depending on the parameter g. The energy level E(g),

corresponding to the state φg, will be supposed isolated and non-

degenerate. The state φg is assumed to be normalized and obeying

〈φg, dφgdg 〉 = 0. Finally, we will assume thatdnφgdgn |g=g0 = 0 for

n = 1, 2, . . . .

Lemma 4.1. Assuming the conditions above, the solution of equa-

tion (4.12) that coincides with φg0 with g = g0 can be written in the

form

ψ(g) = exp

[− iαC(g)

](φg + αsg + α2r(g, α)),

Page 94: Mathematical Foundations of Quantum Field Theory

April 6, 2020 15:51 Mathematical Foundations of Quantum Field Theory 9in x 6in 11222-ch04 page 47

Operators of Time Evolution S(t, t0) and Sα(t, t0) 47

where C(g) =∫ g

0 E(λ)dλ, sg is defined by the relation

idφgdg

= (Hg − E(g))sg,

⟨dsgdg

, φg

⟩= 0, sg0 = 0,

and the norm of the vector r(g, α) is bounded above by a constant that

does not depend on g or α (here, g lies in the interval g0 ≤ g ≤ g1).

To prove this lemma, let us first use the change of variables ψ(g) =

exp( iαC(g))σ(g) to transform equation (4.12) to the form

iαdσ(g)

dg= (Hg − E(g))σ(g). (4.13)

Let us assume that the solution of equation (4.13) has the form

σ(g) =

∞∑n=0

αnσn(g). (4.14)

By comparing terms with equal powers of α, we obtain the relations

(Hg − E(g))σ0(g) = 0, (4.15)

idσ0(g)

dg= (Hg − E(g))σ1(g),

...

idσn−1(g)

dg= (Hg − E(g))σn(g), (4.16)

σn(g0) = 0 for n ≥ 1. (4.17)

Furthermore, taking the scalar product of these relations with φg, we

obtain ⟨dσ0(g)

dg, φg

⟩= 0, (4.18)⟨

dσ1(g)

dg, φg

⟩= 0, (4.19)

...⟨dσn(g)

dg, φg

⟩= 0. (4.20)

Page 95: Mathematical Foundations of Quantum Field Theory

March 26, 2020 14:5 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch04 page 48

48 Mathematical Foundations of Quantum Field Theory

Equation (4.15) and relation (4.18) are satisfied for σ0(g) = φg. It

then follows that σ1(g) can be found from relations (4.16) and (4.19);

we also obtain that σ1(g) = sg. By the recursive application of

relations (4.17) and (4.20), we can find σn(g). Note that the σn(g)

are specified by these relations uniquely (this follows from the fact

that the eigenvalue E(g) is non-degenerate and isolated) and are

smooth in g. From equationdnφgdgn |g=g0 = 0, it follows that all the

above relations can be satisfied.

Let us show now that

σ(g) =

N∑n=0

αnσn(g) + αN+1rN (g, α), (4.21)

where |rN (g, α)| ≤ K. Indeed, by inserting (4.21) into (4.14), we

obtain the following equations for rN (g, α):

iα∂rN (g, α)

∂g= (Hg − E(g))rN (g, α)− idσN (g)

dg(4.22)

with initial condition rN (g0, α) = 0. Let us denote by V (g) the

unitary operator defined by the relation

iαdV (g)

dg= (Hg − E(g))V (g),

V (g0) = 1.

Now, we can make a change of variables rN (g, α) = V (g)ρN (g, α) in

the equation (4.22). We obtain

iα∂ρN (g, α)

∂g= −iV −1(g)

dσN (g)

dg,

hence

‖rN (g, α)‖ = ‖ρN (g, α)‖ ≤∫ g

g0

∥∥∥∥dρN (g′, α)

dg′

∥∥∥∥ dg′≤ 1

α

∫ g

g0

∥∥∥∥dσN (g′)

dg′

∥∥∥∥ dg′ ≤ const

α. (4.23)

Page 96: Mathematical Foundations of Quantum Field Theory

April 6, 2020 15:51 Mathematical Foundations of Quantum Field Theory 9in x 6in 11222-ch04 page 49

Operators of Time Evolution S(t, t0) and Sα(t, t0) 49

We can now see that

rN (g, α) = σN (g) + αrN+1(g, α).

Applying inequality (4.23) to rN+1, we can derive the following

inequality for rN :

‖rN (g, α)‖ ≤ ‖σN (g)‖+ αconst

α≤ const.

Since r(g, α) = r1(g, α), the above estimate with N = 1 confirms

Lemma 4.1.

Remark 4.1. In the above discussion, the condition of smooth

dependence of Hg and φg on g can be relaxed to the requirement

of three-fold differentiability; the conditiondnφgdgn |g=g0 = 0 can be

relaxed to hold only for n = 1.

Remark 4.2. If we do not require thatdφgdg |g=g0 = 0, then it is

clear, from the above proof, that the proof of Lemma 4.1 applies

to the solution of equation (4.12), satisfying the initial condition

ψg0 = φg0 + αsg0 .

We will now show how to obtain the relation (4.11) using

Lemma 4.1. Let us assume that the family of Hamiltonians Hg satisfy

the conditions of Lemma 4.1, with the exception of the requirement

thatdnφgdgn |g=g0 = 0 for n ≥ 1. Let us consider the operator of evolution

Uα(t, T ) constructed with the Hamiltonian Hα(t) = Hh(αt), where

h(τ) is a smooth function defined on τ ≤ 0 and taking values in the

interval [g0, g1]. Let us further assume that h(τ) = g0 for τ ≤ a.

Then the family of Hamiltonians Hλ = Hh(λ) with λ in [a, 0] and the

vectors φλ = φh(λ) satisfy the conditions of Lemma 4.1, including the

condition dnφλdλn |λ=a = 0 (the latter condition follows from the relation

dnh(λ)dλn |λ=a = 0). Applying the statement of this lemma, we see

that ∥∥∥∥Uα (0,a

α

)φα − exp

(− 1

αC

)φ0

∥∥∥∥→ 0,

Page 97: Mathematical Foundations of Quantum Field Theory

March 26, 2020 14:5 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch04 page 50

50 Mathematical Foundations of Quantum Field Theory

where C =∫ 0a E(h(λ))dλ. We obtain

φh(0) = φ0 = lima→0

exp

(i

αC

)Uα

(0,a

α

)φh(a). (4.24)

Introducing the operator

Sα(t, T ) = exp(iHg0t)Uα(t, T ) exp(−iHg0T ),

we can rewrite the equation (4.14) in the form

φh(0) = lima→0

exp

i

α

∫ 0

a[E(h(λ))− E(h(a))]dλ

(0,a

α

)φh(a).

(4.25)

Since Sα( aα , T ) = 1 for T < aα and h(λ) = h(a) for λ < a, using the

relation (4.25), we can show that

φh(0) = lima→0

exp

i

α

∫ 0

−∞[E(h(λ))− E(h(−∞))]dλ

×Sα (0,−∞)φh(−∞). (4.26)

To obtain (4.11) from (4.26), we must choose for the family of

Hamiltonians Hg the family H0 + gV with 0 ≤ g ≤ g1, and for

h(τ) choose the function g exp(−α|τ |). The function exp(−α|τ |) does

not vanish for τ 0 as required of the function h(τ), therefore,

strictly speaking, we cannot use equation (4.26). However, a slight

modification of the above proof, based on Remark 4.2, allows us to

verify (4.11).

Remark 4.3. If the family of Hamiltonians Hg also depends on

another parameter Ω, then it is not difficult to outline the conditions

under which the limit in (4.26) is uniform in Ω (for this, it is

necessary to give uniform in Ω estimates in the proof of the lemma).

In particular, if HΩg = HΩ

0 +gV Ω (0 ≤ g ≤ g1,Ω ∈ O), then the limit

in (4.26) will be uniform in g and in Ω if the norm of the operators

V Ω is bounded by a constant not depending on Ω and it is possible

to find a δ in such a way that the interval (EΩ(g)− δ, EΩ(g) + δ) for

any Ω and g contains no eigenvalues of the operator HΩg except for

EΩ(g) (see Tyupkin and Shvarts, 1972).

Page 98: Mathematical Foundations of Quantum Field Theory

March 26, 2020 14:41 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch05 page 51

Chapter 5

The Theory of Potential Scattering

5.1 Formal scattering theory

Let H and H0 be two self-adjoint operators on the space R. The

Møller matrices S+ and S− of the operator pair (H,H0) are defined

as strong limits

S− = slimt→−∞

exp(itH) exp(−itH0) = slimt→−∞

S(0, t), (5.1)

S+ = slimt→+∞

exp(itH) exp(−itH0) = slimt→+∞

S(0, t). (5.2)

The operators S− and S+ are isometries, as strong limits of

unitary operators, however, they are not necessarily unitary. If they

are unitary, one can check that

S∗− = slimt→−∞

exp(itH0) exp(−itH),

S∗− = slimt→+∞

exp(itH0) exp(−itH)

(see Appendix A.5).

The S-matrix of the pair of operators (H,H0) is defined by the

formula

S = S∗+S−.

If the operators S+ and S− are unitary, then the S-matrix is also

unitary and can be written in the form

S = slimt→∞t0→−∞

S(t, t0), (5.3)

51

Page 99: Mathematical Foundations of Quantum Field Theory

March 26, 2020 14:41 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch05 page 52

52 Mathematical Foundations of Quantum Field Theory

where

S(t, t0) = exp(iH0t) exp(−iH(t− t0)) exp(−iH0t0).

However, the S-matrix can sometimes be unitary when the Møller

matrices are not unitary. Namely, for the S-matrix to be unitary, it is

sufficient (and necessary) to assume that the ranges of the operators

S− and S+ coincide: S−R = S+R.

In what follows, we will see that, under certain conditions, one

can use the S-matrix to describe the process of scattering; in these

conditions, it is usually possible to prove that the S-matrix is unitary.

Let us assume that the operators H and H0 have the same

domain; we will denote by the symbol V the difference H − H0.

Let us prove a simple sufficient condition for the existence of Møller

matrices (Cook’s condition).

If the integral∫∞

0 ‖V exp(−iH0t)x‖dt converges for all vectors x

in a dense set T ⊂ R, then the Møller matrices S+ and S− exist.

Proof. Consider the vector Φx(t) = exp(iHt) exp(−iH0t)x. Note

that∥∥∥∥dΦx(t)

dt

∥∥∥∥ = ‖i exp(iHt)V exp(−iH0t)x‖ = ‖V exp(−iH0t)x‖.

Hence, we have that

‖Φx(t1)− Φx(t2)‖ =

∥∥∥∥∫ t2

t1

dΦx(t)

dtdt

∥∥∥∥ ≤ ∫ t2

t1

∥∥∥∥dΦx(t)

dt

∥∥∥∥ dt=

∫ t2

t1

‖V exp(−iH0t)x‖dt

and therefore, it follows from our assumptions that for x ∈ T , we

have

limt1,t2→±∞

‖Φx(t1)− Φx(t2)‖ = 0.

This means that the limits of Φx(t) for t→ ±∞ exist for x ∈ T and

therefore, the limits exist for all x ∈ R (see Appendix A.5).

Let us consider the relation of Møller matrices S± and the

S-matrices S to the adiabatic Møller matrices Sα± and the adiabatic

Page 100: Mathematical Foundations of Quantum Field Theory

March 26, 2020 14:41 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch05 page 53

The Theory of Potential Scattering 53

S-matrices Sα defined in Section 4.1. Let us assume that for any t

and t0,

slimα→0

Sα(t, t0) = S(t, t0).

It is easy to check that this condition is satisfied if the operator V

is bounded (in this case, we even have convergence in norm); the

condition is also satisfied in a much larger class of situations.

We will now show that the Cook condition implies

slimα→0

Sα+ = S+, (5.4)

slimα→0

Sα− = S−. (5.5)

If the operator S+ is unitary, then the relations (5.4) and (5.5) imply

slimα→0

Sα = S. (5.6)

To prove these statements, let us consider the vectors

Φαx(t) = Sα(0, t)x = S∗α(t, 0)x,

where x ∈ T . It is clear that∥∥∥∥dΦαx(t)

dt

∥∥∥∥ = ‖iUα(0, t)V exp(−α|t|) exp(−iH0t)x‖

= exp(−α|t|)‖V exp(−iH0t)x‖,

from which it follows that

‖Φαx(t1)− Φα

x(t2)‖ ≤∫ t2

t1

‖V exp(−iH0t)x‖ exp(−α|t|)dt

≤∫ t2

t1

‖V exp(−iH0t)x‖dt.

It follows from this inequality that the limits Φαx(±∞) exist, hence

the operators Sα± also exist. Furthermore, it follows from this

Page 101: Mathematical Foundations of Quantum Field Theory

March 26, 2020 14:41 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch05 page 54

54 Mathematical Foundations of Quantum Field Theory

inequality that the limit

limt→±∞

Φαx(t) = Φα

x(±∞)

is uniform in α; this allows us to take the limit α→ 0 under the limit

sign. Taking this limit, we obtain, for x ∈ T , the relation

limα→0

Sα±x = limα→0

limt→±∞

Φαx(t) = lim

t→±∞limα→0

Sα(0, t)x

= limt→±∞

S(0, t)x = S±x,

implying that the strong limit of Sα± for α→ 0 equals S±.

If V is a bounded operator, then

‖Φαx(t1)− Φα

x(t2)‖ ≤ ‖V ‖ · ‖x‖ ·∣∣∣∣∫ t2

t1

exp(−α|t|)dt∣∣∣∣.

This inequality implies that in the relations (5.4)–(5.6), one can

talk about norm convergence (instead of strong convergence), hence

the operators Sα+, Sα−, Sα are unitary.

Taking the limit of t to ±∞ in the identity,

exp(iHτ)S(0, t) = S(0, t+ τ) exp(iH0τ),

we obtain the important relation

exp(iHτ)S± = S± exp(iH0τ),

which implies

HS± = S±H0, H0S = SH0. (5.7)

If S± are unitary operators, then this relation establishes the unitary

equivalence between H and H0. It follows that if, for example, H0

does not have a discrete spectrum but H does, then the operators

S+ and S− are not unitary (it is easy to check that the ranges of the

operators S+ and S− are orthogonal to eigenvectors in the discrete

spectrum).

Let φλ be a complete system of δ-normalized, generalized eigen-

functions of the operator H0 and Eλ the corresponding energy levels

(for simplicity of notation, let us assume that the operator H0 has no

discrete spectrum). Then the functions ψ±λ = S±φλ are generalized

Page 102: Mathematical Foundations of Quantum Field Theory

March 26, 2020 14:41 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch05 page 55

The Theory of Potential Scattering 55

eigenfunctions of the operator H with the same eigenvalues Eλ. This

follows from the relation

Hψ±λ = HS±φλ = S±H0φλ = EλS±φλ = Eλψ±λ .

The matrix elements of S-matrices in the basis φλ can be easily

expressed in terms of the functions ψ±λ , namely,

〈Sφλ, φµ〉 =⟨S∗+S−φλ, φµ

⟩= 〈S−φλ, S+φµ〉 =

⟨ψ−λ , ψ

⟩.

Therefore, it is convenient to obtain the equations for the functions

ψ±λ . With this goal in mind, consider the operator

Σ±ε = ±ε∫ ±∞

0exp(−ε|t|)S(0, t)dt.

It is easy to check that in the case when the operators S+ and S−exist, we have

limε→+0

Σ±ε = S±.

For the proof, it is sufficient to use the relation

limε→+0

±ε∫ ±∞

0exp(−ε|t|)f(t)dt = f(±∞), (5.8)

which holds if the vector function f(t) is bounded and has the limit

f(±∞) = limt→±∞ f(t).1 It is clear that ψ±λ = limε→+0 ψ±ελ , where

ψ±ελ = Σ±εφλ = ±ε∫ ±∞

0exp(−ε|t|) exp(iHt) exp(−iEλt)φλdt

=±iε

H − Eλ ± iεφλ.

1If ‖f(t)‖ ≤ A and for t ≥ T , we have that ‖f(t)− f(+∞)‖ ≤ δ, then∥∥∥∥ε ∫ ∞

0

exp(−εt)f(t)dt− f(+∞)

∥∥∥∥ = ε

∥∥∥∥∫ ∞

0

exp(−εt)(f(t)− f(+∞))dt

∥∥∥∥≤ 2εTA+ δ.

Page 103: Mathematical Foundations of Quantum Field Theory

March 26, 2020 14:41 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch05 page 56

56 Mathematical Foundations of Quantum Field Theory

We have obtained that ψ±ελ satisfies the equation

(H − Eλ ± iε)ψ±ελ = ±iεφλ,

which can be rewritten in the form

φ±ελ = φλ + (Eλ −H0 ∓ iε)−1V ψ±ελ . (5.9)

If the limit limε→0(Eλ −H0 ± iε)−1 = (Eλ −H0 ± i0)−1 exists in an

appropriate sense, then one can take the limit ε→ 0 in equation (5.9);

we obtain the following equation for ψ±λ :

ψ±λ = φλ + (Eλ −H0 ∓ i0)−1V ψ±λ .

This equation is called the Lippman–Schwinger equation.2 One can

also express the matrix elements of the S-matrix in terms of the

functions ψ+λ only (or of the functions ψ−λ only), namely,

〈Sφλ, φµ〉 =⟨ψ−λ , ψ

⟩= δ(λ− µ)− 2πiδ(Eλ − Eµ)

⟨φλ, V ψ

⟩= δ(λ− µ)− 2πiδ(Eλ − Eµ)

⟨V ψ−λ , φµ

⟩. (5.10)

Indeed, using the Lippman–Schwinger equation, we can show that⟨ψ−λ , ψ

⟩= 〈φλ, φµ〉+

⟨(Eλ −H0 + i0)−1V ψ−λ , φµ

⟩+⟨φλ, (Eµ −H0 − i0)−1V ψ+

µ

⟩+⟨V ψ−λ , (Eλ −H0 − i0)−1(Eµ −H0 − i0)−1V ψ+

µ

⟩= δ(λ− µ) +

⟨V ψ−λ , (Eλ − iH0 − i0)−1φµ

⟩+⟨(Eµ −H0 + i0)−1φλ, V ψ

⟩− 1

Eλ − Eµ(⟨ψ−λ − φλ, V ψ

⟩−⟨V ψ−λ , ψ

+µ − φµ

⟩)

= δ(λ− µ) +

(1

Eλ − Eµ + i0− 1

Eλ − Eµ

)⟨V ψ−λ , φµ

⟩+

(1

Eµ − Eλ + i0+

1

Eλ − Eµ

)⟨φλ, V ψ

⟩. (5.11)

2We will not go further into the delicate question of specifying the precisemeaning of the Lippman–Schwinger equation. The calculations at the end of thissection are also informal.

Page 104: Mathematical Foundations of Quantum Field Theory

March 26, 2020 14:41 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch05 page 57

The Theory of Potential Scattering 57

In the above, we used the formula

(Eλ −H0 − i0)−1(Eµ −H0 − i0)−1 = − 1

Eλ − Eµ× [(Eλ −H0 − i0)−1 − (Eµ −H0 − i0)−1]. (5.12)

Both parts of this formula can be seen as generalized operator

functions of λ and µ. The numerical generalized function 1Eλ−Eµ can

be understood as one of the functions (Eλ −Eµ + i0)−1, (Eλ −Eµ −i0)−1 or P 1

Eλ−Eµ (all these functions coincide up to a summand of

the form cδ(Eλ − Eµ)); formula (5.12) is correct in all cases, since

δ(Eλ − Eµ)[(Eλ −H0 − i0)−1 − (Eµ −H0 − i0)−1] = 0).

Replacing 1Eλ−Eµ with 1

Eλ−Eµ+i0 or 1Eλ−Eµ−i0 in (5.11), we obtain

the formula we wanted to prove.

5.2 Single-particle scattering

Let us consider the scattering of non-relativistic particles without

spin in the potential field V(x) that decays to zero at infinity. The

state space R in this case is the space L2(E3) and the Hamiltonian

has the form H = p2/2m+V(x). When the particle is very far from

the scattering center, we can describe the motion of the particle with

the Hamiltonian H0 = p2/2m, assuming that the potential energy is

equal to zero. Accordingly, the scattering of the particle in the field

V(x) can be described by the scattering matrix S corresponding to

the pair of operators (H,H0).

Let us check that in the case when the potential V(x) is square-

integrable, the Cook condition is satisfied, hence the Møller matrices

exist. Note that for a dense set of functions f ∈ R, we have the

inequality |ft(x)| ≤ C|t|−3/2, where ft = exp(−iH0t)f (for the

proof of this inequality, see Section 10.3, Lemma 10.5). This implies

that the norm of the function ψt = V exp(−iH0t)f does not exceed

C|t|−3/2√∫|V(x)|2dx, since ψt(x) = V(x)ft(x). This shows that the

Cook condition is satisfied. One can also prove this condition with

weaker assumptions.

Page 105: Mathematical Foundations of Quantum Field Theory

March 26, 2020 14:41 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch05 page 58

58 Mathematical Foundations of Quantum Field Theory

If the potential V(x) is simultaneously square and absolutely

integrable (V ∈ L2 ∩ L1), then one can show that the S-matrix is

unitary. We will not give the proof of this non-trivial fact. We will

only note that the unitarity of the S-matrix in different assumptions

was considered in numerous papers. The proof of the theorem in the

above formulation is given by Kuroda (1959).

Let us introduce the notation

S(p,q) = 〈Sφq, φp〉,

where φp = (2π)−3/2 exp(ipx) is a generalized eigenfunction of the

momentum operator (S(p,q) can be considered as a kernel of the

operator S in the momentum representation).

Since the operator S commutes with the operator H0, the function

S(p,q) takes the form

S(p,q) = S1(p,q)δ(p2 − q2) =S1(p,q)

2pδ(p− q).

The expression

dσ = π2|S1(k,p)|2dω, (5.13)

where dω = sin θdθdφ is the element of solid angle and k is a

vector of the same length as p and directed at the angle dω has the

physical meaning of differential cross-section, it gives the number of

outgoing particles in the solid angle dω under the condition that the

incoming particles have momentum p (we assume that the beam of

incoming particles has unit flux).3

By the formula (5.10),

S(k,p) = 〈Sφp, φk〉 = δ(k− p)− 2πiδ

(p2

2m− k2

2m

)⟨V ψ−p , φk

⟩,

and therefore the formula for the differential cross-section can be

written in the form

dσ = 16m2π4|⟨V ψ−p , φk

⟩|2dω.

To understand the physical meaning of the quantity (5.13), let

us assume that the incoming particles are described by the wave

3We consider only the scattering on non-zero angles.

Page 106: Mathematical Foundations of Quantum Field Theory

March 26, 2020 14:41 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch05 page 59

The Theory of Potential Scattering 59

function exp(−iH0t)φ with t → −∞ (in other words, we assume

that the particles are described by the function φ at time t = 0 if the

interaction with the potential field is neglected). From the definition

of the S-matrix, it follows that for t → +∞, the wave function has

the form exp(−iH0t)Sφ. In the momentum representation, we have

(Sφ)(p) = ψ(p) =

∫S(p,q)φ(q)dq,

(exp(−iH0t)Sφ)(p) = exp

(−i p

2

2mt

)ψ(p),

hence the probability of having the momentum of outgoing particles

directed in the solid angle Ω is equal to∫Ω|ψ(p)|2dp =

∫Ωdp

∫dqdq′S(p,q)S(p,q′)φ(q)φ(q′).

In actual scattering experiments, we never know the wave function

of the incoming particle. In classical mechanics, this means that we

know the initial momentum p0 of the incoming particle, however, we

do not know its impact parameter (recall that the impact parameter

is the distance of the particle trajectory from the scattering center

in the case when we neglect the interaction and assume that the

trajectory is a straight line). Therefore, in classical mechanics, one

considers a particle beam with particles having the same initial

momentum p0 but different impact parameters; we assume that the

beam has a unit flux (i.e. the number of particles going through a unit

area orthogonal to p0 in unit time is equal to one). The differential

cross-section in the solid angle dΩ is the number of outgoing particles

that are directed into the solid angle dΩ.

In quantum mechanics, the notion of the impact parameter cannot

be defined. However, we can consider a family of wave functions

ρa(p) = α(p) exp(ipa),

where α(p) is a normalized wave function that doesn’t vanish only

in a small neighborhood of p0 and the vector a is orthogonal

to p0. Then one can say that the vector a is the analog of the

impact parameter (recall that multiplication by exp(ipa) in the

Page 107: Mathematical Foundations of Quantum Field Theory

March 26, 2020 14:41 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch05 page 60

60 Mathematical Foundations of Quantum Field Theory

momentum representation is equivalent to a shift by the vector a

in the coordinate representation). The scattering cross-section in the

angle Ω of the incoming particle with initial momentum p0 can be

defined as

σΩ =

∫a⊥p0

da

∫Ωdp|ψa(p)|2,

where ψa = Sρa. Using

ψa(p) =

∫S(p,q)ρa(q)dq =

∫S1(p,q)

δ(p− q)

2pα(q) exp(iqa)dq,

we can say that

σΩ =

∫a⊥p0

da

∫Ωdp

∫dqdq′S1(p,q)S1(p,q′)

× δ(p− q)

2p· δ(p− q′)

2pα(q)α(q′) exp(i(q− q′)a),

Integrating over a, we obtain

σΩ = (2π)2

∫Ωdp

∫dqdq′S1(p,q)S1(p,q′)

δ(p− q)

4p2

×α(q)α(q′)δ(q− q′)δ(qT − q′T );

here qT , q′T are projections of the vectors q and q′ onto the plane

orthogonal to the vector p0; we used the fact that∫a⊥p0

exp(i(q− q′)a)da = (2π)2δ(qT − q′T )].

It is easy to check that

δ(q − q′)δ(qT − qT ′) = 2qδ(q2 − q′2)δ(qT − qT ′)

= 2qδ(q2n − q′2n )δ(qT − q′T )

=q

qnδ(qn − q′n)δ(qT − q′T )

+q

qnδ(qn + q′n)δ(qT − q′T )

=q

qn[δ(q− q′) + δ(Iq− q′)],

where qn, q′n are the projections of the vectors q, q′ on the vector p0

and I is the symmetry with respect to the plane orthogonal to p0.

Page 108: Mathematical Foundations of Quantum Field Theory

March 26, 2020 14:41 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch05 page 61

The Theory of Potential Scattering 61

Using the δ-functions entering the integrand, we can do the integral

over dq′ and using spherical coordinates in the integral over dp, we

can instead integrate over dp. We obtain

σΩ = π2

∫ω

sin θdθdφ

∫dq|S1(q, θ, φ|q)|2 q

qn|α(q)|2

+π2

∫ω

sin θdθdφ

∫dqS1(q, θ, φ|q)S1(q, θ, φ|Iq)

×(− q

qn

)α(q)α(Iq)

(when p runs over Ω, then the spherical coordinates θ, φ run over

ω; if p is a vector with spherical coordinates (p, θ, φ), then we use

the notation S1(p,q) = S1(p, θ, φ|q)). If the function S1(p,q) is

continuous when q changes in a neighborhood of the point p0 and p

changes in the angle Ω, we obtain that

σΩ ≈ π2

∫ω|S1(p0, θ, φ|p0)|2 sin θdθdφ.

This agrees with the expression for the differential cross-section that

was written earlier. We have used that for a normalized function α(q)

with support in a small neighborhood of the point p0 and continuous

function f(q), we have that∫f(q)|α(q)|2dq ≈ f(p0).

If p0 6= 0, then ∫f(q)α(q)α(Iq)dq ≈ 0.4

4It is not difficult to convert the above considerations into a rigorous proof.To be precise, one needs to consider a sequence of normalized functions αn withsupport tending towards the point p0 6= 0 and use the fact that for such a sequenceand a continuous function f(q), we have∫

f(q)|αn(q)|2dq→ f(p0),∫f(q)αn(q)αn(Iq)dq→ 0.

Page 109: Mathematical Foundations of Quantum Field Theory

March 26, 2020 14:41 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch05 page 62

62 Mathematical Foundations of Quantum Field Theory

Let us write down the Lippman–Schwinger equations in this case.

For the full system of generalized eigenfunctions of the Hamiltonian

H0, we will take the unctions φp(x) = (2π)−3/2 exp(ipx). In the

momentum representation, the Hamiltonian H0 can be considered as

multiplication by p2

2m , the operator (Eq−H0± iε)−1 can be identified

with multiplication by ( q2

2m −p2

2m ± iε)−1, and the operator (Eq −H0 ± i0)−1 can be identified with multiplication by the generalized

function ( q2

2m −p2

2m ± i0)−1. Hence, the momentum representation of

the Lippman–Schwinger equation takes the form

ψ±q (p) = δ(p− q) +

∫V(p− p′)ψ±q (p′)dp′

q2

2m −p2

2m ∓ i0, (5.14)

where

V(p) = (2π)−3/2

∫V(x) exp(−ipx)dx.

Taking into account that∫exp(ipx)dpq2

2m −p2

2m ± i0= −4mπ2 exp(±iqx)

x,

we can write the Lippman–Schwinger equation in the coordinate

representation

ψ±q (x) = φq(x)− m

∫dx′

exp(∓iq|x− x′|)|x− x′|

V(x′)ψ±q (x′). (5.15)

Let us calculate the asymptotic behavior of the function ψ±q (x) for

x → ∞. We have assumed that the potential V(x) tends to zero

sufficiently fast at infinity, hence we can assume that in the integral

in the formula (5.15), we have x x′, and hence in the exponential,

we can replace the |x−x′| in the numerator with x− xxx′ and |x−x′|

in the denominator with x. We obtain that for large x,

ψ±(x) ≈ φq(x) + f±q

(∓xx

) exp(∓iqx)

x,

where

f±q (e) = −m2π

∫dx′ exp(−iqex′)V(x′)ψ±q (x′).

Page 110: Mathematical Foundations of Quantum Field Theory

March 26, 2020 14:41 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch05 page 63

The Theory of Potential Scattering 63

Hence, it follows that

f±q (e) = −m√

2π⟨V ψ±q , φk

⟩, (5.16)

where k = qe. Formula (5.16) together with the relations (5.10)

and (5.13) allows us to express the matrix entries of the S-matrix

and the differential cross-section in terms of the function f−q (e) that

is specified by the asymptotic behavior at infinity of the function

ψ−q (x). For example, the scattering cross-section in the solid angle

dω is equal to

dσ = (2π)3|f−p (e)|2dω; (5.17)

here p is the momentum of the incoming particles and e is a unit

vector directed in the solid angle dω. The function ψ+q (x) satisfies

the equation (− 1

2m∆ + V(x)

)ψ±q (x) =

q2

2mψ±q (x) (5.18)

and the boundary condition at infinity

ψ±q (x) ≈ φq(x) + f±q

(∓xx

) exp(∓iqx)

x. (5.19)

It is easy to check that these two conditions characterize the

function ψ±q (x). In other words, the function ψ−q (x) can be described

as the solution of the stationary Schrodinger equation that can be

represented at infinity as a sum of a plane wave and an outgoing

spherical wave. The function ψ+q (x) can be described similarly, but

the outgoing spherical wave should be replaced by an incoming

spherical wave.

Hence, we come to the stationary formulation of the scattering

problem for non-relativistic particles. In this formulation, one can

find the scattering cross-section by solving the stationary Schrodinger

equation with the boundary condition (5.19). It is useful to note that

in the case that V(x) = V(−x),

ψ+q (x) = ψ−q (−x)

(this follows from the fact that the complex conjugate of the incoming

spherical wave is the outgoing spherical wave and equation (5.18) is

Page 111: Mathematical Foundations of Quantum Field Theory

March 26, 2020 14:41 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch05 page 64

64 Mathematical Foundations of Quantum Field Theory

invariant with respect to complex conjugation). Therefore, f+q (e) =

f−q (e), and hence the scattering cross-section can be written in the

form

dσ = (2π)3|f+p (e)|2dω.

5.3 Multi-particle scattering

The multi-particle scattering problem can be considered on the Fock

space F , corresponding to the measure space E3 × B, where B is a

finite set. The momentum operator P on this space is defined by the

formula (3.15). Let us consider a translation-invariant HamiltonianH

on the space F (i.e. a self-adjoint operator H, commuting with the

momentum operator P).

The definition of the scattering matrix given in Section 5.1 is

not always applicable to the problem of scattering in a system of

n particles. This is because composite particles can be generated

in the process of collision. For example, if we start with protons

and neutrons, then it is possible that after collision we will have a

deuteron (a bound state of a proton and neutron) or an α-particle

(a bound state of two protons and two neutrons). Therefore, we

should first give a general definition of a particle that includes the

objects we will call composite particles.

The vector Φ(k) that describes a single-particle state with

momentum k should be an eigenvector of the operators P and H

(in the single-particle state, having the momentum k, the energy

has definite value ω(k)). However, the operator P has only one

normalized eigenvector (the Fock vacuum), therefore, the vector

function Φ(k) should be a generalized function. This consideration

allows us to give the following definition of a particle.

We define a particle corresponding to the Hamiltonian H, as a

generalized vector function Φ(k), obeying

HΦ(k) = ω(k)Φ(k), (5.20)

PΦ(k) = kΦ(k), (5.21)⟨Φ(k),Φ(k′)

⟩= δ(k− k′). (5.22)

Page 112: Mathematical Foundations of Quantum Field Theory

March 26, 2020 14:41 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch05 page 65

The Theory of Potential Scattering 65

The function ω(k) is called the energy of a single-particle state or

the dispersion law.

For example, the particles corresponding to the Hamiltonian

H0 =

n∑s=1

∫εs(k)a+

s (k)as(k)dk (5.23)

are generalized vector functions

Φs(k) = a+s (k)θ. (5.24)

The same generalized vector functions can also be considered as

particles corresponding to the Hamiltonian

H = H0 +W, (5.25)

where

W =∑

m≥2,n≥2

∫Wm,n(k1, . . . ,km|l1, . . . , ln)a+(k1)

. . . a+(km)a(l1) . . . a(ln)dmkdnl. (5.26)

These particles are called elementary particles. For translation-

invariant Hamiltonians of the form H0 + V , where V is an arbitrary

self-adjoint operator, a particle is called an elementary particle if

it can be obtained from the particle a+s (k)θ, corresponding to the

Hamiltonian H0 by means of perturbation theory (in other words,

if the Hamiltonian H0 + gV has a particle Ψg(k) that depends

continuously on the parameter g in the interval [0, 1], and we also

have Ψ0(k) = a+s (k)θ, then the particle Ψ1(k) is called an elementary

particle of the Hamiltonian H0 + V ).

The particles Φ(k),Φ′(k) are orthogonal if 〈Φ(k),Φ′(k)〉 = 0; the

system of particles Φ1(k), . . . ,ΦN (k) is called complete if there is

no particle orthogonal to every particle Φi(k).

For example, for the Hamiltonian, H0 =∫ε(k)a+(k)a(k)dk,

where ε(k) is a strongly convex function, the system of particles

consisting of a single particle Φ(k) = a+(k)θ (elementary particle),

Page 113: Mathematical Foundations of Quantum Field Theory

March 26, 2020 14:41 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch05 page 66

66 Mathematical Foundations of Quantum Field Theory

is complete. To prove this, let us consider a particle

Ψ(k) =∑n

∫ψn(k,k1, . . . ,kn)a+(k1) . . . a+(kn)θ

corresponding to the Hamiltonian H0. From the condition PΨ(k) =

kΨ(k), it follows that

(k1 + · · ·+ kn)ψn(k,k1, . . . ,kn) = kψn(k,k1, . . . ,kn),

and hence

ψn(k,k1, . . . ,kn) = δ(k− k1 − · · · − kn)φn(k1, . . . ,kn).

The condition HΨ(k) = ω(k)Ψ(k) gives the relation

(ε(k1) + · · ·+ ε(kn))ψn(k,k1, . . . ,kn) = ω(k)ψn(k,k1, . . . ,kn),

from which it is clear that

(ω(k1 + · · ·+ kn)− ε(k1)− · · · − ε(kn))φn(k1, . . . ,kn) = 0.

(5.27)

Equation (5.27) allows us to conclude that φn ≡ 0 for n > 1 (to

prove this, we should note that a strongly convex function cannot be

constant on a set of positive measure and that the function ε(p) +

ε(k − p) is a strongly convex function of the variable p). Hence,

Ψ(k) = φ1(k)a+(k)θ; this concludes the proof.

The complete system of particles does not always consist only

of elementary particles. It is possible that there exist particles that

are orthogonal to every elementary particle (such particles are called

composite particles).

We will consider, for example, the Hamiltonian

H =

∫k2

2ma+(k)a(k)dk +

1

2

∫W(k1 − k3)a+(k1)

× a+(k2)a(k3)a(k4)δ(k1 + k2 − k3 − k4)dk1dk2dk3dk4

(5.28)

that describes a system of identical spinless bosons (see Section 3.2).

This Hamiltonian commutes with N (the operator of number of

Page 114: Mathematical Foundations of Quantum Field Theory

March 26, 2020 14:41 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch05 page 67

The Theory of Potential Scattering 67

particles), therefore, it is sufficient to find only particles that belong

to the n-particle state Fn. Recall that Fn can be represented as the

space of symmetric functions φ(k1, . . . ,kn), where ki ∈ E3. Let us

introduce new variables

p =k1 + · · ·+ kn

n,

pi = ki − kn (i = 1, . . . , n− 1).

The space F ′n will be defined as the space of functions ψ(p1, . . . ,pn−1)

obeying the condition that the function f(p)ψ(p1, . . . ,pn−1), where

f(p) ∈ L2(E3), belongs to the space Fn. Functions from the space F ′ncan be considered as wave functions of relative motion of n particles.

The consideration of motion of n particles in terms of functions from

the space F ′n corresponds to the separation of the motion of the center

of inertia. Mathematically, the possibility to separate the motion of

the center of inertia for the Hamiltonian Hn means that there exists

a Hamiltonian H ′n in the space F ′n, that for every function,

f(p)ψ(p1, . . . ,pn−1) ∈ Fn,

where f ∈ L2(E3), ψ ∈ Fn, we have

Hn(f(p)ψ(p1, . . . ,pn−1)) =

(p2

2nmf(p)

)ψ(p1, . . . ,pn−1)

+ f(p)(H ′nψ)(p1, . . . ,pn−1).

Let us assume that ψ ∈ F ′n is a normalized eigenvector of

the operator H ′n with eigenvalue E. It is easy to check that the

generalized vector function

Φ(k) = δ(k− p)ψ(p1, . . . ,pn−1)

is a particle in the sense of the above definition. It is also easy to

check that

HΦ(k) =

(k2

2nm+ E

)Φ(k).

For n > 1, the particle Φ(k) is a composite particle (a bound state

of a system of n particles).

Page 115: Mathematical Foundations of Quantum Field Theory

March 26, 2020 14:41 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch05 page 68

68 Mathematical Foundations of Quantum Field Theory

Let us now ask what we should consider to be a scattering matrix

for a Hamiltonian of the form (5.25) (for more general translation-

invariant Hamiltonians, the definition of the scattering matrix will

be given in Chapter 9). It is natural to try to define the scattering

matrix of the Hamiltonian (5.25) as an S-matrix corresponding to the

pair of operators (H,H0). However, this works only in the case when

elementary particles Φs(k) = a+s (k)θ constitute a complete system of

particles (in other words, there exist no composite particles). From

the physical viewpoint, it is clear that the S-matrix of the operators

(H,H0) cannot describe the scattering of composite particles. If

composite particles are present, then the S-matrix corresponding to

the pair of operators (H,H0) will not be a unitary operator.

To formulate the definition of scattering matrix, we will restrict

ourselves to the case when all particles are bosons.

Note, first of all, that every particle Φ(k) can be written in the

form

Φ(k) =

∞∑n=1

∑i1,...,in

∫δ(k− k1 − · · · − kn)

×φn(k1, i1, . . . ,kn, in)a+i1

(k1) . . . a+in

(kn)θdk1 . . . dkn.

Let us assign to the particle Φ(k) an operator generalized function

defined by the formula

A(k) =∞∑n=1

∑i1,...,in

∫δ(k− k1 − · · · − kn)

×φn(k1, i1, . . . ,kn, in)ai1(k1) . . . ain(kn)θdk1 . . . dkn.

The operator generalized function A(k) satisfies the conditions:

(1) A+(k)θ = Φ(k) (the operator A+(k) creates the particle Φ(k)

when applied to the vacuum) and, (2) A(k) is a superposition of

operators ai1(k1) . . . ain(kn), i.e. it can be represented in the form

A(k) =

∞∑n=1

∑i1,...,in

∫λn(k,k1, i1, . . . ,kn, in)

× ai1(k1) . . . ain(kn)dk1 . . . dkn.

Moreover, these conditions specify A(k) uniquely.

Page 116: Mathematical Foundations of Quantum Field Theory

March 26, 2020 14:41 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch05 page 69

The Theory of Potential Scattering 69

Let us now define the in- and out-operators Ain(k, τ ) and

Aout(k, τ ), corresponding to the particle Φ(k), as the limits

Ain(k, τ ) = limt→−∞

exp(iω(k)(t− τ))A(k, t), (5.29)

Aout(k, τ ) = limt→+∞

exp(iω(k)(t− τ))A(k, t), (5.30)

where A(k, t) = exp(iHt)A(k) exp(−iHt) and ω(k) is the dispersion

law for the particle Φ(k). The limit is understood in the sense of the

limit of generalized functions (in other words, we assume that for

every function f(k) ∈ S(E3), we have∫f(k)A in

out(k, τ)dk = lim

t→∓∞

∫f(k) exp(iω(k)(t− τ))A(k, t)dk

in the sense of strong operator limit on the linear subspace D, where

D is defined as the smallest linear subspace containing all spaces Fn).

When talking about the operators Ain and Aout at the same time,

we will use the notation Aex.

Under certain conditions on the Hamiltonian H (in particular,

for the Hamiltonians of the form (5.28), under the assumption that

interaction potential is square integrable), we can prove that

(1) the limits (5.29) and (5.30) exist;

(2) if the particles Φ1(k), . . . ,Φm(k) are orthogonal, then the corre-

sponding operators Aex(k, i, τ), A+ex(k, i, τ), where i = 1, . . . ,m,

for fixed τ obey the canonical commutation relations (CCR)

[A+ex(k, i, τ), A+

ex(k′, i′, τ)] = [Aex(k, i, τ), Aex(k′, i′, τ)] = 0;

[Aex(k, i, τ), A+ex(k′, i′, τ)] = δiiδ(k− k′);

(3) Aex(k, i, τ) = exp(iHτ)Aex(k, i) exp(−iHτ) = exp(−ωi(k)τ)

Aex(k, i), where Aex(k, i) = Aex(k, i, 0);

(4) Aex(k, i)θ = 0, A+ex(k, i)θ = Φi(k).

The proof of these statements can be found, for example, in Hepp

and Epstein (1971) (the last two statements are trivial).

The generalized vector functions

Ψin(k1, i1, . . . ,kn, in) = A+in(k1, i1) . . . A+

in(kn, in)θ,

Ψout(k1, i1, . . . ,kn, in) = A+out(k1, i1) . . . A+

out(kn, in)θ

are called in- and out-states.

Page 117: Mathematical Foundations of Quantum Field Theory

March 26, 2020 14:41 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch05 page 70

70 Mathematical Foundations of Quantum Field Theory

Sometimes it is useful to note that the vectors Ψin and Ψout can

also be represented in the form

Ψex(k1, i1, . . . ,kn, in) = limt→±∞

exp(−i(ωi1(k1) + · · ·+ ωin(kn))t)

×A+(k1, i1, t) . . . A+(kn, in, t)θ. (5.31)

This follows immediately from the definition of the operators A+in

and A+out. Applying the formula (5.8), we can get from (5.31) the

following representation of in- and out-states:

Ψ inout

(k1, i1, . . . ,kn, in) = ± limα→0

iα(H − (ωi1(k1)

+ · · ·+ ωin(kn))± iα)−1A+(k1, i1) . . . A+(kn, in)θ.

Let us fix a complete orthonormal system of particles Φ1(k),

. . . ,ΦN (k) and the operators Aex(k, 1), . . . , Aex(k, N) corresponding

to the particles of this system.

We define the matrix elements of the S-matrix (or scattering

amplitudes) as functions

Sm,n(k1, i1, . . . ,km, im|l1, j1, . . . , ln, jn)

=⟨A+

in(l1, j1) . . . A+in(ln, jn)θ,A+

out(k1, i1) . . . A+out(km, im)θ

⟩.

Knowing these matrix elements, we can express the probability that

by collision of n particles with quantum numbers l1, j1, . . . , ln, jn, we

get particles with quantum numbers k1, i1, . . . ,km, im (more general

situations will be discussed in Chapter 10).

Let us consider the relation of the definition of the scattering

matrix in terms of the in- and out-operators, with the definition of the

S-matrix given in Section 5.1. Let us suppose that the Møller matrices

S∓, corresponding to the pair of operators (H,H0) are unitary (this

assumption is always satisfied when there are no composite particles).

The operators A(k, s) corresponding to elementary particles Φs(k) =

a+s (k)θ are clearly equal to as(k). Let us show that

Ain(k, s) = ain(k, s) = S−as(k)S∗−, (5.32)

Aout(k, s) = aout(k, s) = S+as(k)S∗+. (5.33)

Page 118: Mathematical Foundations of Quantum Field Theory

March 26, 2020 14:41 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch05 page 71

The Theory of Potential Scattering 71

To check these equalities, we will use the relations (5.1) and (5.2). It

follows from these relations that

S∓as(k)S+∓ = slim

t→∓∞exp(iHt) exp(−iH0t)as(k)

× exp(iH0t) exp(−iHt)

= slimt→∓∞

exp(iεs(k)t)as(k, t) = a inout

(k, s).

The relations (5.32) and (5.33) imply that⟨a+

in(l1, σ1) . . . a+in(ln, σn)θ, a+

out(k1, s1) . . . a+out(km, sm)θ

⟩=⟨S−a

+σ1(l1) . . . a+

σn(ln)θ, S+a+s1(k1) . . . a+

sm(km)θ⟩

=⟨Sa+

σ1(l1) . . . a+σn(ln)θ, a+

s1(k1) . . . a+sm(km)θ

⟩(we have used S∗−θ = S∗+θ = θ). In other words, the matrix elements

of the S-matrix defined in the present section under the assumption

we have made coincide with the matrix elements of the operator S

in the generalized basis a+s1(k1) . . . a+

sm(km)θ.

In the case when there exist bound states, we can also construct

an operator S having the functions Sm,n as matrix elements. This

operator S acts on the space Fas that is called the space of asymptotic

states. Fas can be defined as a Fock space constructed from the

measure space E3×N , where N is a set of types of particles. (Let us

recall that we have fixed a complete orthonormal system of particles.)

The operators of creation and annihilation in Fas will be denoted

b+(k, i) and b(k, i) (here, i ∈ N). The vectors from the space Fas can

be considered as initial and final states of the scattering process (for

example, the vector b+i1(f1) . . . b+in(fn)θ, where bi(f) =∫f(k)b(k, i)dk

corresponds to a state with the particles of the types i1, . . . , in with

the wave functions f1, . . . , fn). Let us define the isometries S− and

S+ acting from the space Fas into the space F by means of the

relations

Ain(k, i)S− = S−b(k, i), S−θ = θ, (5.34)

Aout(k, i)S+ = S+b(k, i), S+θ = θ. (5.35)

(Such operators do exist and they are defined by the relations (5.34)

and (5.35) in a unique way; the proof of this fact is given in

Page 119: Mathematical Foundations of Quantum Field Theory

March 26, 2020 14:41 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch05 page 72

72 Mathematical Foundations of Quantum Field Theory

Section 6.1.) The scattering matrix is defined as the operator S =

S∗+S− acting on the space Fas. It is obvious that

Sm,n(k1, i1, . . . ,km, im|l1, j1, . . . , ln, jn)

=⟨A+

in(l1, j1) . . . A+in(ln, jn)θ,A+

out(k1, i1) . . . A+out(km, im)θ

⟩=⟨S−b

+(l1, j1) . . . b+(ln, jn)θ, S+b+(k1, i1) . . . b+(km, im)θ

⟩=⟨Sb+(l1, j1) . . . b+(ln, jn)θ, b+(k1, i1) . . . b+(km, im)θ

⟩.

In other words, the matrix elements of the operator S in the general-

ized basis b+(k1, i1) . . . b+(km, im)θ coincide with the functions Sm,n(the scattering amplitudes).

Hamiltonians of the form (5.28) commute with the operator of

the number of particles, therefore, one can study separately the

n-particle scattering for n = 2, 3, . . . . For n = 2, this problem can

be reduced to the problem of potential scattering considered in the

preceding section. The proof of unitarity of the scattering matrix

for the three-particle problem was obtained by Faddeev (1963).

Faddeev’s method is based on the construction of equations for

the in- and out-states. Faddeev equations are also very useful for

calculations. They were also generalized in different ways for the

case of n-particle scattering problem for n > 3. The unitarity of the

scattering matrix for any number of non-relativistic particles was

proven in Sigal and Soffer (1987), see Hunziker and Sigal (2000) for

review.

In conclusion, let us present the definition of the Møller matrices

S± and the scattering matrix S in the form that is more convenient

in general situations (Chapter 10). Let us say that the operator B is

a good operator if it can be represented as∑m

∑s1,...,sm

∫f(k1, s1, . . . ,km, sm)

× a+s1(k1) . . . a+

sm(km)dk1 . . . dkm

and obeys the condition

Bθ =

∫φ(k)Φi(k)dk,

Page 120: Mathematical Foundations of Quantum Field Theory

March 26, 2020 14:41 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch05 page 73

The Theory of Potential Scattering 73

where Φi(k) is one of the particles in the fixed complete system

Φ1(k), . . . ,Φn(k). The isometric operator S−(S+) acting on the

asymptotic space Fas into the space F is called the Møller matrix if for

all good operators B1, . . . , Bm and smooth functions with compact

support f1(p), . . . , fm(p), we have

limt→∓∞

B1(f1, t) . . . Bm(fm, t)θ

= S∓b+(φ1, f1, i1) . . . b+(φm, fm, im)θ.

Here, the operators Bα(fα, t) are defined by the formula

Bα(fα, t) =

∫fα(x, t)Bα(x, t)dx,

where

Bα(x, t) = exp(iHt− iPx)Bα exp(−iHt+ iPx),

fα(x, t) =

∫exp(−iωiα(p)t+ ipx)fα(p)

dp

(2π)3,

and the numbers iα and the functions φα(p), ωiα(p) are defined by

the relations

Bαθ =

∫φα(p)Φiα(p)dp,

HΦiα(p) = ωiα(p)Φiα(p).

The scattering matrix S, as always, is defined in terms of the Møller

matrices S = S∗+S−.

To check that this definition of Møller matrices is equivalent to the

above definition, we note that a good operator Bα can be represented

in the form

Bα =

∫φα(p)A+

iα(p)dp,

where Ai(p) is an operator generalized function corresponding to the

particle Φi(p). Using the obvious relation

exp(−iPx)A+i (k) exp(iPx) = exp(−ikx)A+

i (k),

Page 121: Mathematical Foundations of Quantum Field Theory

March 26, 2020 14:41 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch05 page 74

74 Mathematical Foundations of Quantum Field Theory

we see that

Bα(fα, t) =

∫fα(x, t) exp(−ikx)φα(k)A+

iα(k, t)dk

=

∫exp(−iωiα(k)t)fα(k)φα(k)A+

iα(k, t)dk,

hence

slimt→∓∞

Bα(fα, t) =

∫fα(k)φα(k)A+

inout

(k, iα)dk. (5.36)

To finish the proof of equivalence of the two definitions of the Møller

matrices, we note that it follows from (5.36), (5.34) and (5.35) that

limt→∓∞

B1(f1, t) . . . Bn(fn, t)θ =

∫f1(k1)φ1(k1)A+

inout

(k1, i1)

. . . fn(kn)φn(kn)A+in

out

(kn, in)dnkθ = S∓b+(f1, φ1, i1)

. . . b+(fnφn, in)θ.

Page 122: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 75

Chapter 6

Operators on Fock Space

6.1 The representations of canonical and

anticommutation relations: Fock representation

Let us assume that we have assigned to every vector f of a pre-Hilbert

space B two conjugate operators a(f) and a+(f) acting on a dense

subspace of the Hilbert space H, in such a way that a(f) linearly

depends on f (i.e. a(λf + µg) = λa(f) + µa(g)). If the operators

satisfy the relations

[a(f), a(g)] = [a+(f), a+(g)] = 0, [a(f), a+(g)] = 〈f, g〉 (6.1)

(where f, g ∈ B and λ, µ are complex numbers), then we say that

these operators specify a representation of canonical commutation

relations, or CCR, on the space H. If the operators satisfy the

analogous relations

[a(f), a(g)]+ = [a+(f), a+(g)]+ = 0, [a(f), a+(g)]+ = 〈f, g〉, (6.2)

then we say that these operators specify a representation of canonical

anticommutation relations or CAR.1

We will also introduce the notation a(f,−1) = a(f), a(f,+1) =

a+(f). Using this notation, we can write down CCR and CAR in the

following form:

[a(f, ε), a(g, ε′)]∓ = Aεε′ 〈f, g〉, (6.3)

1In the definition of the representation of CCR and CAR, one should assumethat the operators a(f), a+(f) are defined on the same linear subspace D that isdense in the space H. They should transform D into itself. The operators a(f)and a+(f) are conjugate (i.e. 〈a(f)x, y〉 =

⟨x, a+(f)y

⟩for all x, y ∈ D).

75

Page 123: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 76

76 Mathematical Foundations of Quantum Field Theory

where Aεε′ is a matrix with the elements A11 = A−1−1 = 0, A−11 =

1, A1−1 = ∓1 (the upper sign is used in the case of CCR and the

lower sign is used in the case of CAR).

In the case of CAR, the operators a(f) and a+(f) are

bounded (the equation a(f)a+(f) + a+(f)a(f) = 〈f, f〉 implies that

〈a(f)x, a(f)x〉 + 〈a+(f)x, a+(f)x〉 = 〈f, f〉 〈x, x〉, hence ‖a(f)‖ ≤‖f‖, ‖a+(f)‖ ≤ ‖f‖). Therefore, one can assume that the operators

a(f), a+(f) are defined on the whole space H. In the case of CCR,

the operators a(f), a+(f) are unbounded.

Speaking simultaneously of CCR and CAR, we will use the term

“canonical relations” (CR). When it is necessary to stress that CR

are constructed by means of a pre-Hilbert space B, we will use the

notation CR(B). If in the space B we have fixed an orthonormal

basis φn, then the operators an = a(φn) and a+n = a+(φn) satisfy the

relations

[am, an] = [a+m, a+n ] = 0, [am, a

+n ] = δm,n (6.4)

in the case of CCR and the relations

[am, an]+ = [a+m, a+n ]+ = 0, [am, a

+n ]+ = δm,n (6.5)

in the case of CAR (these relations are also called CR).

The most important example of operators satisfying CR includes

the operators of creation a+(f) and annihilation a(f), defined in

Section 3.2. (They specify a representation of CCR in the bosonic

case and the representation of CAR in the fermionic case.) The

operators a+(f), a(f) are called creation and annihilation operators

in other cases as well; note, however, that these names don’t always

agree with their physical meaning.

One more important representation of CCR is given by the

following operators:

a(f) =

N∑n=1

fnan,

a+(f) =

N∑n=1

fna+n ,

Page 124: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 77

Operators on Fock Space 77

where a+n , an are operators satisfying the relations (6.4) that were

constructed in Section 2.6 where we studied the system of coupled

oscillators (here f = (f1, . . . , fN ) runs over a complex N -dimensional

space).

The representation of CR is called the Fock representation if one

can find a cyclic vector θ in the space H that satisfies the condition

a(f)θ = 0. More precisely, we assume that the vector θ is a cyclic

vector with respect to the operators a+(f), a(f) (this means that

applying these operators to θ and taking linear combinations we get

a subset of H that is dense in H). The vector θ is called the vacuum

vector2 (see Section 3.2).

Let us prove that two Fock representations are equivalent. This

means that there exists a unitary operator α mapping the space

H1 onto the space H2 and satisfying the conditions αa1(f) =

a2(f)α, αθ1 = θ2 (here ai(f) is a Fock representation of CR in the

spaceHi with the vacuum vector θi). In the case of CCR, the equation

αa1(f) = a2(f)α should be satisfied on a dense subset of the space

H1 (not necessarily on the whole domain of a1(f)).

Let us construct an operator α. For vectors of the form

a1(f1, ε1) . . . a1(fn, εn)θ1, (6.6)

we define the operator α, assuming that

α(a1(f1, ε1) . . . a1(fn, εn)θ1) = a2(f1, ε1) . . . a2(fn, εn)θ2.

If x and y are two vectors of the form (6.6), then 〈x, y〉 = 〈αx, αy〉.This follows from the fact that the scalar product of vectors of

the form (6.6) can be calculated using only CR and the relation

ai(f)θi = 0. Let us denote by S1 the linear subspace consisting

of linear combinations of vectors of the form (6.6). The operator

α, by linearity, can be extended to S1 and satisfies the condition

〈x, y〉 = 〈αx, αy〉 for all x, y ∈ S1. Using the assumption that the

vector θi is cyclic with respect to the operators ai(f, ε), we obtain

2Using CR, it is easy to check that every vector of the form a(f1, ε1) . . . a(fn, εn)θcan be represented as a linear combination of vectors of the forma+(φ1) . . . a+(φm)θ. Hence, θ is also a cyclic vector with respect to the familyof operators a+(f).

Page 125: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 78

78 Mathematical Foundations of Quantum Field Theory

that the set S1 is dense in H2 and the set S2 = αS1 is dense in H2.

The operator α can be extended to the closure S1 = H1 continuously;

the density of the set S2 in H2 implies the relation αH1 = H2. It is

clear that on the set S1 we have αa1(f, ε) = a2(f, ε)α. This completes

the proof of the equivalence of two Fock representations.3

Hence, it is sufficient to describe just one Fock representation; this

was done in Section 3.2 (and will be done again in slightly different

terms later).

The space of Fock representation CR(B) will be called Fock space

and denoted as F (B).

Let us assume that the pre-Hilbert space B is realized as a dense

subset of the space L2(Er) (the space of square-integrable functions

on the Euclidean space Er). Then, together with the operators

a+(f), a(f), it is convenient to introduce the operator generalized

functions a+(x), a(x) on Er, where

a(f) =

∫f(x)a(x)dx, a+(f) =

∫f(x)a+(x)dx.

The functions a+(x), a(x) satisfy the following relations:

[a(x), a(y)]∓ = [a+(x), a+(y)]∓ = 0,

[a(x), a+(y)]∓ = δ(x, y); a(x)θ = 0.(6.7)

This definition of the generalized operator functions a+(x), a(x) can

also be used in the case when B is a dense subset of the space

L2(X), where X = Er × B consists of pairs (e, s) with e ∈ Er a

point in Eucliean space and s ∈ B from a finite set B (then the

integration over X is understood as integration over Eucliean space

and summation over a finite set). We can also use this definition

in the more general case when X is an arbitrary measure space;

however, we do not need this level of generality. Later in this section,

3If the first representation of CR is a Fock representation, but the secondrepresentation has a vector θ2 ∈ R2 satisfying the condition a2(f)θ2 = 0 (notnecessarily cyclic), then precisely in the same way, we can construct the operatorα satisfying the relation αa1(f) = a2(f)α, αθ1 = θ2. However, in general, α will bean isometry that is not necessarily unitary (this statement is used in Section 5.3to construct the operators S− and S+).

Page 126: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 79

Operators on Fock Space 79

we assume that X is any measure space; however, for simplicity, the

reader can assume that X = Er or X = Er ×B.

Operator generalized functions a+(x), a(x), in the case of the Fock

representations of CR, were considered in Section 3.2. In particular,

we have shown in this section that an expression of the form (3.4) can

be understood as an operator acting on Fock space. For an arbitrary

representation of CR, the expression (3.4) can be easily defined in the

case when the function entering this expression has the form (3.5).

If f is an arbitrary function, then it can be represented as a limit of

functions fk of the form (3.5) and the operator A corresponding to

the function f should be defined, in this case, as a limit of operators

Ak corresponding to the functions fk. Of course, this definition of the

operator A is not completely precise because we have not defined the

notion of limit for the functions fk and the operators Ak. We will not

be concerned with these definitions; however, we will note that in the

case when B = S(E3) and the functional 〈a(f, ε)u, v〉 (for arbitrary

u, v ∈ D) continuously depends on f in the topology of the space

S(E3), then using the operator analog of the kernel theorem (see

Appendix A.5), one can define the operator (3.4) for the functions

f(x1, . . . ,xm|y1, . . . ,yn) from the space S(E3(m+n)).

In the case when B = L2(X), every element Φ of Fock space can

be written in a unique way in the form

Φ =∑n

∫fn(x1, . . . , xn)a+(x1) . . . a

+(xn)θdx1 . . . dxn, (6.8)

where the functions fn(x1, . . . , xn) are symmetric in the case of CCR

and antisymmetric in the case of CAR in the arguments x1, . . . , xn ∈X. The norm of the vector Φ is equal to∑

n

n!

∫|fn(x1, . . . , xn)|2dx1 . . . dxn.

Conversely, to every sequence f0, f1(x), . . . , fn(x1, . . . , xn), . . . of

symmetric (antisymmetric) functions obeying∑n

n!

∫|fn(x1, . . . , xn)|2dx1 . . . dxn <∞,

we can assign a vector in Fock space. This description of the Fock

space can be obtained from the formulation described in Section 3.2,

Page 127: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 80

80 Mathematical Foundations of Quantum Field Theory

namely, it is easy to check that the Fock state (Fock column)

Φ =

φ0

φ1(x1)...

φn(x1, . . . , xn)...

can be represented in the form

Φ =∑n

1√n!

∫φn(x1, . . . , xn)a+(x1) . . . a

+(xn)θdx1 . . . dxn.

This statement can also be formulated in slightly different

terms. Namely, one can say that vector generalized functions

a+(x1) . . . a+(xn)θ, where n = 0, 1, 2, . . . , constitute a generalized

basis of the space F (L2(X)) in the following sense: every vector

from F (L2(X)) can be decomposed as a linear combination of these

vector generalized functions (i.e. represented in the form (6.8)) and if

the coefficient functions fn are symmetric (antisymmetric), then this

decomposition is unique. (Note that we used the term generalized

basis not precisely in the form that is described in A.7.) Generalized

functions

Akl(x1, . . . , xk|y1, . . . , yl) =⟨Aa+(y1) . . . a

+(yl)θ, a+(x1) . . . a

+(xk)θ⟩

are called matrix elements or matrix entries of the operator A in the

generalized basis a+(x1) . . . a+(xn)θ.

In conclusion, we will describe an example of a representation

of CCR that is not a Fock representation. Let us consider, in the

space F (L2(En)), the operator generalized functions b(x) = a(x) +

φ(x), b+(x) = a+(x) + φ(x), where φ ∈ S ′ is a numerical generalized

function. The operators b(x), b+(x) specify a representation of CCR

(we assign the operators

b(f) =

∫f(x)b(x)dx = a(f) +

∫f(x)φ(x)dx,

b+(f) =

∫f(x)b+(x)dx = a+(f) +

∫f(x)φ(x)dx

Page 128: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 81

Operators on Fock Space 81

to every function f ∈ S). It is easy to see that this representation

of CCR can be a Fock representation only in the case when the

function φ is square integrable. To check this, consider the vector

ξ ∈ F (L2(En)) obeying the condition

b(x)ξ = a(x)ξ + φ(x)ξ = 0. (6.9)

Then ⟨ξ, a+(x)θ

⟩= 〈a(x)ξ, θ〉 = −φ(x) 〈ξ, θ〉.

For every vector ξ ∈ F (L2(En)), the function 〈ξ, a+(x)θ〉 is square

integrable because this function enters the Fock column correspond-

ing to the vector ξ.4 This means that in the case when φ /∈ L2(En),

equation (6.9) does not have a solution in the Fock space.

A broader class of representations of CCR is described by the

following formulas:

b(f) = a(Φf) + a+(Ψf) +

∫f(x)φ(x)dx,

b+(f) = a(Ψf) + a+(Φf) +

∫f(x)φ(x)dx

(6.10)

(for concreteness, we assume that f ∈ S(En), φ ∈ S ′(En), and Φ,Ψ

are operators transforming the space S(En) into the space L2(En)).

It is easy to check that the relations (6.10) specify a representation

of CCR in the case when

Φ∗Φ−Ψ∗Ψ = 1,

Φ∗Ψ−Ψ∗Φ = 0,

where the operators Φ,Ψ are defined by the formulas Φf = Φf,Ψf =

Ψf . The relation (6.10) can be written in the form

b(x) =

∫Φ(y, x)a(y)dy +

∫Ψ(y, x)a+(y)dy + f(x),

b+(x) =

∫Ψ(y, x)a(y)dy +

∫Φ(y, x)a+(y)dy + f(x),

(6.11)

4We have used the fact that 〈ξ, θ〉 6= 0. If 〈ξ, θ〉 = 0, then 〈ξ, a+(x1)a+(x2) . . .a+(xn)θ〉 = −φ(x1)

⟨ξ, a+(x2), . . . , a+(xn)θ

⟩. By induction on n, we derive⟨

ξ, a+(x1) . . . a+(xn)θ⟩

= 0, hence ξ = 0.

Page 129: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 82

82 Mathematical Foundations of Quantum Field Theory

where Φ(x, y) and Ψ(x, y) are the kernels of the operators Φ

and Ψ.

One can prove that the representation of CCR specified in (6.10)

is equivalent to the Fock representation if and only if the functions

Ψ(x, y) and f(x) are square integrable (see Berezin (2012), where this

statement and the analogous statement for CAR were formulated and

proved).

The transition from the operators a+(x), a(x) to the operators

b+(x), b(x), described by the formulas (6.11), is called a canonical

transformation in the case when the new operators also satisfy

CR. If the operators b+(f), b(f) generate a Fock representation

of CCR, then one can find a unitary operator U satisfying the

condition

Ua(f)U−1 = b(f),

Ua+(f)U−1 = b+(f).

(This follows from the uniqueness of Fock representation that we

have proven.) Otherwise, such a unitary operator doesn’t exist.

6.2 The simplest operators on Fock space

Let us fix an orthonormal basis φk in the space B, where the index k

runs over the set M . The operators a+(φk), a(φk) in the Fock space

F (B) are denoted a+k , ak. We will use the notation a(k, ε) = aεk, where

ε = ±1, assuming that a1k = a(k, 1) = a+k , a−1k = a(k,−1) = ak.

Let us consider the operators Nk = a+k ak in the Fock space F (B).

It is easy to check that all the operators Nk commute. Let us find

their common eigenvectors.

We consider first the case of CCR. Then the commutation

relations of the operators Nk with the operators a+l , al have the form

[Nk, a+l ] = δk,la

+l , [Nk, al] = −δk,lal.

If φ is a common eigenvector of the operators Nk, in other words

Nkφ = nkφ, then Nka+l φ = a+l Nkφ = nka

+l φ for l 6= k,Nka

+k φ =

a+k (Nk + 1)φ = (nk + 1)a+k φ. Similarly, Nkalφ = nkalφ for l 6=k,Nkakφ = (nk − 1)akφ. We conclude that vectors of the form

Page 130: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 83

Operators on Fock Space 83

a+n1k1

. . . a+nrkr

θ, where ni are non-negative integers and ki 6= kj ,

form a complete orthogonal5 (not orthonormal) system of common

eigenvectors for the operators Nk. The eigenvalue of the operator Nk

in the state a+n1k1

. . . a+nrkr

θ is equal to ni, if k = ki, and equal to zero,

if k does not coincide with any ki.

Let us consider the case of CAR. The operators Nk satisfy the

relation N2k = Nk in this case, which implies that the eigenvalues of

Nk are equal to 0 or 1. Using the equations Nka+l = a+l Nk, Nkal =

alNk for l 6= k, and a+k Nk = 0, Nkak = 0 and Nka+k = a+k (1 −

Nk), akNk = (1−Nk)ak, we can analyze the action of the operators

a+k , ak on the common eigenvector φ of the operators Nk. Namely,

if Nkφ = nkφ and nl = 1, then a+l φ = 0; if nl = 0, then a+l φ is

an eigenvector of the operators Nk; and Nka+l φ = nka

+l φ for l 6=

k,Nka+k φ = a+k φ. Similarly, for nl = 0 we have alφ = 0, and for

nl = 1 we obtain that alφ is an eigenvector of the operators Nk,

and we have Nkalφ = nkalφ for l 6= k,Nkakφ = 0. We see that the

vectors a+k1 . . . a+knθ, where all ki are different, constitute a complete

orthogonal6 (but not orthonormal) system of common eigenvectors

of the operators Nk, namely we have Nkia+k1. . . a+krθ = a+k1 . . . a

+krθ

and Nka+k1. . . a+krθ = 0, if k is not equal to any ki. When we consider

simultaneously the case CCR and CAR, we will denote the vectors

of the complete system that we have constructed as a+n1k1

. . . a+nsks

θ;

however, we should remember that in the case of CAR the numbers

ni cannot be greater than one (i.e. ni = 0, 1).

The operator Nk has the physical meaning of the operator of the

number of particles in the state φk, and the operators a+k and akare called the creation and annihilation operators of a particle in

the state φk. The properties of the operators Nk, a+k , ak agree with

this terminology. The operator N =∑

kNk =∑

k a+k ak is called the

operator of the number of particles and the state x ∈ F (B) satisfying

Nx = nx is called an n-particle state.

5More precisely, two vectors of this kind are either proportional or orthogonal.The completeness of this system of eigenvectors follows from the cyclicity of thevector θ with respect to the family of operators a+(f).6More precisely, two vectors of this kind are either orthogonal or proportional.

Page 131: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 84

84 Mathematical Foundations of Quantum Field Theory

It is easy to see that these definitions agree with the definitions of

Section 3.2. In particular, the set of n-particle states coincides with

the subspace Fn. This implies that the operator of the number of

particles N doesn’t depend on the choice of orthonormal basis φk.

Let us consider the Hamiltonian

H0 =∑

ωka+k ak =

∑ωkNk, (6.12)

where ωk are real numbers.7 Note that the Heisenberg operators

ak(t) = exp(−iH0t)ak exp(−iH0t) and a+k (t) = exp(iH0t)a+k

exp(−iH0t) can be calculated; namely, it is clear that the operators

ak(t) = ak exp(−iωkt) and a+k (t) = a+k exp(iωkt) satisfy the following

Heisenberg equations:

dak(t)

dt= i[H0, ak(t)] = −iωkak(t),

da+k (t)

dt= i[H0, a

+k (t)] = iωka

+k (t).

The operator H0 commutes with the operators Nk, therefore

the system of vectors a+n1k1

. . . a+nrkr

θ that we have constructed is

a complete system of eigenvectors of the operator H0 and the

eigenvalues are equal to∑r

i=1 ωkini.

Let us find the ground state of the Hamiltonian H0. In the case

of CCR, if at least one ωk < 0, then the spectrum of energies is

unbounded from below and therefore the ground state does not exist.

If all ωk ≥ 0, then all energy levels are non-negative and the vacuum

vector θ is the ground state. In the case of CAR, the lowest energy

level is equal to E0 =∑

k∈L ωk, where the sum is taken over the set L

consisting of k, for which ωk < 0. To obtain the ground state Φ from

the vacuum vector θ, we should apply all operators a+k with k ∈ Lto the vacuum vector (in other words, Φ = (

∏k∈L a

+k )θ). If the set

L is infinite, then the ground state doesn’t exist. Let’s introduce the

7The operator H0 is an essentially self-adjoint operator for any choice of realnumbers ωk (this follows from the fact that the operator H0 has a completesystem of eigenvectors). In particular, the operator of the number of particles Nis essentially self-adjoint. As usual, we identify the essentially self-adjoint operatorwith its self-adjoint extension.

Page 132: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 85

Operators on Fock Space 85

operators bk in the following way: bk = a+k for k ∈ L and bk = ak for

k 6∈ L. The operators b+k , bk satisfy the relations (6.5). To check that

these operators generate a representation of CAR, we should define

the operators b+(f), b(f) for all f ∈ B. It is easy to do this, setting

b(f) = a+(Pf) + a((1 − P )f), where Pf =∑

k∈L 〈f, φk〉φk. It is

convenient to express the Hamiltonian H0 in terms of the operators

b+k , bk. If the set L is finite, then

H0 =∑|ωk|b+k bk +

∑k∈L

ωk,

and then the ground state of the Hamiltonian H0 is the vacuum

vector for the operators bk, i.e. it satisfies the conditions bkΦ = 0. If

the set L is infinite, then the vector satisfying the condition bkΦ =

0 in the space F (B) doesn’t exist. In other words, the operators

b+(f), b(f) specify a representation of CAR that is not equivalent to

the Fock representation.

The operators acting in Fock space F (B), where B = L2(X), can

be expressed in terms of operator generalized functions a+(x), a(x),

satisfying the relations (6.7). The statements proven in this section

can also be proven for the operators in the space F (B) that are

expressed in terms of a+(x), a(x). In particular, one can introduce

the commuting operators (or more precisely operator generalized

functions) N(x) = a+(x)a(x). The operator of the number of

particles N can be expressed in terms of the operators N(x) by the

formula

N =

∫N(x)dx

(more details about the operator of the number of particles can be

found in Section 3.2). A complete system of generalized eigenvectors

(generalized eigenbasis) of the operator

H0 =

∫ω(x)a+(x)a(x)dx =

∫ω(x)N(x)dx (6.13)

consists of vectors (vector generalized functions) a+(x1) . . . a+(xn)θ.

Operators of the form (6.12), (6.13) are among the simplest

operators in Fock space. Slightly more general operators (3.9) are

Page 133: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 86

86 Mathematical Foundations of Quantum Field Theory

considered in Section 3.2. The analysis of Hamiltonians of the form

H =

∫A(x, y)a+(x)a(y)dxdy +

∫B(x, y)a+(x)a+(y)dxdy

+

∫B(x, y)a(y)a(x)dxdy

+

∫C(x)a+(x)dx+

∫C(x)a(x)dx (6.14)

(quadratic Hamiltonians) is more complicated. Under certain condi-

tions, they can be reduced to the form (6.13) by means of a linear

canonical transformation of the form (6.11). A complete analysis of

the Hamiltonian (6.14) can be found in the book by Berezin (2012).

Much harder is the situation for the Hamiltonians of degree higher

than two, with respect to the operators of creation and annihilation.

Very rarely one can find precisely the eigenvalues or the eigenvectors

of these operators and hence it is necessary to apply other methods.

In Chapter 11, we will explain a method of calculating the evolution

operator and the eigenvalues of the Hamiltonian H = H0 + gV ,

where H0 =∑ωka

+k ak, in the form of power series with respect to

the parameter g (perturbation theory).

In this section, we will consider the conditions under which the

formal expression

H =∑m

∑ki,εi

Γmk1,ε1,...,km,εmaε1k1. . . aεmkm (6.15)

defines an operator on the Fock space F (B) (here, as always, a1k =

a+k = a+(φk), a−1k = ak = a(φk), where φk is an orthonormal basis

for B, ki are taken from the set M , and εi = ±1). In the case of CAR,

one can prove the following statement. If the numerical series∑m

∑ki,εi

Γmk1,ε1,...,km,εm (6.16)

is absolutely convergent, then the operator series (6.15) is also

absolutely convergent (in the sense of norm convergence) and specifies

a bounded operator in F (B). This follows from the remark that

in the case of CAR we have that ‖aεk‖ ≤ ‖φk‖ = 1 and hence,

‖aε1k1 . . . aεmkm‖ ≤ 1.

Page 134: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 87

Operators on Fock Space 87

If, in addition, the expression (6.15) is formally Hermitian, i.e.

obeys

Γmk1,ε1,...,km,εm = Γmk1,−ε1,...,km,−εm , (6.17)

then the operator H defined by the formula (6.15) is self-adjoint

(since bounded Hermitian operators are self-adjoint).

In the case of CCR, we can use the following estimates for an

n-particle state x:

‖aεkx‖ ≤√n‖x‖,

‖aε1k1 . . . aεmkmx‖ ≤

√n√n+ 1 . . .

√n+m− 1‖x‖ (6.18)

(the first of which can be obtained from the relation

‖akx‖2 =⟨a+k akx, x

⟩≤

⟨∑k

a+k akx, x

⟩= 〈Nx, x〉 = n‖x‖2,

and the second follows from the first).

Using the estimate (6.18), one can check that the expression (6.15)

specifies an operator with everywhere dense domain in F (B), if this

expression is polynomial (Γmk1,ε1,...,km,εm = 0 for m > s) and the

numerical series (6.16) is absolutely convergent. This follows from

the remark that the series

s∑m=1

∑ki,εi

Γmk1,ε1,...,km,εmaε1k1. . . aεmkmx

converges for x ∈ Fn, and hence the domain of the operator H

contains D — the union of all linear spaces Fn. If the operator H is

formally Hermitian, i.e. the condition (6.17) is satisfied, then H will

be Hermitian on D. However, it is not clear whether this operator

is essentially self-adjoint on D; results in this direction can be found

in the paper by Hepp (1966). The operators of interest for physics

are usually bounded from below and hence they define a self-adjoint

operator by means of Friedrich’s extension (see Appendix A.5). The

analysis of the operator H defined by the expression (6.15) is much

easier when Γmk1,ε1,...,km,εm 6= 0 only for ε1 + · · · + εm = 0 (i.e.

Page 135: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 88

88 Mathematical Foundations of Quantum Field Theory

if the operator H commutes with the operator of the number of

particles N). Under this condition, one can consider the operator H

separately on each subspace Fn. One can extend the class of operators

of the form (6.15) that defines self-adjoint operators, by noting that

adding to a self-adjoint operator (for example, to H0) a bounded

self-adjoint operator, we again obtain a self-adjoint operator.

We will not analyze the conditions under which expressions of the

form

H =∑m

∑ε1,...,εm

∫Γm(x1, ε1, . . . , xn, εn)a(x1, ε1)

. . . a(xn, εn)dx1 . . . dxn (6.19)

define an operator on the space F (B) = F (L2(X)). Note that only

for the case X = Er some results in this direction are proven in

Section 3.2.

In the case of CAR an operator corresponding to a physical quantity

(in particular, a Hamiltonian) should always contain an even number

of creation and annihilation operators (i.e. Γm = 0 for odd m).

6.3 The normal form of an operator: Wick’s theorem

Let us say that an operator C in the Fock space F (B) is presented

in normal form if the operators of creation are to the left of the

annihilation operators:

C =∑m,n

Γm,n(k1, . . . , km|l1, . . . , ln)a+k1 . . . a+kmal1 . . . aln (6.20)

(here ak = a(φk), where φk is an orthonormal basis in B. As usual, we

sum over repeated indices ki, lj). The sum in the expression above can

be infinite, in which case the convergence of the series is understood

in the sense of strong operator convergence. The set of indices for

the operators ak will be denoted by M . Since M is countable, we

can identify M with the set of natural numbers; however, this is not

always convenient.

The product of an arbitrary number of operators a+k , al, taken in

any order, as well as any finite sum of such operators, can be written

in normal form by means of CR. It can be shown that any bounded

Page 136: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 89

Operators on Fock Space 89

operator can be written in normal form if the convergence in (6.20)

is understood in the sense of weak convergence.

The representation of an operator in normal form is especially

convenient by the calculation of the vacuum average (vacuum

expectation value) 〈Cθ, θ〉 of the operator C. It is clear that 〈Cθ, θ〉 =

Γ0,0 (the constant term in the representation of the operator C in

normal form).

Let us assume that C1, . . . , Ck are operators written in normal

form. Let us consider the following question: How can we write the

product C1 . . . Ck of these operators in normal form?

Before we answer this question, let us consider a simpler (though,

in some sense, more general) situation. Let A1, . . . , Am, B1, . . . , Bnbe operators with the property that the commutators [Aα, Bβ]

are numbers (i.e. they can be written as product of the iden-

tity operator and a number). The number BβAα = [Aα, Bβ] is

called the contraction of the operators Aα and Bβ. The prod-

uct B1 . . . BnA1 . . . Am with one contraction BβAα will be under-

stood as this product with the operators Bβ and Aα deleted

and with the factor BβAα inserted instead of these operators;

to denote the product with the contraction BβAα we will write

B1 . . . Bβ . . . BnA1 . . . Aα . . . Am. The product B1 . . . BnA1 . . . Am

with k contractions Bβ1Aα1 , . . . , BβkAαkwill be understood as a

product where the operators Bβ1 , . . . , Bβk , Aα1 , . . . , Aαkare deleted

and the contractions Bβ1Aα1 , . . . , BβkAαkare included in their

place. Hence, in order to define a product with k contractions, we

should select k operators Bβ1 , . . . , Bβk (the order of the selection is

unimportant), and to every one of these operators, we should assign

one of the operators A1, . . . , An, assuming that different indices

β1, . . . , βk correspond to different indices α1, . . . , αk.

Let us now formulate the following statement, which we will call

Wick’s theorem, noting that it presents a simplified form of this

theorem:

The product A1 . . . AmB1 . . . Bn is equal to the product B1 . . . BnA1

. . . Am with additional summand terms obtained by all possible

contractions.

Page 137: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 90

90 Mathematical Foundations of Quantum Field Theory

For example,A1A2B1B2 = B1B2A1A2 +B1B2A1A2 +B1B2A1A2 +B1B2A1A2

+B1B2A1A2 +B1B2A1A2 +B1B2A1A2

+B1B2A1A2 .

This statement is almost obvious: when we transfer an operator

Bβ to the left, we use the relation AαBβ = BβAα + BβAα, and

therefore every time we change the order of Bβ and Aα, we will add

two summands: one with the same number of contractions as in the

original summand and another where the number of contractions is

greater by one.

In order to give a formal proof, we can start by demonstrating

the following statement (either directly or by induction over m):

A1 . . . AmB = BA1 . . . Am +BA1 . . . Am +BA1A2 . . . Am

+ · · ·+BA1 . . . Am . (6.21)

In other words, we check Wick’s theorem for n = 1. Then, we can

use mathematical induction with respect to the number of operators

Bβ; at every induction step, we use the formula (6.21).

Let us now consider the reduction to normal form of the product

of two operators K and L, both represented in normal form. Without

loss of generality, we can assume that K and L have definite degree

with respect to creation and annihilation operators:

K = K(k1, . . . , km|l1, . . . , ln)a+k1 . . . a+kmal1 . . . aln ,

L = L(p1, . . . , pr|q1, . . . , qs)a+p1 . . . a+praq1 . . . aqs

(as always, we have in mind a summation over the repeated indices

from the set M).

Let us begin with the case of CCR. In this case, applying Wick’s

theorem to the product al1 . . . alna+p1 . . . a

+pr , one can show that KL

is equal to

K(k1, . . . , km|l1, . . . , ln)L(p1, . . . , pr|q1, . . . , qs)

×a+k1 . . . a+kma+p1 . . . a

+pral1 . . . alnaq1 . . . aqs ,

Page 138: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 91

Operators on Fock Space 91

with the addition of summands obtained from this expression by

means of arbitrary contractions between the operators a+pi and alj(recall that a+p al = [al, a

+p ] = δpl).

For example, if K = K(k|l)a+k al and L = L(p1, p2|q1, q2)a+p1a+p2aq1aq2 , then

KL = K(k|l)L(p1, p2|q1, q2)a+k a+p1a

+p2alaq1aq2

+K(k|l)L(p1, p2|q1, q2)a+k a+p1a

+p2al aq1aq2

+K(k|l)L(p1, p2|q1, q2)a+k a+p1 a

+p2al aq1aq2

= K(k|l)L(p1, p2|q1, q2)a+k a+p1a

+p2alaq1aq2

+K(k|l)L(l, p2|q1, q2)a+k a+p2aq1aq2

+K(k|l)L(p1, l|q1, q2)a+k a+p1aq1aq2 .

The functions K(k1, . . . , km|l1, . . . , ln) and L(p1, . . . , pr|q1, . . . , qs)can be assumed to be symmetric with respect to every group of

indices. In what follows, we will assume that this symmetry condition

is always satisfied. Using this assumption, we can considerably

reduce the number of summands in the expression for KL, since

the summands corresponding to the same number of contractions

are identical. In our example, using the symmetry of coefficients, one

notices that the second and third summands are equal.

The reduction to normal form can be represented graphically by

means of diagrams suggested by Feynman. The operator

K = K(k1, . . . , km|l1, . . . , ln)a+k1 . . . a+kmal1 . . . aln

will be denoted by a diagram consisting of one internal vertex with

m dotted incoming lines and n dotted outgoing lines (we consider

the lines to be directed topological intervals). We will say that such a

diagram is a star (see Fig. 6.1). The beginning of an incoming line will

be called an in-vertex and the end of an outgoing line will be called an

out-vertex. We will assign the indices k1, . . . , km ∈M to in-vertices;

to out-vertices, we will assign the indices l1, . . . , ln ∈ M ; and to the

internal vertex, we will assign the function K(k1, . . . , km|l1, . . . , ln).

The operator L is depicted by a similar diagram.

Page 139: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 92

92 Mathematical Foundations of Quantum Field Theory

Fig. 6.1 Star: Feyman diagram of one internal vertex.

The normal form of the operator KL can be expressed now as

a collection of all different diagrams, consisting of diagrams of the

operators K and L, where some of the out-vertices of the diagram for

the operator K are connected with solid lines with the in-vertices of

the diagram for the operator L. These solid lines will be called edges

of the diagram and we assign the contractions a+p al = [al, a+p ] = δlp

to the edges. (Here, l is the index of the end of the edge and p is the

index of the beginning of the edge.) The direction of the edge will

always be fixed as the direction from the out-vertex of the diagram of

the operator K to the in-vertex of the diagram of the operator L. For

every diagram, we construct a product in which the internal vertices

of the diagram contribute the factors K(k1, . . . , km|l1, . . . , ln) and

L(p1, . . . , pr|q1, . . . , qs), where k1, . . . , km, p1, . . . , pr are the indices

of the ends of the lines entering the vertex and l1, . . . , ln, q1, . . . , qsare the indices of the ends of the lines exiting the in-vertex. The

factor a+p al = δlp is assigned to every connecting edge. Finally, the

lines with a free end contribute the operator a+k or ak, if the line is

incoming or outgoing, respectively. We assume a summation over the

set M for each index. The operators a+k , al are assumed to be written

in normal order.

It is easy to check that the operator constructed by the diagram

described above coincides with one of the summands in the product

of the operators K and L written in normal form. Therefore, the

Page 140: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 93

Operators on Fock Space 93

operator KL is equal to the sum of operators corresponding to all the

possible diagrams (more shortly, to the sum of diagrams).

By induction, it is easy to obtain diagrams for the reduction to

normal form of a product of r operators K1, . . . ,Kr, where

Ki = Ki(k1, . . . , kmi |l1, . . . , lni)a+k1. . . a+kmi

al1 . . . alni.

The normal form of the operator K1 . . .Kr is depicted by a sum of

diagrams consisting of the diagrams of the operators K1, . . . ,Kr and

some number of lines connecting the outgoing vertices of the diagram

for the operator Ki with the incoming vertices of the diagram of the

operator Kj (here i, j = 1, . . . , r; i < j). The prescription that assigns

to a diagram an operator remains the same.

If the operator Ki =∑

m,nKm,ni , where

Km,ni = Km,n

i (k1, . . . , km|l1, . . . , ln)a+k1 . . . a+kmal1 . . . aln ,

then the diagram of the operator Ki is a set of several stars, the

diagrams of the operators Km,ni .

The normal form of the operator K1 . . .Kr is represented then

by the collection of diagrams consisting of r stars — a diagram of

every operator K1, . . . ,Kr — and several lines connecting outgoing

vertices of the star diagram of one of the operators Ki with incoming

vertices of the star diagram of one of the operators Kj , j > i.

The diagram technique in Fig. 6.1 can be applied to reduce

to the normal form a product of operators expressed in terms

of creation and annihilation operators satisfying CCR. A similar

diagram technique can also be constructed in the case of CAR; the

only difference is in some signs in front of the diagrams. We do not

describe explicitly the rules for these sign factors; however, we will

formulate accurately the analog of Wick’s theorem for the CAR case.

Let A1, . . . , Am, B1, . . . , Bn be operators having the property that

the anticommutators [Aα, Bβ]+ are numbers. The number BβAα =

[Aα, Bβ]+ is called a contraction of the operators Aα and Bβ. The

product B1 . . . BnA1 . . . Am with k contractions Bβ1Aα1 , . . . , BβkAαk

is defined as the operator obtained by deleting from the product

Page 141: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 94

94 Mathematical Foundations of Quantum Field Theory

the operators Bβ1 , . . . , Bβk , Aα1 , . . . , Aαkand including instead of

them the contractions Bβ1Aα1 , . . . , BβkAαkand the sign factor

(−1)mn+λ+µ, where λ is the sum of the lengths of all contractions

and µ is the number of pairs of overlapping contractions (the length

of a contraction BβAα is defined as n−β+α; two contractions BβAα

and Bβ′Aα′ are called overlapping if (β′ − β)(α′ − α) > 0).

Given these definitions, one can formulate Wick’s theorem in the

form that is applicable in both situations.

The product A1 . . . AmB1 . . . Bn is equal to the product

B1 . . . BnA1 . . . Am with all possible contractions (we do not exclude

the case where the number of contractions is equal to zero).

For example, in the case when the anticommutators [Aα, Bβ]+ are

numbers, we obtain

An operator, acting on the Fock space F (B), where B = L2(X),

is said to be represented in normal form if it is written in the form

A =∑m,n

∫km,n(x1, . . . , xm|y1, . . . , yn)a+(x1) . . . a

+(xm)

× a(y1) . . . a(yn)dmxdny (6.22)

(here X is a measure space and a+(x), a(x) are operator generalized

functions satisfying the relation (6.7)). The diagram techniques

discussed above can also be used to represent in normal form

the product of operators written in the form (6.22). The only

modification that is necessary in the case in question is that the

role of the indices k ∈ M is played by the elements of the space

Page 142: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 95

Operators on Fock Space 95

X, correspondingly the summation with respect to M is replaced by

integration over the space X.8

Let us use the above results to prove some useful relations. For

definiteness, we will consider the case of CCR.

First of all, we will calculate the inner product⟨a+(f1) . . . a

+(fm)θ, a+(g1) . . . a+(gn)θ

⟩=⟨a(gn) . . . a(g1)a

+(f1) . . . a+(fm)θ, θ

⟩.

The operator a(gn) . . . a(g1)a+(f1) . . . a

+(fm) can be reduced to

normal form by means of Wick’s theorem; it is equal to the sum where

every summand is the product a+(f1) . . . a+(fm) × a(gn) . . . a(g1)

with some contractions. Remembering that the vacuum average of

an operator is equal to the constant term in the representation of the

operator in normal form, we see that the quantity we would like to

calculate is equal to the sum of the summands, where all operators

are contracted (i.e. m = n is equal to the number of contractions).

Noting that

a+(fi)a(gj) = [a(gj), a+(fi)] = 〈gj , fi〉 = 〈fi, gj〉,

thus, we obtain the necessary formula⟨a+(f1) . . . a

+(fm)θ, a+(g1) . . . a+(gn)θ

⟩= δnm

∑P

〈f1, gi1〉〈f2, gi2〉 . . . 〈fm, gim〉 (6.23)

8It is useful to note that an orthonormal basis φk, with k ∈ M , for the spaceB, specifies an isomorphism between the spaces B and L2(M). (The set M isconsidered as a measure space, equipped with the counting measure, where themeasure of a finite subset is equal to the number of its elements; then the integralof f(k) over the set M is equal to

∑k∈M f(k).) This remark allows us to say

that the representation of an operator in the form (6.20) is a particular case ofrepresentation in the form (6.22); correspondingly, the diagram technique for theoperators of the form (6.22) is a generalization of the techniques for operators ofthe form (6.20).

Page 143: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 96

96 Mathematical Foundations of Quantum Field Theory

(where the sum is taken over all permutations P = (i1, . . . , im)). The

formula (6.23) can also be expressed in the form⟨a+(x1) . . . a

+(xm)θ, a+(y1) . . . a+(yn)θ

⟩= δnm

∑P

δ(x1, yi1) . . . δ(xm, yim), (6.24)

where a+(x), a(y) are operator generalized functions satisfying the

relations (6.7).

Using the statements proved in this section, we can find the

relation between the coefficient functionsKm,n(x1, . . . , xm|y1, . . . , yn)

in the representation of an operator A in the form (6.22) (normal

form) and the matrix elements of the operator A in the generalized

basis a+(x1) . . . a+(xm)θ. To find these relations, we can, for example,

write the operator Aa+(y1) . . . a+(yn)θ in normal form by means of

Wick’s theorem, and then we can calculate the matrix element

Am,n(x1, . . . , xm|y1, . . . , yn)

=⟨Aa+(y1) . . . a

+(yn)θ, a+(x1) . . . a+(xm)θ

⟩using the formula (6.24). As a result, we obtain the following formula:

Am,n(x1, . . . , xm|y1, . . . , yn) = Km,n(x1, . . . , xm|y1, . . . , yn)m!n!

+∑i,j

Km−1,n−1(x1, . . . , xi−1, xi+1, . . . , xm|y1, . . . , yj−1, yj+1,

. . . , yn)δ(xi − yj)(m− 1)!(n− 1)! + · · ·

+∑B,B′,α

Km−r,n−r(xi, i∈B|yj , j∈B′)

×∏i∈B

δ(xi − yα(i))(m− r)!(n− r)! + · · ·

(here B and B′ consist of r elements, B ⊂ 1, . . . ,m, B′ ⊂1, . . . , n, α is a one-to-one correspondence between B and B′,

and we are summing over all possible B,B′, α; Km−r,n−r(xi, i∈B|yj ,j∈B′) denotes the function Km−r,n−r with arguments xi and yjsatisfying i∈B, j∈B′).

Page 144: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 97

Operators on Fock Space 97

To conclude this section, we will formulate some useful definitions.

Let K and L be operators represented in normal form as follows:

K =∑m,n

∫Km,n(k1, . . . , km|l1, . . . , ln)a+(k1)

. . . a+(km)a(l1) . . . a(ln)dmkdnl,

L =∑m,n

∫Lm,n(k1, . . . , km|l1, . . . , ln)a+(k1)

. . . a+(km)a(l1) . . . a(ln)dmkdnl.

Then the normal product of operators K and L in the case of CCR

is the operator

N(KL) =∑

m,n,r,s

∫Km,n(k1, . . . , km|l1, . . . , ln)Lr,s(p1,

. . . , pr|q1, . . . , qs)a+(k1) . . . a+(km)a+(p1) . . . a

+(pr)a(l1)

. . . a(ln)a(q1) . . . a(qs)dmkdnldrpdsq,

and in the case of CAR is the operator

N(KL) =∑

m,n,r,s

(−1)nr∫Km,n(k1, . . . , km|l1, . . . , ln)Lr,s(p1,

. . . , pr|q1, . . . , qs)a+(k1) . . . a+(km)a+(p1) . . . a

+(pr)a(l1)

. . . a(ln)a(q1) . . . a(qs)dmkdnldrpdsq

(instead of using the symbol N(KL) to denote normal product, one

can also use the symbol :KL:). In other words, the normal product

is obtained if we reduce the product KL to the normal form using

the prescriptions above and delete all summands containing at least

one contraction (i.e. we delete all diagrams having at least one non-

dotted line). The normal product of n operators K1, . . . ,Kn can be

defined by induction

N(K1, . . . ,Kn) = N(N(K1, . . . ,Kn−1)Kn).

Page 145: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 98

98 Mathematical Foundations of Quantum Field Theory

The normal exponent of the operator K is the operator

N(expK) =∞∑n=0

1

n!N(Kn).

6.4 Diagram techniques

Let us consider the Hamiltonian H = H0 + gV (t), where

H0 =∑

ε(k)a+k ak, (6.25)

where

V (t) =∑m,n

∑k1,...,km

vm,n(k1, . . . , km|p1, . . . , pn|t)a+k1 . . . a+km

×ap1 . . . apn

(here, a+k = a+(φk), ak = a(φk) are operators in the Fock space F (B)

corresponding to the orthonormal basis φk ∈ B and the index k runs

over the set M). For concreteness, we consider the case of CCR;

however, the consideration of this section can also be performed

in the case of CAR (in the case of CAR, the Hamiltonian H is

assumed to be Fermi-even, i.e. it is assumed that every summand

in the operator H contains an even number of operators a+ and a

(m+ n is even)).

Using the representation of the operator S(t, t0) in the form of

T -exponent (4.4) and Wick’s theorem (Section 6.3), we can construct

diagram techniques for the calculation of the operator S(t, t0). More

precisely, we obtain the decomposition of the normal form of S(t, t0)

as a power series in g.

Let us recall that a star is a diagram consisting of a point with m

incoming lines and n outgoing lines. The lines belonging to a star are

depicted as dotted lines (Fig. 6.1 depicts a star with three incoming

and two outgoing lines).

A diagram of the operator S(t, t0) is a collection of several stars

and several edges that are depicted as directed lines. Every edge

starts with an out-vertex of some star and ends in an in-vertex of

another star. Let us assume that two lines belonging to a diagram

cannot have common internal points and can have only one common

Page 146: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 99

Operators on Fock Space 99

vertex. Two vertices belonging to the same edge cannot belong to

the same star.

The vertices belonging to edges are called internal vertices. Other

vertices are called external vertices.

All vertices of a diagram should be numbered in such a way that

the vertices belonging to the same star are numbered by neighboring

numbers. Two diagrams are equivalent if there exists a topological

equivalence between these diagrams transforming a vertex with some

number to a vertex with the same number.

Some examples of diagrams are drawn in Fig. 6.2. The diagrams

(a) and (b) are topologically equivalent, but are distinct, while the

diagrams (a) and (c) are equivalent.

For every diagram, we can construct an operator in the following

way. Let us assume that to every vertex of the diagram we have

assigned an index ki ∈ M and a real number ti (time). We assign

the same time to every vertex belonging to one star. Every star

contributes a factor vm,n(k1, . . . , km|p1, . . . , pn|t), where t is the time

of this star and k1, . . . , km are indices of in-vertices of this star

and p1, . . . , pn are indices of the out-vertices of the star. An edge

contributes

exp(−iε(k1)(t1 − t2))δk2k1θ(t1 − t2),

where the index k1 ∈ M and the time t1 correspond to the

beginning of the edge and k2 ∈ M and t2 correspond to the end

of the edge. To every free in-vertex, we assign an operator a+k (t) =

exp(iε(k)t)a+k , and to every free out-vertex, we assign the operator

ak(t) = ak exp(−iε(k)t) (here k and t stand for the index and the

time corresponding to the vertex, respectively).

To every diagram, we assign an operator obtained in the following

way. First of all, we take the product of functions corresponding to

the stars and edges of the diagram and the operators corresponding

to the external in- and out-vertices. The operator is obtained from

this product by means of summing over the indices of all vertices and

integration over the times of all stars (calculating S(t, t0) we should

integrate over the interval [t0, t]). The operators corresponding to the

external vertices are taken in normal order. We include the factor1n!g

n in the operator corresponding to the diagram with n stars.

Page 147: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 100

100 Mathematical Foundations of Quantum Field Theory

Fig. 6.2 Examples of diagrams.

Let us write down, for example, the operator corresponding to

Fig. 6.2(a):

1

3!g3

∑k1,...,k9

∫dt1dt2dt3v1,2(k1|k2, k3|t1)v2,1(k4, k5|k6|t2)

× v2,1(k8, k7|k9|t3) exp(−iε(k3)(t1 − t3))δk8k3θ(t1 − t3)

× exp(−iε(k2)(t1− t2))δk4k2θ(t1 − t2) exp(−iε(k6)(t2 − t3))

× δk7k6 (t2 − t3) exp(iε(k1)t1 + iε(k5)t2

− iε(k9)t3)a+k1a+k5ak9 .

Page 148: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 101

Operators on Fock Space 101

Let us prove that the operator S(t, t0) can be expressed as a

sum of all different diagrams, or more precisely, as a sum of all the

operators corresponding to these diagrams. Of course, this statement

(and other statements of this kind) means only that the sum of the

diagrams gives a decomposition S(t, t0) as a series in g; we cannot

say anything about the convergence of this series.

Before giving the proof of this statement, let us make several

remarks about the structure of the diagram representation of the

operator S(t, t0) and consider an example. Let us note first that we

can omit all diagrams containing n stars joined in a cyclic order by

edges. (We say that n stars are connected in cyclic order by edges if

they can be ordered in such a way that for every i, 1 ≤ i ≤ n, there

exists an edge starting in a vertex of the ith star and ending in the

vertex of the (i+1)-th star; we identify the (n+1)th star with the first

star.) Let us denote the times of these stars by τ1, . . . , τn; then the

edges introduce the number θ(τ1− τ2) . . . θ(τn−1− τn)θ(τn− τ1), and

hence in the integral defining our diagram, the integrand is equal

to zero unless τ1 = τ2 = · · · = τn, which means that the integral

vanishes.

It is easy to check that two topologically equivalent diagrams

specify the same operator, therefore one usually draws only one of

the class of topologically equivalent diagram and this diagram should

be multiplied by the number of topologically equivalent diagrams.

Let us consider as an example the Hamiltonian of the system

of non-relativist identical bosons in a box of the volume L3. Here,

H = H0 + V , where

H0 =∑ k2

2ma+k ak,

V =1

2

(2π

L

)3 ∑k1,k2,p1,p2

W(k1 − p2)δk1+k2p1+p2 × a+k1a

+k2ap1ap2 ,

and k, k1, k2, p1, p2 run over a lattice in three-dimensional space with

a spacing of 2π/L.

Let us write down the summands corresponding to the diagrams

of Fig. 6.3 in the decomposition of the adiabatic S-matrix Sα. The

Page 149: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 102

102 Mathematical Foundations of Quantum Field Theory

Fig. 6.3 More examples of diagrams.

first of these two diagrams represents the operator

1

2

(2π

L

)3 ∫ ∞−∞

∑k1,k2,p1,p2

W(k1 − p2)δk1+k2p1+p2 exp(−α|τ |)

× a+k1(τ)a+k2(τ)a+p1(τ)a+p2(τ)dτ

=1

2

(2π

L

)3 ∑k1,k2,p1,p2

W(k1 − p2)δk1+k2p1+p2

× 2α(k212m +

k222m +

p212m +

p222m

)2+ α2

a+k1a+k2ap1ap2 .

The operator corresponding to the second diagram has the form

1

2l

∫dτ1dτ2

∑k1,k2,p1,p2

∑k′1,k

′2,p′1,p′2

1

4

(2π

L

)6

W(k1 − p2)

× δk1+k2p1+p2 exp(−α|τ1|) exp

(−i p

21

2m(τ1 − τ2)

)δp1k′1θ(τ1 − τ2)

× exp

(−i p

22

2m(τ1 − τ2)

)θ(τ1 − τ2)δp2k2W(k′1 − p′2)δ

k′1+k′2

p′1+p′2

× exp(−α|τ2|)a+k1(τ1)a+k2

(τ1)ap′1(τ2)ap′2(τ2)

(where ki, pi and k′i, p′i are summed over a lattice).

Page 150: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 103

Operators on Fock Space 103

The third diagram is equivalent to the second; we take this into

account by multiplying the second diagram by 2.

We can reduce the number of diagrams that should be calculated

if we note that the operator corresponding to a disconnected diagram

is equal to the normal product of the operators corresponding to each

connected component.

Let us start now the derivation of the diagram representation of

the operator S(t, t0). First of all, we note that for the representation

of the operator V (τ1) . . . V (τn) in normal form, we can use the

technique developed in Section 6.3 with some modifications. (This

follows from the formula

V (τ) = exp(iH0τ)V (τ) exp(−iH0τ)

=∑m,n

∑k1,...,km

vm,n(k1, . . . , km|p1, . . . , pn|τ)a+k1(τ)

. . . a+km(τ)ap1(τ) . . . apn(τ)

=∑m,n

∑k1,...,km

exp

i m∑j=1

ε(kj)−n∑j=1

ε(pj)

τ

×vm,n(k1, . . . , km|p1, . . . , pn|τ)a+k1 . . . a

+kmap1 . . . apn).

Let us describe the diagram representation of the operator

V (τ1) . . . V (τn) in more detail. The representation consists of dia-

grams that constitute the collection of n stars and some edges that

point in the direction of increasing order (recall that the vertices of

the diagram are ordered).

To every vertex of the diagram, we assign the index k ∈ M .

To every free in-vertex (out-vertex), we assign the operator a+k (τi)

(correspondingly the operator ak(τi)). To every edge, we assign the

function exp(−iε(k)(τi − τj))δkl , where k, l are indices of the vertices

at the beginning and end of the edge and i, j are the numbers of

stars that contain the beginning and end of the edge. The operator

corresponding to the diagram is obtained from the product of the

Page 151: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 104

104 Mathematical Foundations of Quantum Field Theory

operators and functions that correspond to free vertices, stars, and

edges. Namely, one should sum over the indices of all vertices (the

operators are written in normal order).

It follows from the calculation of Section 6.3 that the operator

V (τ1) . . . V (τn) is the sum of operators corresponding to all the

different diagrams.

Using the diagram techniques for the normal form of the operator

V (τ1) . . . V (τn), we can get the diagram technique for representing

in normal form the operator T (V (τ1) . . . V (τn)), where all times are

different. The only difference in the techniques is that we should

consider arbitrary diagrams with n stars without restriction on the

direction of the edges and to every edge we should assign the function

exp(−iε(k)(τi − τj))δkl θ(τi − τj). This follows from the remark that

the factors θ(τi − τj) in the functions assigned to the edges are

equal to 1 in the diagrams where the edges go in the direction

of decreasing time and are equal to 0 in all other diagrams. This

means that the diagrams that are not equal to zero coincide with

the diagrams from the diagram representation for V (τi1) . . . V (τin),

where i1, . . . , in is a permutation where τi1 > · · · > τin . Noting that

V (τi1) . . . V (τin) = T (V (τ1) . . . V (τn)), we obtain the representation

that we have claimed.

To get a diagram representation of the operator S(t, t0), it is

sufficient to refer to the expression of this operator as a T -exponent

(4.4) and to note that in the integral over τ1, . . . , τn the value of the

integrand on the set where at least two variables τ1, . . . , τn coincide

is irrelevant.

The diagram representation of the operator S(t, t0) can be easily

obtained in the case when the index k runs over an arbitrary measure

space X. More precisely, let us consider the Fock space F (L2(X)),

where X is a measure space, and in this space we define the operator

generalized functions a+(k), a(k), obeying

[a(k), a(k′)] = [a+(k), a+(k′)] = 0,

[a(k), a+(k′)] = δ(k, k′); a(k)θ = 0.

Page 152: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:10 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch06 page 105

Operators on Fock Space 105

Let us fix the Hamiltonian H = H0 + gV (t), where

H0 =

∫ε(k)a+(k)a(k)dk,

V (t) =∑m,n

∫vm,n(k1, . . . , km|p1, . . . , pn|t)a+(k1)

. . . a(pn)dmkdnp.

We would like to find the normal form of the operator S(t, t0)

for this Hamiltonian. The diagram technique to solve this problem

is analogous to the technique described above. The only difference is

that to every vertex of the diagram we should assign a point k ∈ Xinstead of k ∈M and we should replace the summation with respect

to M with integration with respect to X.

Page 153: Mathematical Foundations of Quantum Field Theory

b2530 International Strategic Relations and China’s National Security: World at the Crossroads

b2530_FM.indd 6 01-Sep-16 11:03:06 AM

This page intentionally left blankThis page intentionally left blankThis page intentionally left blankThis page intentionally left blank

Page 154: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch07 page 107

Chapter 7

Wightman and Green Functions

7.1 Wightman functions

Let us consider a Hilbert space H, a Hamiltonian H with the ground

state Φ, and a family of operators a(λ)λ∈Λ closed with respect

to Hermitian conjugation. More precisely, let us assume that the

operators a(λ) are defined on a dense subspace D ⊂ H and for every

operator a(λ), there exists an operator a(λ) that satisfies the relation

〈a(λ)x, y〉 = 〈x, a(λ)y〉 for every x, y ∈ D. The symbol a(λ, t) will

denote the Heisenberg operators a(λ, t) = exp(iHt)a(λ) exp(−iHt).We will assume that the operators a(λ) and exp(iHt) transform

the set D into itself. Note that under these assumptions, the

operators a(λ1, t1) . . . a(λn, tn) also transform D into itself and that

(a(λ1, t1) . . . a(λn, tn))+ = a+(λn, tn) . . . a+(λ1, t1)

= a(λn, tn) . . . a(λ1, t1).

We will assume that the ground state Φ belongs to the set D; its

energy will be denoted by E0.

Definition 7.1. The Wightman function wn(λ1, t1, . . . , λn, tn) of the

Hamiltonian H is the expectation value of the operator

a(λ1, t1) . . . a(λn, tn)

in the ground state Φ of the Hamiltonian H:

wn(λ1, t1, . . . , λn, tn) = 〈a(λ1, t1) . . . a(λn, tn)Φ,Φ〉.

107

Page 155: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch07 page 108

108 Mathematical Foundations of Quantum Field Theory

Everything in this chapter can also be easily proved in the case

when a(λ) is a generalized operator function of the parameter λ.

Naturally, in this case, the Wightman function will be a generalized

function of the variables λ1, . . . , λn.

It is easy to prove the following properties of Wightman functions:

(1) Invariance with respect to time translation:

wn(λ1, t1 + τ, . . . , λn, tn + τ) = wn(λ1, t1, . . . , λn, tn).

(7.1)

(2) Hermiticity:

wn(λ1, t1, . . . , λn, tn) = wn(λn, tn, . . . , λ1, t1). (7.2)

(3) Positivity: Let fn(λ1, t1, . . . , λn, tn) denote a sequence of func-

tions that doesn’t vanish only for finitely many values of n. Let

us also assume that the functions belonging to this sequence do

not vanish only on finitely many values of λ1, t1, . . . , λn, tn. Then∑m,n,λi,µj ,ti,τj

fm(λ1, t1, . . . , λm, tm)fn(µ1, τ1, . . . , µn, τn)

×wm+n(λ1, t1, . . . , λm, tm, µn, τn, . . . , µ1, τ1) ≥ 0

(7.3)

(the sum is taken over λi, µj ∈ Λ and m,n = 0, 1, . . . ; −∞ <

ti, τj <∞).

(4) Spectral property:∫exp(−iωa)wn(λ1, t1, . . . , λk, tk, λk+1, tk+1

+ a, . . . , λn, tn + a)da = 0 (7.4)

if the number ω + E0 does not belong to the spectrum of the

Hamiltonian H (in particular, the relation (7.4) is true for every

ω < 0).

The proof of property (1) follows immediately from exp(−iHt)Φ = exp(−iE0t)Φ. Property (2) follows from 〈a(λ1, t1) . . .

Page 156: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch07 page 109

Wightman and Green Functions 109

a(λn, tn)Φ,Φ〉 = 〈Φ, a(λn, tn) . . . a(λ1, t1)Φ〉. Property (3) follows

from applying the inequality 〈AA+Φ,Φ〉 ≥ 0 to the operator

A =∑m,λi,ti

fm(λ1, t1, . . . , λm, tm)a(λ1, t1) . . . a(λm, tm),

where the sum is taken over m = 0, 1, . . . ; λi ∈ Λ; −∞ < ti <∞.

The fourth property follows from the relation∫exp(−iβa) 〈exp(iBa)Ψ1,Ψ2〉 da = 0,

which is correct if the number β does not belong to the spectrum of

the self-adjoint operator B (see Appendix A.5) (we should assume

that B = H, Ψ1 = UΦ, Ψ2 = V +Φ, where

U = a(λk+1, tk+1) . . . a(λn, tn), V = a(λ1, t1) . . . a(λk, tk)).

If the Hamiltonian H acts on a representation space of CR (on

the Fock space F (B), for example), we can consider as the family

of operators a(λ) the family a(k, ε) of annihilation and creation

operators that correspond to some orthonormal basis φk ∈ B (here,

k ∈M , ε = ±1, a(k, ε) = a(φk, ε)).

In other words, the Wightman functions of the Hamiltonian H

acting on a representation space of CR are functions

wn(k1, ε1, t1, . . . , kn, εn, tn)

= 〈a(k1, ε1, t1) . . . a(kn, εn, tn)Φ,Φ〉 .

In the case of CAR, the operators a(k, ε, t) are bounded and

hence the Wightman functions are well defined. In the case of CCR,

we assume that the ground state Φ belongs to the domain D of

operators a(k, ε) and that the domain D is invariant with respect to

the operators exp(iHt).

For Hamiltonians acting on a representation space of CR, we can

prove one more property of Wightman functions.

(5) Permutation of arguments: Let the functions w(i)n be the func-

tions obtained from wn by swapping (ki, εi, ti) and (ki+1, εi+1, ti+1)

Page 157: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch07 page 110

110 Mathematical Foundations of Quantum Field Theory

(with a change of sign in the case of CAR). If ti = ti+1, then

w(i)n (k1, ε1, t1, . . . , kn, εn, tn)

= wn(k1, ε1, t1, . . . , kn, εn, tn) +Aεiεi+1δki,ki+1

wn−2

× (k1, ε1, t1, . . . , ki−1, εi−1, ti−1, ki+2, εi+2, ti+2, . . . , kn, εn, tn)

(7.5)

(here, Aεε′ is the matrix defined in formula (6.3)).

To check this equation, we note that from the commutation

relations in the form (6.3), we can obtain

[a(k, ε, t), a(k′, ε′, t)]∓ = [a(k, ε), a(k′, ε′)]∓ = Aεε′δkk′ .

Wightman functions can be easily calculated for the Hamiltonian

H0 =∑ω(k)a+(k)a(k) in Fock space because for this Hamiltonian,

we have a(k, ε, t) = exp(iεω(k)t)a(k, ε), and hence,

wn(k1, ε1, t1, . . . , kn, εn, tn)

= exp(i(ε1ω(k1)t1 + · · ·+ εnω(kn)tn)) 〈a(k1, ε1)

. . . a(kn, εn)Φ,Φ〉 .

In the case of CCR, the operator H0 is bounded from below

when ωk ≥ 0; then Φ = θ (see Section 6.2) and the functions

w2(k1, ε1, t1, k2, ε2, t2) do not vanish only in the case when ε1 =

−1, ε2 = 1; in this case, we have

w2(k1 − 1, t1, k2 − 1, t2)

= δk1,k2 exp(−iω(k1)t1 + iω(k2)t2).

In the case of CAR, the ground state is Φ =∏ω(k)<0 a

+(k)θ. In

this case, the function w2(k1, ε1, t1, k2, ε2, t2) does not vanish only if

ε2ω(k2) > 0, ε1ω(k1) < 0; then

w2(k1, ε1, t1, k2, ε2, t2) = δk1,k2 exp(iε1ω(k1)t1 + iε2ω(k2)t2).

In other words,

w2(k1, ε1, t1, k2, ε2, t2)

= θ(ε2ω(k2))θ(−ε1ω(k1))δk1k2

× exp(i(ε1ω(k1)t1 + ε2ω(k2)t2)).

Page 158: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch07 page 111

Wightman and Green Functions 111

For Hamiltonians acting on the space of representations CR(B),

one can give a more invariant definition of Wightman functions that

does not depend on the choice of basis in the space B. Namely, for the

Wightman function wn(f1, ε1, t1, . . . , fn, εn, tn), where fi ∈ B, εi =

±1,−∞ < ti < ∞, we can consider the expectation value of the

operator a(f1, ε1, t1) . . . a(fn, εn, tn) with respect to the ground state

Φ of the Hamiltonian H (wn is a linear functional of f1, . . . , fn). If

the space B = L2(X), then we can write the operator a(f, ε) in the

form a(f, ε) =∫f(x)a(x, ε)dx, where a(x, ε) is a generalized operator

function (x ∈ X, ε = ±1). It is easy to check that

wn(f1, ε1, t1, . . . , fn, εn, tn)

=

∫f1(x1) . . . fn(xn)wn(x1, ε1, t1, . . . , xn, εn, tn)dx1 . . . dxn,

where wn(x1, ε1, t1, . . . , xn, εn, tn) is the Wightman function con-

structed by means of the generalized operator function a(x, ε) and

the Hamiltonian H.

7.2 Green functions

Let us define Green functions for Hamiltonians acting on the space

of representations of canonical commutation or anticommutation

relations. As in Section 7.1, we use the notation a(k, ε) = a(φk, ε),

where φk is an orthonormal basis in B.

Let us define the T -product of Heisenberg operators T (a(k1, ε1, t1)

. . . a(kn, εn, tn)) as the product of the operators a(k1, ε1, t1) . . .

a(kn, εn, tn) in chronological order (in order of decreasing times). In

the case of CAR, this product is taken with a minus sign if we need

an odd permutation to put the operators in chronological order; if

we need an even permutation, we take it with a plus sign. In other

words,

T (a(k1, ε1, t1) . . . a(kn, εn, tn))

= (−1)γa(ki1 , εi1 , ti1) . . . a(kin , εin , tin), (7.6)

where P = (i1, . . . , in) is the permutation obeying ti1 > . . . tin , γ = 0

in the case of CCR and γ is equal to the parity of the permutation P

Page 159: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch07 page 112

112 Mathematical Foundations of Quantum Field Theory

in the case of CAR (here, ki ∈ M , εi = ±1 and all the times ti are

distinct).

For the case of n = 2, we can write

T (a(k1, ε1, t1)a(k2, ε2, t2))

= θ(t1 − t2)a(k1, ε1, t1)a(k2, ε2, t2)

± θ(t2 − t1)a(k2, ε2, t2)a(k1, ε1, t1),

where we use the upper sign in the case of CCR and we use the lower

sign in the case of CAR.

The Green function of the Hamiltonian H is defined as an

expectation value of the T -product T (a(k1, ε1, t1) . . . a(kn, εn, tn))

with respect to the ground state Φ of the Hamiltonian H:

Gn(k1, ε1, t1, . . . , kn, εn, tn)

= 〈T (a(k1, ε1, t1) . . . a(kn, εn, tn))Φ,Φ〉.

Green functions can be easily expressed in terms of Wightman

functions, e.g.,

G2(k1, ε1, t1, k2, ε2, t2) = θ(t1 − t2)w2(k1, ε1, t1, k2, ε2, t2)

± θ(t2 − t1)w2(k2, ε2, t2, k1, ε1, t1).

The general expression of Green functions in terms of Wightman

functions can be written in the following way:

Gn(k1, ε1, t1, . . . , kn, εn, tn)

=∑π

(−1)γ(π)θπ(t)wπn(k1, ε1, t1, . . . , kn, εn, tn), (7.7)

where the sum is taken over all permutations π = (i1, i2,

. . . , in), θπ(t) = θ(ti1−ti2)θ(ti2−ti3) . . . θ(tin−1−tin) and the functions

wπn are defined by the relation

wπn(k1, ε1, t1, . . . , kn, εn, tn)

= w(ki1 , εi1 , ti1 , . . . , kin , εin , tin),

γ(π) = 0 in the case of CCR, and γ is equal to the parity of the

permutation π in the case of CAR.

Page 160: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch07 page 113

Wightman and Green Functions 113

The definition of the functions Gn(k1, ε1, t1, . . . , kn, εn, tn) is

meaningful only when all the times t1, . . . , tn are distinct. The set

where a Green function is not defined has measure zero, therefore, the

integral of a Green function multiplied by a conventional function of

t1, . . . , tn is well defined. This means, for example, that the function

Gn can be considered as a well-defined generalized function of the

variables t1, . . . , tn. We will need to specify the definition of a Green

function in the case when some of the arguments t1, . . . , tn coincide

only in one situation in Section 7.4. We will do this by requiring that

when the time variables coincide, the corresponding operators in the

T -product are written in normal order (i.e. the T -product is defined

with respect to such a permutation of indices (i1, . . . , in), such that

ti1 ≥ · · · ≥ tin and in the case of tik = til , we have εik ≥ εil).Let us formulate the following simple properties of Green

functions:

(1) Symmetry: Green functions Gn(k1, ε1, t1, . . . , kn, εn, tn) do not

change under permutation of variables in the case of CCR. In

the case of CAR, the Green function changes only by a sign

under a permutation of (ki, εi, ti) and (kj , εj , tj).

(2) Invariance with respect to time translation:

Gn(k1, ε1, t1 + a, . . . , kn, εn, tn + a)

= Gn(k1, ε1, t1, . . . , kn, εn, tn).

(3) Gn(k1, ε1, t+ 0, k2, ε2, t, k3, ε3, t3, . . . , kn, εn, tn)

∓Gn(k1, ε1, t1, k2, ε2, t+ 0, k3, ε3, t3, . . . , kn, εn, tn)

= δk1,k2Aε1ε2Gn−2(k3, ε3, t3, . . . , kn, εn, tn).

The first property follows from the fact that we can transpose

factors under the sign of T -product. The third property follows from

commutation relations. As for Wightman functions, one can give a

definition of Green functions which does not depend on the choice of

basis, namely, we should set

Gn(f1, ε1, t1, . . . , fn, εn, tn)

= 〈T (a(f1, ε1, t1) . . . a(fn, εn, tn))Φ,Φ〉

Page 161: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch07 page 114

114 Mathematical Foundations of Quantum Field Theory

(here, fi ∈ B, εi = ±1,−∞ < ti <∞). If B = L2(X), then

Gn(f1, ε1, t1, . . . , fn, εn, tn)

=

∫f1(x1) . . . fn(xn)Gn(x1, ε1, t1, . . . , xn, εn, tn)dx1 . . . dxn,

where Gn(x1, ε1, t1, . . . , xn, εn, tn) is a Green function constructed by

means of the generalized operator function a(x, ε).

It is easy to construct a perturbation series for Green functions as

well as a diagrammatic representation for this series. Let us describe

this construction.

Suppose that the Hamiltonian H is written in the form H =

H0 + gV . The ground states of the Hamiltonians H and H0 will be

denoted by Φ and Φ0 correspondingly. Let us use the relation

Φ = lima→0

C1(α)Sα(0,−∞)Φ0 = lima→0

C2(α)Sα(0,+α)Φ0 (7.8)

that follows from (4.11) (our goal is to get the representation of Green

functions in terms of perturbation series, therefore it is sufficient to

know that the relation (7.8) is true in the framework of perturbation

theory in the case when the ground states Φ0 of the Hamiltonian H0

is not degenerate). If (i1, . . . , in) is a permutation obeying ti1 > · · · >tin , then

Gn(k1, ε1, t1, . . . , kn, εn, tn)

= limα→0

C1(α)C2(α)× 〈a(ki1 , εi1 , ti1) . . . a(kin , εin , tin)

×Sα(0,−∞)Φ0, Sα(0,+∞)Φ0〉.

Let us introduce the operators

a(k, ε, t) = exp(iH0t)a(k, ε) exp(−iH0t).

It is easy to check that

a(k, ε, t) = exp(iHt) exp(−iH0t)a(k, ε, t) exp(iH0t) exp(−iHt)

= S(0, t)a(k, ε, t)S(t, 0).

For fixed t and α→ 0, we can write the approximate equations

S(t, 0) ≈ Sα(t, 0), S(0, t) ≈ Sα(0, t),

a(k, ε, t) ≈ Sα(0, t)a(k, ε, t)Sα(t, 0).

Page 162: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch07 page 115

Wightman and Green Functions 115

Using these formulas, we obtain

Gn(k1, ε1, t1, . . . , kn, εn, tn)

= limα→0

C1(α)C2(α)

× 〈Sα(0, ti1)a(ki1 , εi1 , ti1)Sα(ti1 , 0)Sα(0, ti2) . . .

× a(kin , εin , tin)Sα(tin , 0)Sα(0,−∞)Φ0, Sα(0,+∞)Φ0〉

= limα→0

〈AαΦ0,Φ0〉〈Sα(∞,−∞)Φ0,Φ0〉

,

where

Aα = Sα(∞, ti1)a(ki1 , εi1 , ti1)Sα(ti1 , ti2)

. . . a(kin , εin , tin)Sα(tin ,−∞). (7.9)

We have used the group property

Sα(t′′, t) = Sα(t′′, t′)Sα(t′, t),

and the relation

1 = 〈Φ,Φ〉 = limα→0

C1(α)C2(α) 〈Sα(0,−∞)Φ0, Sα(0,+∞)Φ0〉

= limα→0

C1(α)C2(α) 〈Sα(∞,−∞)Φ0,Φ0〉

in the derivation of (7.9). Equation (7.9) can be rewritten in the

following way:

Aα = T (a(k1, ε1, t1) . . . a(kn, εn, tn))

× exp

(1

i

∫ ∞−∞

g exp(−α|τ |)V (τ)dτ

)

=

∞∑n=0

1

n!

(gi

)n×∫T (a(k1, ε1, t1) . . . a(kn, εn, tn)

× exp(−α|τ1| − · · · − α|τn|)V (τ1) . . . V (τn))dnτ.

Page 163: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch07 page 116

116 Mathematical Foundations of Quantum Field Theory

One can use the short-hand notation

Aα = T (a(k1, ε1, t1) . . . a(kn, εn, tn)Sα(∞,−∞))

for the operator Aα.

Let us assume now that the operator H = H0 + gV is written in

the form (6.26) with the functions Vm,n that do not depend on time.

For the operator Sα(∞,−∞) corresponding to this Hamiltonian,

one can construct the diagram representation using the results of

Section 6.3. Using this representation, we can also construct a dia-

gram representation for the operator Aα. The diagram representation

for the numbers 〈AαΦ0,Φ0〉 and 〈Sα(∞,−∞)Φ0,Φ0〉 is obtained from

the diagrams of the operators Aα and Sα(∞,−∞) if we exclude from

the diagrams of these operators the diagrams where there are vertices

that do not belong to any edges. Using the fact that disconnected

diagrams decompose as a product of diagrams corresponding to

the connected components, it is easy to construct the diagram

representation for the expression

〈AαΦ0,Φ0〉〈Sα(∞,−∞)Φ0,Φ0〉

. (7.10)

It consists of the diagrams 〈AαΦ0,Φ0〉 that do not contain as a

component a diagram for 〈Sα(∞,−∞)Φ0,Φ0〉. To get a diagram

representation for a Green function Gn(k1, ε1, t1, . . . , kn, εn, tn), we

should take the limits α tending to zero in the diagrams for the

expression (7.10).

7.3 Kallen–Lehmann representation

Let us consider the Wightman function wn of the Hamiltonian H

with respect to the family of operators a(λ). It is easy to express

this function in terms of matrix entries of the operator a(λ) in the

basis of the eigenvectors of the Hamiltonian H. Let us prove such

an expression (called the Kallen–Lehmann representation) for the

function w2. Let us assume first that the operator H has discrete

spectrum, i.e. there exists an orthonormal basis Φn of eigenvectors

of the operator H with the corresponding eigenvalues denoted by En.

Page 164: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch07 page 117

Wightman and Green Functions 117

Then

w2(λ1, t1, λ2, t2) = 〈a(λ1, t1)a(λ2, t2)Φ,Φ〉

=⟨a(λ2, t2)Φ, a(λ1, t1)Φ

⟩=∑n

〈a(λ2, t2)Φ,Φn〉⟨

Φn, a(λ1, t1)Φ⟩

=∑n

〈exp(iHt2)a(λ2) exp(−iHt2)Φ,Φn〉

×⟨

Φn, exp(iHt1)a(λ1) exp(−iHt1)Φ⟩

=∑n

exp(i(En − E0)t2) 〈a(λ2)Φ,Φn〉

× exp(i(E0 − En)t1)⟨

Φn, a(λ1)Φ⟩.

Introducing the notation ρλn = 〈Φn, a(λ)Φ〉, we obtain the

representation

w2(λ1, t1, λ2, t2) =∑n

exp[−i(En − E0)(t1 − t2)]ρλ1n ρλ2n .

If the operator H additionally has continuous spectrum, then the

complete system of eigenvectors consists of normalized eigenvectors

Φn and generalized eigenvectors Φγ (we assume that the vectors

Φn are orthonormal and Φγ are δ-normalized; the corresponding

eigenvalues are denoted by En and Eγ). Repeating the above

calculations and using the relation

〈x, y〉 =∑n

〈x,Φn〉 〈Φn, y〉+

∫〈x,Φγ〉 〈Φγ , y〉 dγ,

we see that

w2(λ1, t1, λ2, t2) =∑n

exp[−i(En − E0)(t1 − t2)]ρλ1n ρλ2n

+

∫exp[−(Eγ − E0)(t1 − t2)]ρλ1γ ρ

−λ2γ dγ,

(7.11)

where

ρλn = 〈Φn, a(λ)Φ〉 , ρλγ = 〈Φγ , a(λ)Φ〉 .

Page 165: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch07 page 118

118 Mathematical Foundations of Quantum Field Theory

It is clear from the above formulas that the asymptotic behavior

of the function w2 as t1− t2 →∞ is determined only by the discrete

spectrum because the contribution of the continuous spectrum for

t1 − t2 →∞ is an integral of quickly oscillating functions.1

Let us note that the statement about the asymptotic behavior

of the function w2 should be modified in the case when a(λ) is a

generalized operator function.

The function w2 can also be expressed in terms of

σλ1,λ2(µ) = 〈a(λ1Eµa(λ)Φ,Φ)〉 ,

where Eµ = eµ(H) is the spectral decomposition of the operator H

(recall that eµ(x) = 1 for x ≤ µ, eµ(x) = 0 for x > µ), namely,

w2(λ1, t1, λ2, t2) =

∫exp(i(µ− E0)(t2 − t1))dσλ1,λ2(µ).

Indeed,

w2(λ1, t1, λ2, t2) = 〈a(λ2, t2)Φ, a(λ1, t1)Φ〉

= 〈exp(iHt2)a(λ2) exp(−iE0t2)Φ, exp(iHt1)

× a(λ1) exp(−iE0t1)Φ〉

= 〈exp[i(H − E0)(t2 − t1)]a(λ2)Φ, a(λ1)Φ〉

=

∫exp[i(µ− E0)(t2 − t1)]d〈Eµa(λ2)Φ, a(λ1)Φ〉

=

∫exp[i(µ− E0)(t2 − t1)]dσλ1,λ2(µ).

The Green function G2 of the operator acting on the representa-

tion space of CR can be expressed in terms of the function w2 by

means of the simple formula

G2(k1, ε1, t1, k2, ε2, t2) = θ(t1 − t2)w2(k1, ε1, t1, k2, ε2, t2)

± θ(t2 − t1)w2(k2, ε2, t2, k1, ε1, t1).

1A rigorous proof of this statement can be given under the assumption ofabsolute continuity of the continuous spectrum.

Page 166: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch07 page 119

Wightman and Green Functions 119

Representations for the function G2 follow immediately from repre-

sentations of w2. For example, for operators with a discrete spectrum,

we have

G2(k1, ε1, t1, k2, ε2, t2)

= θ(t1 − t2)∑n

exp[−i(En − E0)(t1 − t2)]ρk1,−ε1n ρk2,ε2n

± θ(t2 − t1)∑n

exp[−i(En − E0)(t2 − t1)]ρk2,−ε2n ρk1,ε1n

(here, ρk,εn = 〈Φn, a(k, ε)Φ〉 and Φn are the stationary states of the

Hamiltonian H and En are the corresponding eigenvalues).

Let us also consider the functions

w2(k1, ε1, ω1, k2, ε2, ω2)

= (2π)−1

∫exp[i(ε1ω1t1 + ε2ω2t2)]

w2(k1, ε1, t1, k2, ε2, t2)× dt1dt2;

G2(k1, ε1, ω1, k2, ε2, ω2)

= (2π)−1

∫exp[i(ε1ω1t1 + ε2ω2t2)]

×G2(k1, ε1, t1, k2, ε2, t2)dt1dt2

(the Wightman and Green functions in energy representations).

The representations of w2 and G2 imply representations of the

functions w2 and G2. In particular, for operators with discrete

spectrum, we have

G2(k1, 1, ω1, k2,−1, ω2) = G(k1, k2, ω1)δ(ω1 − ω2),

where

G(k1, k2, ω) =

∫exp(iωτ)G2(k1, 1, τ, k2,−1, 0)dτ

= i∑n

ρk1,−1n ρk2,−1

n

ω − (En − E0) + i0∓∑n

ρk2,1n ρk1,−1n

ω + (En − E0)− i0.

(7.12)

Page 167: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch07 page 120

120 Mathematical Foundations of Quantum Field Theory

Hence, the poles of the function G with respect to the variable ω

correspond to points in the discrete spectrum. If H has continuous

spectrum, we should add an integral over the continuous spectrum

to the expression (7.12). In all physical examples, the contribution

of the continuous spectrum is a continuous function of ω.

7.4 The equations for Wightman and Green functions

Let us consider the Hamiltonian H acting on the representation space

of CR (B). We will represent this Hamiltonian in normal form:

H =∑m,n

∑ki,lj

Γm,n(k1, . . . , km|l1, . . . , ln)a+k1. . . a+

kmal1 . . . aln ,

where a+k = a+(φk), ak = a(φk), φk is an orthonormal basis in

B, the functions Γm,n are symmetric with respect to variables

ki and lj in the case of CCR and antisymmetric in the case of

CAR. Then the Heisenberg equations for the operators ak(t) =

exp(iHt)ak exp(−iHt), a+k (t) = exp(iHt)a+

k exp(−iHt) can be writ-

ten in the form1

i

dak(t)

dt= [H, ak(t)]

= −∑

m,n,k1,...,km−1,l1,...,ln

mΓm,n(k, k1, . . . , km−1|l1, . . . , ln)

× a+k1

(t) . . . a+km−1

(t)al1(t) . . . aln(t);

1

i

da+k (t)

dt= [H, a+

k (t)]

=∑

m,n,k1,...,km−1,l1,...,ln

nΓm,n(k, k1, . . . , km|l1, . . . , ln−1, k)

× a+k1

(t) . . . a+km

(t)al1(t) . . . aln−1(t).

These equations immediately imply equations for Wightman

functions

wn(k1, ε1, t1, . . . , kn, εn, tn).

Let us calculate, for example, the expression of the derivative of

the function wn with respect to the variable t1 in terms of Wightman

Page 168: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch07 page 121

Wightman and Green Functions 121

functions (the expressions for the derivatives of wn with respect to

ti have similar form). It is easy to see that

1

i

∂t1wr(k1, ε1, t1, . . . , kr, εr, tr)

= −δε1−1

∑m,n,pi,qj

mΓm,n(k1, p1, . . . , pm−1|q1, . . . , qn)

×wr+m+n−1(p1, 1, t1, . . . , pm−1, 1, t1, q1,−1, t1,

. . . , qn,−1, t1, k2, ε2, t2, . . . , kr, εr, tr)

+ δε1+1

∑m,n,pi,qj

nΓm,n(p1, . . . , pm|q1, . . . , qn−1, k1)

×wr+m+n−1(p1, 1, t1, . . . , pm, 1, t1, q1,−1, t1,

. . . , qn−1,−1, t1, k2, ε2, t2, . . . , kr, εr, tr). (7.13)

It is easy to derive the equations for Green functions from (7.13) or

directly from Heisenberg equations. Let us calculate, for example,

∂t1G2(k1, ε1, t1, k2, ε2, t2)

=∂

∂t1[w2(k1, ε1, t1, k2, ε2, t2)θ(t1 − t2)

±w2(k2, ε2, t2, k1, ε1, t1)θ(t2 − t1)]

= θ(t1 − t2)∂

∂t1w2(k1, ε1, t1, k2, ε2, t2)

± θ(t2 − t1)∂

∂t1w2(k2, ε2, t2, k1, ε1, t1)

+ δ(t1 − t2)[w2(k1, ε1, t1, k2, ε2, t2)

∓w2(k2, ε2, t2, k1, ε1, t1)].

Using CR (or property (5) of Wightman functions), we can

simplify the last summand:

δ(t1 − t2)(w2(k1, ε1, t1, k2, ε2, t2)∓ w2(k2, ε2, t2, k1, ε1, t1))

= Aε1ε2δ(t1 − t2)δk1,k2 .

Page 169: Mathematical Foundations of Quantum Field Theory

March 26, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch07 page 122

122 Mathematical Foundations of Quantum Field Theory

The first two summands in the formula for ∂G2∂t1

can be expressed

in terms of Green functions (one should use (7.13)). As a result, we

obtain the equation

1

i

∂t1G2(k1, ε1, t1, k2, ε2, t2) =

1

iδ(t1 − t2)Aε1ε2δk1,k2

− δε1−1

∑m,n,pi,qj

mΓm,n(k1, p1, . . . , pm−1|q1, . . . , qn)

×Gm+n(p1, 1, t1, . . . , pm−1, 1, t1, q1,−1, t1,

. . . , qn,−1, t1, k2, ε2, t2)

+ δε1+1

∑m,n,pi,qj

nΓm,n(p1, . . . , pm|q1, . . . , qn−1, k1)

×Gm+n(p1, 1, t1, . . . , pm, 1, t1, q1,−1, t1,

. . . , qn−1,−1, t1, k2, ε2, t2).

Similarly, one can derive the equations for Green functions Gn. We

will not write down the form of these equations since, in more general

situations, they are very complicated.

Page 170: Mathematical Foundations of Quantum Field Theory

March 27, 2020 8:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch08 page 123

Chapter 8

Translation-Invariant Hamiltonians

8.1 Translation-invariant Hamiltonians in Fock space

Let us consider operators acting on the Fock space F (B), where

B = L2(E3) is the space of square-integrable functions ψ(k) on the

Euclidean space E3 with variable k denoting the momentum.1

It is convenient to express these operators in terms of operator

generalized functions a+(k), a(k) satisfying the commutation rela-

tions

[a(k), a(k′)]∓ = [a+(k), a+(k′)]∓ = 0,

[a(k), a+(k′)]∓ = δ(k− k′).

These operator generalized functions were defined in Section 3.2.

The Hamiltonian H is called translation-invariant if it commutes

with the momentum operators P =∫ka+(k)a(k)dk. An example

of a translation-invariant Hamiltonian is the Hamiltonian for a

system of interacting non-relativistic identical particles considered

in Section 3.2.

We point out the following easy-to-check statement: the vac-

uum vector θ is always an eigenvector of a translation-invariant

Hamiltonian H.

1The case B = L2(E3) corresponds to identical spinless particles. If we considerparticles with spin, or if we have several types of particles, we should considerB = L2(E3 × B), where B is a finite set. It is easy to generalize all the resultsof the present chapter and the following chapters to this case, hence there is noneed to describe the generalization in detail. It is important to note, however, thatfermions always have spin, therefore, in the case of CAR, the situation B = L2(E3)that we consider does not correspond to any particle existing in nature.

123

Page 171: Mathematical Foundations of Quantum Field Theory

March 27, 2020 8:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch08 page 124

124 Mathematical Foundations of Quantum Field Theory

To prove this statement, we note that the momentum operator P

has precisely one (up to a factor) eigenvector θ that corresponds to

a zero eigenvalue (here, we consider conventional — normalized —

eigenvectors; of course, P has generalized eigenvectors). Since we

have assumed that the operator H commutes with the operator

P, it transforms the single eigenvector of the operator P into an

eigenvector of P. Therefore, we see that Hθ = λθ, which proves our

statement.

Let us assume that the Hamiltonian H is written in the normal

form

H =∑m,n

∫Γm,n(p1, . . . ,pm|q1, . . . ,qn)a+(p1) . . . a+(pm)a(q1)

. . . a(qn)dmpdnq. (8.1)

It is easy to check that it is translation-invariant only in the case

when the functions Γm,n have the form

Γm,n(p1, . . . ,pm|q1, . . . ,qn) = Λm,n(p1, . . . ,pm|q1, . . . ,qn)

× δ(p1 + · · ·+ pm − q1 − · · · − qn).

The statement we have proven implies that expression (8.1) cannot

define a self-adjoint operator on Fock space if for some m the function

Γm,0 6≡ 0 (in physical terms, the Hamiltonian H generates vacuum

polarization in this case).2

The above statement implies that we should consider translation-

invariant operators with vacuum polarization outside the Fock space

framework. It is not that we should consider these Hamiltonians to

be bad, but that Fock space is too narrow for these Hamiltonians.3

We see that we should consider translation-invariant Hamil-

tonian (8.1) as a formal expression, composed of the symbols

a+(k), a(k). We will show how we can construct various physical

2Indeed, in the case of vacuum polarization, the expression (8.1) does notdetermine an operator on Fock space at all (see Section 3.2 for details).

3To construct an operator on Fock space corresponding to a translation-invariantHamiltonian with vacuum polarization, we should make the volume or, in otherwords, the infrared cutoff (see below).

Page 172: Mathematical Foundations of Quantum Field Theory

March 27, 2020 8:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch08 page 125

Translation-Invariant Hamiltonians 125

quantities for such Hamiltonians. In particular, in Chapter 9, we will

give a definition for the scattering matrix of a translation-invariant

Hamiltonian.

We will use the following construction. Let H be a translation-

invariant Hamiltonian, i.e. a formal expression of the form

H =∑m,n

∫Λm,n(k1, . . . ,km|l1, . . . , ln)

× δ(k1 + · · ·+ km − l1 − · · · − ln)

× a+(k1) . . . a+(km)a(l1) . . . a(ln)dmkdnl. (8.2)

We will assume that the expression for H is formally Her-

mitian (i.e. we assume that Λm,n(k1, . . . ,km|l1, . . . , ln) =

Λn,m(ln, . . . , l1|km, . . . ,k1)). Let us define a Hamiltonian H with

volume cutoff Ω. We will denote by Ω the cube with edge length

L in coordinate space (0 ≤ x ≤ L, 0 ≤ y ≤ L, 0 ≤ z ≤ L). Then,

we denote by BΩ = L2(Ω) the space of square-integrable functions

ψ(r), where r ∈ Ω, and by FΩ = F (BΩ), we denote the Fock space

constructed with the Hilbert space BΩ. In the space BΩ, we select an

orthonormal basis of functions φk = L−3/2 exp(−ikr), where k runs

over the lattice TΩ with step size 2πL (i.e. k = 2π

L n, where n is an

integer vector). The operators a+(φk), a(φk) in FΩ corresponding to

this basis will be denoted by a+k , ak.

The operator HΩ on the space FΩ will be defined by the formula

HΩ =∑m,n

∑ki,lj

(2π

L

) 32

(m+n−2)

Λm,n(k1, . . . ,km|l1, . . . , ln)

× δk1+···+km,l1+···+lna+k1. . . a+

kma+l1. . . a+

ln(8.3)

(in other words, we obtain the expression for HΩ from the expression

for H by replacing a+(k), a(k) with ( L2π )3/2a+k , (

L2π )3/2ak, replacing

the integration with summation over the lattice TΩ and multiply-

ing by (2πL )3, and replacing the function δ(k1 + · · ·+km−l1−· · ·−ln)

by ( L2π )3δk1+···+km,l1+···+ln).

Page 173: Mathematical Foundations of Quantum Field Theory

March 27, 2020 8:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch08 page 126

126 Mathematical Foundations of Quantum Field Theory

Let us assume that the formula (8.3) specifies a self-adjoint

operator on the space FΩ. If the functions Λm,n are “good” enough,

then this assumption holds. For example, in the case of CAR, one

can prove the following statements.

If the coefficient functions Λm,n tend to zero at infinity faster

than q−3(m+n)−ε where q = (k21 + · · · + k2

m + l21 + · · · + l2n)1/2, ε > 0,

and Λm,n = 0 for m + n ≥ s, then the formula (8.3) specifies a

bounded self-adjoint operator in the space FΩ. If the Hamiltonian H

is represented in the form H0 + V , where H0 =∫ω(k)a+(k)a(k)dk,

and the coefficient functions of the expression V tend to zero at

infinity sufficiently fast, then the formula (8.3) specifies a self-adjoint

operator4 in the space FΩ.

The proof of these statements can be obtained from the consid-

erations at the end of Section 6.2.

Note that the operator HΩ commutes with the operator PΩ =∑k∈TΩ

ka+k ak (the momentum operator).

We will define the physical quantities corresponding to the formal

Hamiltonian H by taking the limit Ω→∞.

For example, let us say that the number E belongs to the

spectrum (it is an energy level) of the formal Hamiltonian H, if

one can find such eigenvalues EΩ of the Hamiltonian HΩ, such that

limΩ→∞(EΩ − E0Ω) = E (here, E0

Ω denotes the energy of the ground

state ΦΩ of the Hamiltonian H). We will say that the Hamiltonian

H has an energy level E with momentum k, if there exist vectors

ΨΩ ∈ FΩ satisfying the conditions HΩΨΩ = EΩΨΩ;PΩΨΩ =

kΩΨΩ; limΩ→∞(EΩ − E0Ω) = E; limΩ→∞ kΩ = k.

Another way to analyze a translation-invariant Hamiltonian

is based on the construction of an operator realization of the

Hamiltonian.

Let us suppose that commuting self-adjoint operators H and P =

(P1, P2, P3) (energy operator and momentum operator) act on the

Hilbert space H. Let us assume further that the operator functions

a(k, ε, t) are generalized functions with respect to the variable k, obey

4In the case of CCR, with the given assumptions on the coefficient functions,one can prove only that the formula (8.3) defines a Hermitian operator.

Page 174: Mathematical Foundations of Quantum Field Theory

March 27, 2020 8:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch08 page 127

Translation-Invariant Hamiltonians 127

CR and act on the same space (here, k ∈ E3, ε = ±1, a(k, 1, t) =

a+(k,−1, t)).

We will say that the operators H, P, the operator generalized

functions a(k, ε, t), and the vector Φ ∈ H constitute an operator

realization of the formal Hamiltonian (8.2), if

(1) The operator generalized functions a(k, ε, t) satisfy the Heisen-

berg equations that formally correspond to the Hamilto-

nian (8.2):

i∂a(k,−1, t)

∂t=∑m,n

m

∫Λm,n(k,k1, . . . ,km−1|l1, . . . , ln)

× δ(k + k1 + · · ·+ km−1 − l1 − · · · − ln)

× a(k1, 1, t) . . . a(km−1, 1, t)a(l1,−1, t)

. . . a(ln,−1, t)dm−1kdnl; (8.4)

i∂a(k, 1, t)

∂t= −

∑m,n

n

∫Λm,n(k1, . . . ,km|l1, . . . , ln−1,k)

× δ(k1 + · · ·+ km − l1 − · · · − ln−1 − k)

× a(k1, 1, t) . . . a(km, 1, t)a(l1,−1, t)

. . . a(ln−1, 1, t)dmkdn−1l. (8.5)

(2) exp(iτH)a(k, ε, t) exp(−iτH) = a(k, ε, t+ τ);

exp(iαP)a(k, ε, t) exp(−iαP) = exp(iαkε)a(k, ε, t).

(3) The operators a(f, ε, t) =∫f(k)a(k, ε, t)dk, where f ∈ S(E3)

are defined on a dense subset D of the space H and transform

this subset into itself; if Ψ1,Ψ2 ∈ D, then 〈a(f, ε, t)Ψ1,Ψ2〉continuously depends on f ∈ S(E3). The operators a(f, ε, t) also

satisfy CCR for fixed t.

(4) The vector Φ is a ground state of the operator H and satisfies

the condition HΦ = PΦ = 0.

(5) The vector Φ is a cyclic vector of the family of operators a(f, ε, t).

Page 175: Mathematical Foundations of Quantum Field Theory

March 27, 2020 8:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch08 page 128

128 Mathematical Foundations of Quantum Field Theory

Note that it follows from condition (3) that expressions of the

form ∫f(k1, . . . ,kn)a(k1, ε1, t1) . . . a(kn, εn, tn)dk1 . . . dkn,

where f ∈ S(E3n) make sense as operators defined on the set D. This

can be proven by means of an operator analog of the kernel theorem

(see Appendix A.5). This remark allows us to define rigorously

the right-hand side of (8.5) in the case when the function Λ1,1(k)

is smooth and all of its derivatives grow no faster than a power

of k and the other functions Λm,n belong to the space S. The

derivative ∂a(k,ε,t)∂t on the left-hand side is understood as a weak

derivative.

The Wightman and Green functions for the operator realizations

of translation-invariant Hamiltonians can be defined by the formulas

wn(k1, ε1, t1, . . . ,kn, εn, tn) = 〈a(k1, ε1, t1) . . . a(kn, εn, tn)Φ,Φ〉;

Gn(k1, ε1, t1, . . . ,kn, εn, tn) = 〈T (a(k1, ε1, t1) . . . a(kn, εn, tn))Φ,Φ〉.

It would be more precise to say that the functions wn are Wightman

functions in (k, t)-representations. Their Fourier transforms with

respect to the variables k1, . . . ,kn are the functions

wn(x1, ε1, t1, . . . ,xn, εn, tn) = (2π)−32n

∫exp

(i∑

εjxjkj

)× wn(k1, ε1, t1, . . . ,kn, εn, tn)dnk

which are Wightman functions in (x, t)-representations. Their

Fourier transforms with respect to the variables t1, . . . , tn are

given by

wn(k1, ε1, ω1, . . . ,kn, εn, ωn) = (2π)−12n

∫exp

(i∑

εjωjtj

)× wn(k1, ε1, t1, . . . ,kn, εn, tn)dnt

which are Wightman functions in the (k, ω)-representation. Anal-

ogously, we can define Green functions in (x, t) and (k, ω)-

representations; we denote them by Gn and Gn.

Page 176: Mathematical Foundations of Quantum Field Theory

March 27, 2020 8:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch08 page 129

Translation-Invariant Hamiltonians 129

In the following sections of this chapter, we define the Wightman

and Green functions of translation-invariant Hamiltonians by means

of taking the limit of Wightman and Green functions in finite volume.

We will relate these definitions to the definitions of Wightman and

Green functions by means of operator realization of the translation-

invariant Hamiltonian. We will show also how one can construct the

operator realization by taking the limit when the volume tends to

infinity. In Chapter 11, we consider the construction of operator

realization in perturbation theory.

Note that for translation-invariant Hamiltonians, we can

write down the Kallen–Lehmann representation for functions

w2(k1, ε1, t1,k2, ε2, t2) and G2(k1, ε1, t1,k2, ε2, t2).

To derive this representation, we fix a system of generalized

eigenfunctions Ψλ of the operators H and P:

HΨλ = E(λ)Ψλ;

PΨλ = k(λ)Ψλ.

Let us prove, first, that the generalized function 〈a(k, ε, t)Φ,Ψλ〉 can

be represented by the formula

〈a(k, ε, t)Φ,Ψλ〉 = exp(iE(λ)t)δ(εk− k(λ))ρ(ε, λ). (8.6)

Indeed,

〈a(k, ε, t)Φ,Ψλ〉 = 〈exp(iHt)a(k, ε)Φ,Ψλ〉

= 〈a(k, ε)Φ, exp(−iHt)Ψλ〉

= exp(iE(λ)t) 〈a(k, ε)Φ,Ψλ〉 .The function 〈a(k, ε)Φ,Ψλ〉 satisfies the relation

exp(iεαk) 〈a(k, ε)Φ,Ψλ〉 = exp(iαk(λ)) 〈a(k, ε)Φ,Ψλ〉 (8.7)

that can be derived using the transformations

〈exp(iεαk)a(k, ε)Φ,Ψλ〉 = 〈exp(iαp)a(k, ε) exp(−iαp)Φ,Ψλ〉

= 〈a(k, ε)Φ, exp(−iαpΨλ)〉

= exp(iαk(λ)) 〈a(k, ε)Φ,Ψλ〉 .It follows from (8.7) that the function 〈a(k, ε)Φ,Ψλ〉 has the form

ρ(ε, λ)δ(εk− k(λ)); this gives the proof of the statement we need.

Page 177: Mathematical Foundations of Quantum Field Theory

March 27, 2020 8:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch08 page 130

130 Mathematical Foundations of Quantum Field Theory

Using (8.6), we can write down the following representation of

Wightman functions:

w2(k1, ε1, t1,k2, ε2, t2) = 〈a(k1, ε1, t1)a(k2, ε2, t2)Φ,Φ〉

= 〈a(k2, ε2, t2)Φ, a(k1,−ε1, t1)Φ〉

=

∫〈a(k2, ε2, t2)Φ,Ψλ〉〈a(k1,−ε1, t1)Φ,Ψλ〉dλ

=

∫exp(iE(λ)(t2 − t1))δ(ε1k1 + ε2k2)

× δ(ε1k1 + k(λ))ρ(ε2, λ)ρ(−ε1, λ)dλ. (8.8)

The relation

G2(k1, ε1, t1,k2, ε2, t2) = θ(t1 − t2)w2(k1, ε1, t1,k2, ε2, t2)

±θ(t2 − t1)w(k2, ε2, t2,k1, ε1, t1),

and (8.8) implies the representation of the function G2 in terms of the

functions ρ(ε, λ) (the Kallen–Lehmann representation). In particular,

G2(k1, 1, t1,k2,−1, t2) = G(k1, t1 − t2)δ(k1 − k2),

where

G(k, t) = θ(t)

∫exp(−iE(λ)t)δ(k + k(λ))|ρ(−1, λ)|2dλ

±θ(−t)∫

exp(iE(λ)t)δ(−k + k(λ))|ρ(1, λ)|2dλ. (8.9)

In the (k, ω)-representation, we have

G2(k1, 1, ω1,k2,−1, ω2) = G(k1, ω1)δ(ω1 − ω2)δ(k1 − k2),

where

G(k, ω) =

∫exp(iωt)G(k, t)dt

= i

∫|ρ(−1, λ)|2

ω − E(λ) + i0δ(k + k(λ))dλ

∓i∫

|ρ(+1, λ)|2

ω + E(λ)− i0δ(−k + k(λ))dλ. (8.10)

Page 178: Mathematical Foundations of Quantum Field Theory

March 27, 2020 8:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch08 page 131

Translation-Invariant Hamiltonians 131

8.2 Reconstruction theorem

The Wightman functions wn(k1, ε1, t1, . . . ,kn, εn, tn) of a translation-

invariant Hamiltonian H can be defined by the following relation:

wn(k1, ε1, t1, . . . ,kn, ε, tn)

= limΩ→∞

(L

) 32n

wΩn (k1, ε1, t1, . . . ,kn, εn, tn) (8.11)

(here, wΩn (k1, ε1, t1, . . . ,kn, εn, tn) = 〈aε1k1

(t1) . . . aεnkn(tn)ΦΩ,ΦΩ〉 are

Wightman functions of the Hamiltonian HΩ, constructed with

the basis φk, where k ∈ TΩ). The relation (8.11) requires some

explanation because the functions of continuous argument are defined

as the limit of functions of arguments running over a lattice. We will

understand this limit in the sense of generalized functions: for every

test function φ(k1, . . . ,kn), we require∫φ(k1, . . . ,kn)wn(k1, ε1, t1, . . . ,kn, εn, tn)dk1 . . . dkn

= limΩ→∞

(2π

L

) 3n2 ∑

k1,...,kn∈TΩ

φ(k1, . . . ,kn)

×wΩn (k1, ε1, t1, . . . , bkn, εn, tn). (8.12)

We assume that the limit (8.11) exists (the existence of this limit

can be checked in the framework of perturbation theory).

The functions wn(k1, ε1, t1, . . . ,kn, εn, tn) will be considered as

generalized functions with respect to the variables k1, . . . ,kn and

conventional functions with respect to the variables t1, . . . , tn.

Green functions Gn(k1, ε1, t1, . . . ,kn, εn, tn) of the translation-

invariant Hamiltonian H are defined in the same way, as limits

of finite-volume Green functions. Namely, we use the following

definition:

Gn(k1, ε1, t1, . . . ,kn, εn, tn)

= limΩ→∞

(L

) 3n2

GΩn (k1, ε1, t1, . . . ,kn, εn, tn),

Page 179: Mathematical Foundations of Quantum Field Theory

March 27, 2020 8:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch08 page 132

132 Mathematical Foundations of Quantum Field Theory

where GΩn (k1, ε1, t1, . . . ,kn, εn, tn) = 〈T (aε1k1

(t1) . . . aεnkn(tn))ΦΩ,ΦΩ〉are Green functions of the Hamiltonian HΩ constructed with the

basis φk, where k ∈ TΩ. The limit is understood in the same way as

the limit in the definition of Wightman functions.

Green functions Gn can be expressed in terms of Wightman

functions wn. Namely, taking the limit of Ω→∞ in the relation (7.7)

applied to the functions GΩn and wΩ

n , we obtain

Gn(k1, ε1, t1, . . . ,kn, εn, tn)

=∑π

(−1)γ(π)θπ(t)wπn(k1, ε1, t1, . . . ,kn, εn, tn) (8.13)

(the notation is the same as in Section 7.2).

The definition of Green functions in (x, t)-representation Gn and

in the (k, ω)-representation Gn is analogous to similar definitions for

Wightman functions.

It is easy to check that the Wightman functions wn(k1, ε1, t1,

. . . ,kn, εn, tn) of a translation-invariant Hamiltonian H have the

following properties:

(1) Invariance with respect to time translation:

wn(k1, ε1, t1, . . . ,kn, εn, tn)= wn(k1, ε1, t1+τ, . . . ,kn, εn, tn+τ).

(2) Hermiticity:

wn(k1, ε1, t1, . . . ,kn, εn, tn) = wn(kn,−εn, tn, . . . ,k1,−ε1, t1).

(3) Positive definiteness: For every sequence fn(k1, ε1, t1,

. . . ,kn, εn, tn) of test functions that do not vanish for only a

finite number of indices n, we have∑m,n

∑εα,σβ

∫fm(k1, ε1, t1, . . . ,km, εm, tm)fn(qn,−σn, τn,

. . . ,q1,−σ1, τ1)wm+n(k1, ε1, t1, . . . ,km, εm, tm,

q1, σ1, τ1, . . . ,qn, σn, τn)dmkdnqdmtdnτ ≥ 0.

Page 180: Mathematical Foundations of Quantum Field Theory

March 27, 2020 8:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch08 page 133

Translation-Invariant Hamiltonians 133

(4) Spectral property:∫exp(−iωa)wn(k1, ε1, t1, . . . ,kr, εr, tr,kr+1, εr+1, tr+1 + a,

. . . ,kn, εn, tn + a)da = 0, (8.14)

if ω < 0. The relation (8.14) is also valid under a weaker

condition — it is sufficient to assume that ω does not belong

to the spectrum of the Hamiltonian H.

(5) Symmetry with respect to the permutation of arguments: Let us

introduce the notation

w(i)n (k1, ε1, t1, . . . ,ki, εi, ti,ki+1, εi+1, ti+1, . . . ,kn, εn, tn)

= ±wn(k1, ε1, t1, . . . ,ki+1, εi+1, ti+1,ki, εi, ti,

. . . ,kn, εn, tn)

(here, the plus sign corresponds to the case of CCR and the

minus sign corresponds to the case of CAR). In other words, w(i)n

is obtained from wn by permuting ki, εi, ti with ki+1, εi+1, ti+1

(with a change of sign in the case of CAR).

If ti = ti+1, then

w(i)n (k1, ε1, t1, . . . ,kn, εn, tn)

= wn(k1, ε1, t1, . . . ,kn, εn, tn) +Aεε′δ(ki − k′i)

× wn−2(k1, ε1, t1, . . . ,ki−1, εi−1, ti−1,ki+2, εi+2, ti+2,

. . . ,kn, εn, tn) (8.15)

(the definition of the matrix Aε′ε can be found in Section 6.1).

(6) Translation-invariance:

wn(k1, ε1, t1, . . . ,kn, εn, tn)

= vn(k1, ε1, t1, . . . ,kn, εn, tn)δ(∑

εjkj

)(in the (x, t)-representation this property takes the form

wn(x1, ε1, t1, . . . ,xn, εn, tn) = wn(x1+a, ε1, t1, . . . ,xn+a, εn, tn)).

Page 181: Mathematical Foundations of Quantum Field Theory

March 27, 2020 8:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch08 page 134

134 Mathematical Foundations of Quantum Field Theory

All the above listed properties (with the exception of (6)) can

be easily proven if we recall that wn are obtained as limits of the

functions wΩn , which have similar properties proven in Section 7.1.

Theorem 8.1. (Reconstruction theorem.)

Existence: Given a family of functions wn(k1, ε1, t1, . . . ,

kn, εn, tn) obeying properties (1) − (6), one can construct a Hilbert

space H, four commuting self-adjoint operators H, P, operator func-

tions a(k, ε, t) acting on H that are generalized 5 functions with

respect to the variable k, and the vector Φ ∈ H in such a way that

the following requirements are met:

(a) wn(k1, ε1, t1, . . . ,kn, εn, tn) = 〈a(k1, ε1, t1) . . . a(kn, εn, tn) Φ,Φ〉.(b) exp(iτH)a(k, ε, t) exp(−iτH) = a(k, ε, t+ τ),

exp(iαp)a(k, ε, t) exp(−iαp) = exp(iεαk)a(k, ε, t);

(c) The operators a(f, ε, t) =∫f(k)a(k, ε, t)dk are defined on the

dense set D of the space H and transform the subset into itself;

the operators a(f, ε, t) for fixed t specify a representation of CR;

the expression 〈a(f, ε, t)Ψ1,Ψ2〉 continuously depends on f in the

topology of the space S for all Ψ1,Ψ2 ∈ D.

(d) The vector Φ is the ground state of the operator H and satisfies

the conditions HΦ = 0 and PΦ = 0.

(e) The vector Φ is a cyclic vector with respect to the operators

a(f, ε, t).

Uniqueness: If Hi, Hi,Pi, ai(k, e t) and Φi are two sets of objects

(i = 1, 2) satisfying the conditions (a), (b), and (c) of the

reconstruction theorem, then there exists a unitary operator U that

maps the space H1 onto the space H2 and obeys the conditions

UΦ1 = Φ2, UH1 = H2U,UP1 = P2U , as well as the relations

Ua1(f, ε, t) = a2(f, ε, t)U on some dense subset of the space H1 (in

other words, these two sets are isomorphic).

Let us start the proof from the second part of the theo-

rem. Let us consider the set N of sequences of functions f =

5We will consider the space of test functions to be S(E3) (i.e. all functions f inthe formulation of the theorem satisfy f ∈ S(E3)).

Page 182: Mathematical Foundations of Quantum Field Theory

March 27, 2020 8:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch08 page 135

Translation-Invariant Hamiltonians 135

fn(k1, ε1, t1, . . . ,kn, εn, tn) such that (1) every function fn is a

linear combination of functions of the form

λ1(k1, ε1)δ(t1 − τ1) . . . λm(km, εm)δ(tm − τm),

where λi(ki, εi) are test functions and (2) only a finite number of

functions fn do not vanish. Let us assign the vectors Ψif , where i =

1, 2, to every sequence f ∈ N by the formula

Ψif =

∞∑n=1

∑ε1,...,εn

∫fn(k1, ε1, t1, . . . ,kn, εn, tn)

× ai(k1, ε1, t1) . . . ai(kn, εn, tn)dnkdnt.

The set of vectors Ψif , where f ∈ N , will be denoted by Di. The

inner product 〈Ψif ,Ψ

ig〉 of two vectors from the set Di can be easily

expressed in terms of the sequences f, g and Wightman functions:⟨Ψif ,Ψ

ig

⟩=∑m,n

∑εα,σα

∫fm(k1, ε1, t1, . . . ,km, εm, Lm)fn(q1, σ1, τ1, . . . ,

qn, σn, τn)wm+n(k1, ε1, t1, . . . ,km, εm, tm,qn,−σn,

τn, . . . ,q1,−σ1, τ1)dmkdnqdmtdnτ. (8.16)

It is easy to check that

ai(φ)Ψif = Ψi

φf (8.17)

(here, φ = φ(k, ε, t) = λ(k, ε)δ(t−τ), where λ(k, ε) is a test function,

ai(φ) =∑

ε

∫φ(k, ε, t)ai(k, ε, t)dkdt =

∑ε

∫λ(k, ε)a(k, ε, τ))dk,

and φf ∈ N is a sequence whose nth entry is equal to

φ(k1, ε1, t1)fn−1(k2, ε2, t2, . . . ,kn, εn, tn)). Condition (b) implies that

exp(−iHiτ)Ψif = Ψi

Vτf ; exp(−iαP)Ψif = Ψi

Wαf , (8.18)

where Vτf ∈ N and Wαf ∈ N are sequences of functions with the

nth entry given by

fn(k1, ε1, t1 − τ, . . . ,kn, εn, tn − τ),

exp(−i∑

αkjεj

)fn(k1, ε1, t1, . . . ,kn, εn, tn).

Page 183: Mathematical Foundations of Quantum Field Theory

March 27, 2020 8:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch08 page 136

136 Mathematical Foundations of Quantum Field Theory

Let us now construct the operator U on the set D1 assuming that

UΨ1f = Ψ2

f .

It follows from (8.16) that this operator preserves inner products,

therefore it can be extended by continuity to the unitary operator

mapping H1 into H2. It follows from the relations (8.17) and (8.18)

that the constructed operator obeys the conditions we need.

The above proof of the second part of the theorem gives us a path

to proving the first part of the theorem.

Let us start with the sequence of Wightman functions w. Let us

introduce the inner product 〈f, g〉 on the set N by means of the

formula

〈f, g〉 =∑m,n

∑εα,σβ

∫fm(k1, ε1, t1, . . . ,kn, εn, tn)fn(q1, σ1, τ1, . . . ,

qn, σn, τn)wm+n(k1, ε1, t1, . . . ,km, εm, tm,qn,−σn, τn, . . . ,

q1,−σ1, τ1)dmkdnqdmtdnτ.

Then, it follows from (3) in Section 8.2 that 〈f, f〉 ≥ 0. The elements

f, g ∈ N we consider to be equivalent (f ∼ g), if 〈f − g, f − g〉 =

0. The set of equivalence classes will be denoted by D and the

equivalence class of the element f ∈ N will be denoted by Ψf . In

other words, the set D consists of symbols Ψf , where f ∈ N , and two

symbols Ψf ,Ψg specify the same element of the set D if f ∼ g. The

inner product of the elements Ψf ,Ψg ∈ D is defined by the formula

〈Ψf ,Ψg〉 = 〈f, g〉. An element of the set D can be represented in

different ways in the form Ψf , however the inner product in D does

not depend on the choice of the representative because from the

relation f ∼ f ′, g ∼ g′, it follows that 〈f, g〉 = 〈f ′, g′〉. (Similar

considerations can be applied to other operations in D.) A linear

combination of elements in D is defined by the formula

λΨf + µΨg = Ψλf+µg,

where the linear combination λf +µg of the sequences f, g is defined

in the usual way. Hence, the set D can be considered a pre-Hilbert

space.

Page 184: Mathematical Foundations of Quantum Field Theory

March 27, 2020 8:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch08 page 137

Translation-Invariant Hamiltonians 137

In the space D, we can define the operator generalized functions

a(f, ε, t) =∫f(k)a(k, ε, t)dk, assuming that a(f, ε, t)Ψg = Ψφg,

where φ(k, σ, τ) = f(k)δσ,εδ(t − τ), and we can define the family

of operators Vτ , Wα, satisfying VτΨg = ΨVτg, WαΨg = ΨWαg.

Let us now define the Hilbert space H as the completion of the

pre-Hilbert space D.

The operators Vτ , Wα map the set D into itself and preserve inner

products. Therefore, they can be extended by continuity to unitary

operators on the spaceH. In this way, we obtain a one-parameter and

a three-parameter group of unitary operators on H; the generators

of these groups will be denoted by H and P (in other words, Vτ =

exp(−iHτ), Wα = exp(−iαP)). The symbol Φ will denote the vector

Ψθ, where θ is a function sequence with f0 = 1, fn = 0, for n > 0.

The operator generalized functions a(f, ε, t) are defined on the dense

subset D ⊂ H.

Hence, starting with the Wightman functions, we have con-

structed the objects that were described in the reconstruction theo-

rem. It is easy to check that they have all the necessary properties.

The only point we will consider in detail is the proof that the vector

Φ is the ground state of the Hamiltonian H. We derive this fact from

the following lemma.

Lemma 8.1. The number ω does not belong to the spectrum of the

operator H if and only if for all Wightman functions we have∫exp(−iωτ)wn(k1, ε1, t1, . . . ,ki, εi, ti,ki+1, εi+1, ti+1 + τ,

. . . ,kn, εn, tn + τ)dτ = 0. (8.19)

Then, by property (4) of Wightman functions and the lemma, it

follows that the operator H is non-negative. Taking into account that

HΦ = 0, we see that Φ is the ground state.

To prove (2), we first note that the functions wn are Wightman

functions of the operator H with respect to the operator generalized

function a(k, ε, t) (in the sense of the definition in Section 7.1).

Therefore, it follows from property (4) (Section 7.1) that for every

ω that does not belong to the spectrum of the operator H, we

have (8.19). To prove the inverse statement, it is sufficient to check

Page 185: Mathematical Foundations of Quantum Field Theory

March 27, 2020 8:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch08 page 138

138 Mathematical Foundations of Quantum Field Theory

that for a smooth finite function χ(ω) that does not vanish only for

ω satisfying (8.19), we have∫χ(t) exp(iHt)dt = 0, where χ(t) =∫

exp(−iωt)χ(w)dω (see Appendix A.5). It is easy to check that∫χ(t)〈exp(iHt)Ψ1,Ψ2〉dt = 0,

if Ψi = a(f(i)1 , ε

(i)1 , t

(i)1 ) . . . a(f

(i)ni , ε

(i)ni , t

(i)ni )Φ; to prove this, we should

express∫χ(t) 〈exp(iHt)Ψ1,Ψ2〉 dt =

∫χ(ω)〈exp(i(H − ω)t)Ψ1,Ψ2〉dωdt

in terms of Wightman functions. By using the cyclicity of the

vector Φ, we can see that the relation we have proven implies∫χ(t) exp(iHt)dt = 0.

This finishes the proof of the reconstruction theorem.

It follows from the lemma that every point of the spectrum of

the operator H belongs to the spectrum of the translation-invariant

Hamiltonian H in the sense of Section 8.1. Analogously, one can

prove the following statement: if the point (k, E) belongs to the

joint spectrum of a family of commuting operators (P, H), then the

Hamiltonian H has an energy level E with momentum k.

In conclusion, we will check that under certain conditions, the

space H, the operators H, P, the operator generalized functions

a(k, ε, t), and the vector Φ, constructed as in the reconstruction theo-

rem, can be considered as an operator realization of the translation-

invariant Hamiltonian H in the sense of Section 8.1. Namely, we

will show that this statement is correct if in formula (8.2), that

specifies the Hamiltonian H, the function Λ1,1(k) is smooth, all

derivatives of this function do not grow faster than a polynomial,

and the remaining functions Λm,n belong to the space S (as noted

in Section 8.1, equations (8.4) and (8.5) have precise meaning in this

case). To give the proof, it is sufficient to check that the operator

generalized functions a(k, ε, t) in the reconstruction theorem satisfy

the Heisenberg equations (8.4) and (8.5) that correspond formally

to the Hamiltonian H (all other conditions in the definition of an

operator realization follow from the reconstruction theorem). It is

Page 186: Mathematical Foundations of Quantum Field Theory

March 27, 2020 8:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch08 page 139

Translation-Invariant Hamiltonians 139

easy to give this proof, using the remark that the Wightman functions

of the Hamiltonian H satisfy

1

i

∂t1wr(k1, ε1, t1, . . . ,kr, εr, tr)

= δε11

∑m,n

n

∫Λm,n(p1, . . . ,pm|q1, . . . ,qn−1,k1)δ(p1 + · · ·+ pm

−q1 − · · · − qn−1 − k1)wr+m+n−2(p1, 1, t1, . . . ,pm, 1, t1,q1,

−1, t1, . . . ,qn−1,−1, t1,k2, ε2, t2, . . . ,kr, εr, tr)dmpdn−1q

−δε1−1

∑m,n

m

∫Λm,n(k1,p1, . . . ,pm−1|q1, . . . ,qn)

× δ(k1 + p1 + · · ·+ pm−1 − q1 − · · · − qn)wr+m+n−2(p1, 1, t1,

. . . ,pm−1, 1, t1,q1,−1, t1, . . . ,qn,−1, t1,k2, ε2, t2,

. . . ,kr, εr, tr)dm−1pdnq (8.20)

and satisfy similar conditions for derivatives with respect to other

time variables. (Equation (8.20) can be obtained if we use the

equations for the functions wΩr that follow from the considerations

in Section 7.4 and take the limit Ω → ∞ in these equations. The

conditions we have imposed on the functions Λm,n imply that we

can take this limit because we have assumed that the functions wΩr

tend to wr in the sense of generalized functions.) Using (8.20), we

can derive the Heisenberg equations (8.4) and (8.5); it is sufficient to

express in terms of Wightman functions the quantity⟨d

dta(f, t)Ψα,Ψβ

⟩,

where α, β ∈ N , using equations similar to (8.20).

It follows from the above statement that the Wightman functions

for a translation-invariant Hamiltonian defined in Section 8.2 are

Wightman functions of the operator realizations of this Hamiltonian

in the sense of Section 8.1. A similar statement can be proven

for Green functions. In the following, we will use the definition in

Section 8.1, but use the terminology from Section 8.2.

Page 187: Mathematical Foundations of Quantum Field Theory

March 27, 2020 8:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch08 page 140

140 Mathematical Foundations of Quantum Field Theory

Let us define Schwinger functions that, together with Wightman

and Green functions, play an important role in quantum field theory.

Let us consider the operators exp(iHτ), where τ is a complex

number. It follows from the non-negativity of the operator H that

these operators are bounded for τ in the upper half plane (Im τ ≥ 0).

Therefore, it is natural to assume that under the condition Im τ ≥ 0,

the operators exp(iHτ) transform the set D into itself; then, in

particular, the operator exp(−Hσ), where σ ≥ 0 has this property.

Under this assumption, we can define Schwinger functions using the

formula

Sn(k1, ε1, t1, . . . ,kn, εn, tn) = 〈a(k1, ε1, 0) exp(−H(t1 − t2))

× a(k2, ε2, 0) . . . exp(−H(tn−1 − tn))

× a(kn, εn, 0)Φ,Φ〉

(we assume that t1 ≥ · · · ≥ tn). Let us note that Schwinger functions

can also be defined without any additional assumptions. Indeed by

the property of time-translation invariance, the Wightman function

wn can be written in the following form:

wn(k1, ε1, t1, . . . ,kn, εn, tn)

= vn(k1, ε1, . . . ,kn, εn, t2 − t1, . . . , tn − tn−1).

The functions vn(k1, ε1, . . . ,kn, εn, τ1, . . . , τn−1) can be analytically

continued with respect to the variables τ1, . . . , τn−1 in the domain

Imτ1 ≥ 0, . . . , Imτn−1 ≥ 0 (to prove this fact, we should represent vnin the form

vn (k1, ε1, . . . ,kn, εn, τ1, . . . , τn−1)

=

∫exp(i

∑j

ωjτj)vn(k1, . . . , εn, ω1, . . . , ωn−1)dn−1ω

and note that by the spectrum condition the support of the function

vn is contained in the set ω1 ≥ 0, . . . , ωn−1 ≥ 0). The analytic

continuation of the function vn will be denoted by the same symbol.

Page 188: Mathematical Foundations of Quantum Field Theory

March 27, 2020 8:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch08 page 141

Translation-Invariant Hamiltonians 141

Schwinger functions can then be defined by the formula

Sn(k1, ε1, t1, . . . ,kn, εn, tn)

= vn(k1, ε1, . . . ,kn, εn, i(t1 − t2), . . . , i(tn−1 − tn))

(where the definition assumes t1 ≥ · · · ≥ tn).

It is easy to formulate properties of Schwinger functions that are

analogous to the properties of Wightman functions and to prove

the analog of the reconstruction theorem for Schwinger functions;

see Osterwalder and Schrader (1973).

8.3 Interactions of the form V (φ)

In this section, we will consider an important class of translation-

invariant Hamiltonians. The Hamiltonians of this class can be

obtained by the quantization of classical systems with an infinite

number of degrees of freedom; they can be written in the form (8.2),

which allows us to apply the results of Section 8.1.

In the present section, it will be convenient not to use the

assumption that ~ = 1, made in the rest of the book.

Let us recall that by quantizing a classical mechanical system with

the Hamiltonian

H(p, q) =n∑k=1

p2k

2+ U(q1, . . . , qn), (8.21)

where pk are generalized momenta and qk are generalized coordinates,

we obtain a quantum mechanical system described by the Hamilto-

nian

H =n∑k=1

1

2p2k + U(q1, . . . , qn), (8.22)

where pk, qk are self-adjoint operators satisfying the canonical com-

mutation relations (CCR)

[pk, pl] = [qk, ql] = 0; [pk, ql] =~iδkl

Page 189: Mathematical Foundations of Quantum Field Theory

March 27, 2020 8:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch08 page 142

142 Mathematical Foundations of Quantum Field Theory

(the operators pk, qk can be realized in the space of square-integrable

functions ψ(q1, . . . , qn) and defined by the formulas pkψ = ~i∂∂qk

ψ;

qkψ = qkψ).

The Heisenberg operators pk(t) = exp(iHt)pk exp(−iHt), qk(t) =

exp(iHt)qk exp(−iHt) satisfy the equations

dpk(t)

dt= −∂U

∂qi(q1(t), . . . , qn(t));

dqk(t)

dt= pk(t).

Let us now consider the analog of a classical system with the

Hamiltonian (8.21) in the case of an infinite number of degrees of

freedom. Namely, we assume that the classical system is described

by the Hamiltonian functional

H(π, φ) =1

2

∫π2(x)dx + V (φ),

V (φ) =∑n

∫Vn(x1, . . . ,xn)φ(x1) . . . φ(xn)dx1 . . . dxn,

where π(x) are the generalized momentum variables and φ(x) are the

generalized coordinates. For definiteness, we assume that x runs over

three-dimensional Euclidean space. We consider only translation-

invariant functionals (i.e. we assume that the function Vn(x1, . . . ,xn)

has the form vn(x1− xn, . . . ,xn−1− xn)). It is natural to conjecture

that by quantizing such a system we will obtain a quantum system

described by the Hamiltonian

H =1

2

∫π2(x)dx +

∑n

∫Vn(x1, . . . ,xn)φ(x1) . . . φ(xn)dx1 . . . dxn,

(8.23)

where π(x), φ(x) are Hermitian operators (more precisely, operator

generalized functions) that obey the commutational relations

[π(x), π(x′)] = [φ(x), φ(x′)] = 0,

[π(x), φ(x′)] =~iδ(x− x′). (8.24)

Page 190: Mathematical Foundations of Quantum Field Theory

March 27, 2020 8:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch08 page 143

Translation-Invariant Hamiltonians 143

However, in trying to correctly define this quantum system, we

encounter some difficulties that we have already encountered in con-

sidering a Hamiltonian of the form (8.2). Namely, there exist many

essentially different systems of operators satisfying (8.24). For simple

constructions of the operators π(x), φ(x), the expression (8.23) does

not specify a well-defined operator. These difficulties can be overcome

by means of the approach in Section 8.1. Here, we will consider only

the method of solving the Heisenberg equations formally written

using the Hamiltonian (8.23).

Namely, we will consider the Hamiltonian H of the form (8.23)

as a formal expression. We define the operator realization of the

Hamiltonian H as a Hilbert space H with the energy operator H,

the momentum operator P, the vector φ, and the operator functions

φ(x, t), that are generalized functions with respect to the variable x

and conventional functions with respect to the variable t, satisfying

the following conditions:

(1)

∂2φ(x, t)

∂t2= −

∑n

n

∫Vn(x,x1, . . . ,xn−1)

× φ(x1, t) . . . φ(xn−1, t)dx1 . . . dxn−1.

(8.25)

(2)

exp

(i

~τH

)φ(x, t) exp

(−i~τH

)= φ(x, t+ τ),

exp

(−i~ap

)φ(x, t) exp

(i

~ap

)= φ(x+ a, t).

(3) The operators φ(f, t) =∫f(x)φ(x, t)dx and π(f, t) = d

dtφ(f, t) =∫f(x) ∂∂tφ(x, t)dx, where f ∈ S(E3), are defined on a dense

subset D of the space H and transform the subset into itself;

if the function f is real, then these operators are Hermitian. The

expressions 〈φ(f, t)Ψ1,Ψ2〉 and 〈π(f, t)Ψ1,Ψ2〉 should depend

Page 191: Mathematical Foundations of Quantum Field Theory

March 27, 2020 8:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch08 page 144

144 Mathematical Foundations of Quantum Field Theory

continuously on the functions f ∈ S(E3) in the topology of the

space S(E3) for all Ψ1,Ψ2 ∈ D. We assume further that for all

t, the conditions

[π(f, t), π(f ′, t)] = [φ(f, t), φ(f ′, t)] = 0,

[π(f, t), φ(f ′, t)] =~i

∫f(x)f ′(x)dx

are satisfied.

(4) The operators H, P1, P2, P3 commute. The vector Φ is the ground

state of the energy operator H and satisfies the conditions HΦ =

0, PΦ = 0.

(5) The vector Φ is a cyclic vector of the family of operators φ(f, t).

In some simple cases, one can define the precise meaning for

equation (8.25) by means of the operator analog of the kernel theorem

(as in Section 8.1).

For Hamiltonians of the form

H0 =1

2

∫π2(x)dx +

1

2

∫v(x− y)φ(x)φ(y)dxdy (8.26)

(free Hamiltonians), it is easy to construct an operator realization.

Let us assume that the function v(k) =∫

exp(−ikx)v(x)dx is

positive almost everywhere (if this condition is not satisfied, then

an operator realization of the Hamiltonian H0 does not exist). Let

H, H, P,Φ, φ(x, t) be an operator realization of the Hamiltonian H0.

Let us construct the operator generalized functions

a(k, ε, t) = ~−1/2

(1√2

√ω(k)φ(εk, t) +

iε√ω(k)

π(εk, t)

),

where ω(k) =√v(k), φ(k, t) = (2π)−3/2

∫exp(−ikx)φ(x, t)dx, and

π(k, t) = ∂∂t φ(k, t). It is easy to check that these operator generalized

functions, together with the operators H, P and the vector Φ, specify

an operator realization of the Hamiltonian ~∫ω(k)a+(k)a(k)dk

in the sense of Section 8.1. This statement prompts the following

construction of an operator realization of the Hamiltonian (8.26).

Page 192: Mathematical Foundations of Quantum Field Theory

March 27, 2020 8:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch08 page 145

Translation-Invariant Hamiltonians 145

As the space H, we should take the Fock space F (L2(E3)), and the

energy operator H and the momentum operator P should be defined

by the formulas

H = ~∫ω(k)a+(k)a(k)dk,

P = ~∫

ka+(k)a(k)dk,

and the operator generalized functions φ(x, t) should be defined by

the relation

φ(x, t) = (2π)−3/2~1/2

∫(a+(k) exp(iω(k)t− ikx)

+ a(k) exp(−iω(k)t+ ikx))dk√2ω(k)

.

The ground state Φ coincides with Fock vacuum θ. It is easy to

check that the objects we have constructed satisfy the conditions for

an operator realization and that any other operator realization is

unitarily equivalent to the realization we have described.

Let us now consider an arbitrary Hamiltonian H of the

form (8.23). Let us express it in terms of the symbols a+(k), a(k)

satisfying CCR, assuming that

φ(x) = (2π)−3/2~1/2

∫(a+(k) exp(−ikx) + a(k) exp(ikx))

dk√2ω(k)

,

π(x) = (2π)−3/2~1/2

∫i√ω(k)√2

(a+(k) exp(−ikx)

−a(k) exp(ikx))dk, (8.27)

where ω(k) is an almost everywhere positive function. This expres-

sion can be written in normal form by means of CCR; we obtain

a quadratic expression plus a (possibly infinite) constant. We will

discard this constant and as a result we will obtain a formal

expression for H of the form (8.2).

It is easy to see that the problem of constructing an operator

realization of the Hamiltonian H is equivalent to the same problem

Page 193: Mathematical Foundations of Quantum Field Theory

March 27, 2020 8:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch08 page 146

146 Mathematical Foundations of Quantum Field Theory

for H. For example, if H, H, P,Φ, a(k, ε, t) is an operator realization

of the Hamiltonian H, then the operator realization of the Hamil-

tonian H can be obtained if we take the same Hilbert space H, the

same operators H, P, and vector Φ, and the operator generalized

functions φ(x, t) can be written in the form

φ(x, t) = (2π)−3/2~1/2

∫(a(k, 1, t) exp(−ikx)

+ a(k,−1, t) exp(ikx))dk√2ω(k)

.

This remark allows us to transfer to Hamiltonians of the form (8.23)

everything we know for Hamiltonians of the form (8.2).

Page 194: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 147

Chapter 9

The Scattering Matrix forTranslation-Invariant Hamiltonians

9.1 The scattering matrix for translation-invariant

Hamiltonians in Fock space

In this chapter, we review the basic facts of scattering theory for

translation-invariant Hamiltonians. Some of these facts are proved,

however, the proofs are not fully rigorous (Sections 9.2, 9.5), while

other facts are conveyed almost without proof (Sections 9.1, 9.3 and

9.4). Most of the results in this chapter will be proved later on the

basis of axiomatic scattering theory (see Chapter 11).

To construct a scattering matrix for a translation-invariant

Hamiltonian, one cannot directly use the general construction of

formal scattering theory (Section 5.1). The first obstacle for applying

this construction is the fact that a translation-invariant Hamiltonian

can define a self-adjoint operator on Fock space only in the case

when vacuum polarization is absent (Section 8.1). However, even

translation-invariant Hamiltonians that specify an operator on Fock

space still have problems. These problems are related to the fact

that in the case under consideration, a natural representation of the

Hamiltonian H as a sum of “free” Hamiltonian H0 and “interaction”

V usually does not exist.

It is reasonable to take the Hamiltonian H0 to be of the form∫ε(k)a+(k)a(k)dk because such a Hamiltonian describes a system

of non-interacting identical particles. We assume, therefore, that

the translation-invariant Hamiltonian H specifying an operator on

147

Page 195: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 148

148 Mathematical Foundations of Quantum Field Theory

Fock space F (L2(E3)) is represented in the form H = H0 + V ,

where

H0 =

∫ε(k)a+(k)a(k)dk, (9.1)

V =∑m,n≥1

∫Vm,n(k1, . . . ,km|p1, . . . ,pn)a+(k1)

. . . a+(km)a(p1) . . . a(pn)dmkdnp (9.2)

(here,

Vm,n(k1, . . . ,km|p1, . . . ,pn) = vm,n(k1, . . . ,km|p1, . . . ,pn)

× δ(k1 + · · ·+ km − p1 − · · · − pn)).

(9.3)

However, such a representation of the Hamiltonian H has physical

meaning only under the condition vm,1 ≡ 0. Otherwise, the choice of

free and interaction parts can be performed in different ways and is

dictated only by convenience.

One can prove the following statement: if Vm,1 6≡ 0 (i.e. the

“interaction” contains summands with one annihilation operator),

then the Møller matrices S± cannot be defined by means of the

definitions of formal scattering theory (5.1) and (5.2). Indeed, if there

exists a limit entering the relations (5.1) and (5.2), then for every

vector x ∈ F , we have

limt1,t2→−∞;(t1,t2→+∞)

∥∥∥∥∫ t2

t1

exp(iHt)V exp(−iH0t)xdt

∥∥∥∥ = 0.

(We can check this using the relation

exp(itH)V exp(−itH0)x =1

i

dξ(t)

dt,

where ξ(t) = exp(iHt) exp(iH0t)x and hence∫ t2

t1

exp(iHt)V exp(−iH0t)xdt =1

i(ξ(t2)− ξ(t1)).

)

Page 196: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 149

The Scattering Matrix for Translation-Invariant Hamiltonians 149

However, if x =∫f(k)a+(k)dkθ, then∥∥∥∥dξdt

∥∥∥∥ = ‖ exp(iHt)V exp(−iH0t)x‖ = ‖V exp(−iH0t)x‖

does not depend on t. This is clear because in this case

exp(iH0t)x =

∫exp(iε(k)t)f(k)a+(k)dkθ,

V exp(−iH0t)x =∑n

∫fn(k1, . . . ,kn|t)a+(k1) . . . a+(kn)dnkθ,

where

fn(k1, . . . ,kn|t) = vn,1(k1, . . . ,kn|k1 + · · ·+ kn)

∫(k1 + · · ·+ kn)

× exp(−itε(k1 + · · ·+ kn)).

Noting that

‖V exp(−iH0t)x‖ =∑n

n!

∫|fn(k1, . . . ,kn|t)|2dnk,

we see that ‖dξdt ‖ does not depend on t. Similar considerations show

that ‖d2ξdt2‖ = ‖ ddt(exp(iHt)×V exp(−iH0t)x)‖ = ‖ exp(−iHt)(HV −

V H0) × exp(−iH0t)x‖ = ‖(HV − V H0) exp(−iH0t)x‖ does not

depend on t. Now, to finish the proof, we should apply the following

lemma to the vector η(t) = dξdt : if ‖η(t)‖ does not depend on t

and ‖dηdt ‖ is bounded above, then∫ t2t1η(t)dt cannot tend to zero as

t1, t2 →∞.

This mathematical statement — the fact that it is impossible

to give the definition of a Møller matrix in the same way as in the

theory of potential scattering — has a clear physical background. The

problem is that the states a+(k)θ (one particle states) are eigenstates

of the Hamiltonian H0 but are not eigenstates of the full Hamiltonian

H. In Section 5.3, we introduced the notion of a particle (single-

particle state) for the Hamiltonian H as a generalized vector function

Φ(k) satisfying the conditions (5.20)–(5.22).

Page 197: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 150

150 Mathematical Foundations of Quantum Field Theory

Using this notion, we can reformulate the above statement in

the following way: the generalized vector function a+(k)θ is a

single-particle state of the Hamiltonian H0 (bare single-particle

state), however, it is not a single-particle state of the Hamiltonian

H (“dressed” single-particle state). The fact that the notions of bare

and dressed particles do not coincide explains the necessity to modify

the definition of a scattering matrix.

In the case when vm,0 ≡ vm,1 ≡ 0, bare particles coincide

with dressed ones, and therefore, it is not necessary to modify the

definition of a scattering matrix.

Let us sketch how to perform the necessary modification of

the definition of a scattering matrix for a translation-invariant

Hamiltonian that specifies an operator on Fock space.

We will give three definitions of Møller matrices S± that can

be used in the situation when dressed particles coincide with bare

particles (i.e. the Hamiltonian has the form (9.1) and vm,1 ≡ 0). We

will show how to modify these definitions for the case at hand (recall

that knowing the Møller matrices S±, we can define the scattering

matrix by the formula S = S∗+S−).

The first definition is the same definition that was used in the

formal scattering theory (see (5.1) and (5.2)). One can also define

Møller matrices S± as strong limits of the adiabatic Møller matrices

Sα± with α → 0. In Section 5.1, we described the conditions that

guaranteed that the second definition is equivalent to the first one

(see (5.4) and (5.5)). Finally, the third definition is based on the

consideration of the in- and out-operators

a inout

(k) = slimt→∓∞

exp(iε(k)t)a(k, t),

a+in

out

(k) = slimt→∓∞

exp(−iε(k)t)a+(k, t). (9.4)

If we know the in- and out-operators, then the Møller matrices can

be defined by the relations

S−ain(k) = a(k)S−, S−θ = θ,

S+aout(k) = a(k)S+, S+θ = θ (9.5)

Page 198: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 151

The Scattering Matrix for Translation-Invariant Hamiltonians 151

(the equivalence of this definition with the first one was discussed in

Section 5.3).

Let us start by generalizing the third definition. The limit in (9.4)

does not exist for vm,1 6≡ 0; this is clear from the fact that

the expression

exp(iHt) exp(−iH0t)

∫f(k)a+(k)θdk

=

∫f(k) exp(−iε(k)t)a+(k, t)θdk

does not have a limit as t→ ±∞ (as was proven already).

However, this fact does not mean that the expression

exp(iω(k)t)a(k, t) does not have a weak limit. Therefore, the in- and

out-operators (operator generalized functions) ain(k), a+in(k), aout(k),

a+out(k) will be defined by means of the following relations:

ain(k) = wlimt→−∞

Λ(k) exp(iω(k)t)a(k, t),

a+in(k) = wlim

t→−∞Λ(k) exp(−iω(k)t)a+(k, t),

aout(k) = wlimt→+∞

Λ(k) exp(iω(k)t)a(k, t),

a+out(k) = wlim

t→+∞Λ(k) exp(−iω(k)t)a+(k, t),

where a positive function ω(k) should be found from the condition of

the existence of the limit and the function Λ(k) from the condition

that the operators ain(k), a+in(k) and aout(k), a+

out(k) satisfy CR:

[a inout

(k), a inout

(k′)]∓ = [a+in

out

(k), a+in

out

(k′)] = 0,

[a inout

(k), a+in

out

(k′)]∓ = δ(k− k′).

It is easy to check (see Section 9.2) that the vector generalized

functions a+in(k)θ and a+

out(k)θ are single-particle states of the

Hamiltonian H and therefore conclude that the functions ω(k) has

meaning of the energy of a single-particle state:

Ha+in

out

(k)θ = ω(k)a+in

out

(k)θ.

Page 199: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 152

152 Mathematical Foundations of Quantum Field Theory

Møller matrices are defined as earlier in terms of in- and out-

operators by (9.5).

The second definition can be generalized in the following way.

The Møller matrices S± are defined as unitary operators that can be

represented in the form

S+ = slimα→0

Sα+U∗α, (9.6)

S− = slimα→0

Sα−Uα, (9.7)

where Uα is an operator of the form

exp

(i

α

∫r(k)a+(k)a(k)dk

), (9.8)

and the function r(k) is chosen from the condition of the existence

of the limits (9.6) and (9.7).

If we want to define directly the scattering matrix, we can use the

relation

S = slimα→0

UαSαUα, (9.9)

where Sα is the adiabatic S matrix, the operator Uα has the

form (9.8) and the function r(k) is chosen from the condition of

existence of the limit (9.9).

Finally, we give the following modification for the first definition:

Møller matrices S± are operators of the form

S± = slimt→±

exp(iHt)T exp(−iHast), (9.10)

where T is an operator satisfying the conditions

Tθ = θ,

Ta+(k)θ = Φ(k)

and having the form T = N(expB), where

B =∑n

∫bn(k1, . . . ,kn)δ

(∑i

ki − k

)a+(k1) . . .

a+(kn)a(k)dkdk1 . . . dkn,

Page 200: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 153

The Scattering Matrix for Translation-Invariant Hamiltonians 153

(Φ(k) denotes a singe-particle state of the Hamiltonian H,Has =∫ω(k)a+(k)a(k)dk, where ω(k) is the energy of the single-particle

state: HΦ(k) = ω(k)Φ(k)).1

In the relation (9.10), one can replace the operator T described

above by other operators. All operators D such that

S± = slimt→±∞

exp(itH)D exp(−itHas)

will be called dressing operators (this term is related to the fact

that they transform the bare single-particle state a+(k)θ into the

dressed single-particle state Φ(k)). One can construct a broad class

of dressing operators (see Sections 9.4 and 9.5).

All three definitions of Møller matrices are equivalent (under

certain conditions). They can be generalized further to translation-

invariant Hamiltonians generating the polarization of vacuum.

Naturally, there are additional complications related to the fact

that the Hamiltonian does not specify an operator in Fock

space.

For the definition by means of in- and out-operators, these

complications can be overcome by considering an operator realization

of the Hamiltonian H. Other definitions should be modified by

considering the volume cutoff Ω of the Hamiltonian H and taking the

limit Ω → ∞ (in this case one should define the scattering matrix

directly because Møller matrices cannot be defined in this case).

In more detail, we will study various definitions of the scattering

matrix for a translation-invariant Hamiltonian in other sections of

this chapter and in Chapter 11.

In Section 5.3, we noted that the definition of scattering matrix

should be modified even for the simplest Hamiltonians if there exist

bound states. Similar modifications are necessary in the situation at

hand if there exist bound states. The definition of scattering matrix

that can be used in the case when bound states exist is given in

Section 11.1.

1The definition of Møller matrices by means of (9.10) was suggested byI. Ya. Arefieva.

Page 201: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 154

154 Mathematical Foundations of Quantum Field Theory

9.2 The definition of scattering matrix by means of

operator realization of a translation-invariant

Hamiltonian

Let H denote a translation-invariant Hamiltonian and let us consider

its operator realization (H, H, P, a(k, ε, t),Φ) (see Section 8.1).

Let us define the in-operators ain(k, ε, t) and out-operators

aout(k, ε, t) as limits

a inout

(k, t) = a inout

(k,−1, t)

= wlimτ→∓∞

Λ∓(k) exp(iω(k)τ)a(k,−1, t+ τ),

a+in

out

(k, t) = a inout

(k, 1, t)

= wlimτ→∓∞

Λ∓(k) exp(−iω(k)τ)a(k, 1, t+ τ),

(9.11)

where ω(k) is found from the condition of the existence of limits and

Λ∓(k) is found from the condition that the operators a+in(k, t) and

ain(k, t) (and, correspondingly, a+out(k, t), aout(k, t)) satisfy CR for

fixed t. (Here, ω(k) is an almost everywhere positive function and the

limit is understood as a weak limit of operator generalized functions,

ε = ±1. Introducing the notation Λ∓(k,−1) = Λ∓(k),Λ∓(k, 1) =

Λ∓(k), we can say that for every function f ∈ S, we have

a inout

(f, ε, t) =

∫f(k)a in

out(k, ε, t)dk

= wlimτ→∓∞

∫f(k)Λ∓(k, ε) exp(−iεω(k)τ)a(k, ε, t+ τ)dk

in the sense of weak limits of operators.)

This definition of in- and out-operators differs from the definition

accepted in the theory of potential scattering (Section 5.3) by

replacing strong limits with weak limits. The factor Λ∓(k) is related

to this modification: strong limits preserve CR, but weak limits

do not.

The question of the existence of in- and out-operators (i.e. the

question of the existence of the functions ω(k) and Λ∓(k) such that

Page 202: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 155

The Scattering Matrix for Translation-Invariant Hamiltonians 155

the limit in (9.11) exists and satisfies CR) will be discussed in Chap-

ters 10 and 11. To denote the in- and out-operators simultaneously,

we will use the notation aex(k, ε, t).

Let us establish a few simple properties of in- and out-operators:

(1) aex(k, ε, t) = exp(iHt)aex(k, ε, 0) exp(−iHt) =

exp(iεω(k)t)aex(k, ε, 0);

(2) exp(iPα)aex(k, ε, t) exp(−iPα) = exp(iεkα)aex(k, ε, t);

(3) aex(k, t)Φ = 0;

(4) the generalized vector function Φ∓(k) = a+ex(k, 0)Φ is

δ-normalized and is an eigenfunction of the operators H and P:

HΦ∓(k) = ω(k)Φ∓(k),

PΦ∓(k) = kΦ∓(k)

(the functions Φ∓(k) satisfying this condition describe a single-

particle state; see Section 5.3).

We first note that (2) follows from the relation

exp(iPα)a(k, ε, t) exp(iPα) = exp(iεkα)a(k, ε, t).

Then,

aex(k, ε, t) = wlimτ→∓∞

Λ∓(k, ε) exp(iεω(k)τ)a(k, ε, t+ τ)

= wlimτ→∓∞

exp(iHt)Λ∓(k, ε) exp(iεω(k)τ)

× a(k, ε, τ) exp(−iHt)

= exp(iHt)( wlimτ→∓∞

Λ∓(k, ε) exp(iεω(k)τ)a(k, ε, τ))

× exp(−iHt) = exp(iHt)aex(k, ε, 0) exp(−iHt).

From the other side, substituting ρ instead of t + τ in (9.11), we

obtain

aex(k, ε, t) = wlimρ→∓∞

Λ∓(k, ε) exp(iεω(k)(ρ− t))a(k, ε, ρ)

= exp(−iεω(k)t)aex(k, ε, 0).

Page 203: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 156

156 Mathematical Foundations of Quantum Field Theory

It follows from (1) that

exp(iHt)aex(k, ε)Φ = exp(iHt)aex(k, ε) exp(−iHt)Φ

= aex(k, ε, t)Φ = exp(iεω(k)t)Φ,

hence,

Haex(k, ε)Φ = εω(k)aex(k, ε)Φ. (9.12)

From (9.12), it follows that aex(k,−1)Φ = 0 (if this condition is

not satisfied for k in the set K having non-zero measure, then the

generalized vector function aex(k,−1)Φ is a generalized eigenfunction

of the operator H, hence the number −ω(k), where k ∈ K, belongs

to the spectrum of the operator H; this is impossible because the

operator H is positive).

Applying CR and (3), we see that the generalized function

Φ∓(k) = aex(k, 1)Φ is δ-normalized (〈Φ∓(k) , Φ∓(k′)〉 =

〈aex(k′,−1)aex(k, 1)Φ,Φ〉 = 〈[aex(k′), a+ex(k)]∓Φ,Φ〉 = δ(k − k′)).

Formula (9.12) implies that Φ∓(k) is a generalized eigenfunction

of the operator H. In order to prove that Φ∓(k) is a generalized

eigenfunction of the operator P, we should recall that by (2), we have

exp(iPα)aex(k, ε, t) exp(−iPα) = exp(iεkα)aex(k, ε, t),

which implies that

exp(iPα)Φ∓(k)=exp(iPα)a+ex(k, 0) exp(−iPα)Φ=exp(εkα)Φ∓(k)

and therefore PΦ∓(k) = kΦ∓(k).

Remark 9.1. We have assumed that the function ω(k) in the

definition of in- and out-operators is almost everywhere positive. One

can replace this condition by the condition that the operator H has

a unique ground state. Then, in the case of CCR, we can modify the

above considerations to check that the function ω(k) is automatically

almost everywhere positive. In the case of CAR, one should introduce

new operators a+(k, t), a(k, t), a+ex(k, t), aex(k, t) also satisfying CAR,

Page 204: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 157

The Scattering Matrix for Translation-Invariant Hamiltonians 157

using the formulas

a(k, t) = θ(ω(k))a(k, t) + θ(−ω(k))a+(−k, t),

aex(k, t) = θ(ω(k))aex(k, t) + θ(−ω(k))a+ex(−k, t).

Then, it is clear that

aex(k, t) = wlimτ→∓∞

Λ′∓(k) exp(i|ω(k)|τ)a(k, t+ τ),

where the function |ω(k)| is almost everywhere positive.

Let us introduce the space of asymptotic states Has as the space

F (L2(E3)) of Fock representation of CR. The symbol b(k, ε) denotes

the operator generalized functions in this space that satisfy the

conditions

[b(k, ε), b(k′, ε′)]∓ = Aεε′δ(k− k′),

b(k,−1)θ = 0, b+(k,−1) = b(k, 1).

Møller matrices S− and S+ are defined as isometric operators

mapping the space Has into the space H and satisfying the equations

ain(k, ε)S− = S−b(k, ε), S−θ = Φ, (9.13)

aout(k, ε)S+ = S+b(k, ε), S+θ = Φ (9.14)

(it follows from the results of Section 6.1 that such operators exist

and are defined by conditions (9.13) and (9.14) uniquely).

The scattering matrix of a translation-invariant Hamiltonian H

is defined as the operator S = S∗+S−.

It is easy to verify that the scattering matrix S will be unitary if

and only if the spaces Hin = S−Has and Hout = S+Has coincide.

We will assume that H = Hin = Hout (in other words, not only is

the S-matrix unitary, but so are the Møller matrices S− and S+).

Some of the relations proven later, in particular (9.22), are correct

without this assumption.

Page 205: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 158

158 Mathematical Foundations of Quantum Field Theory

Let us define the operator Has (asymptotic Hamiltonian) and the

operator Pas in the space Has using the formulas

Has =

∫ω(k)b+(k)b(k)dk,

Pas =

∫kb+(k)b(k)dk

(as usual, b+(k) = b(k, 1), b(k) = b(k,−1)). It is easy to see that

b(k, ε, t) = exp(iHast)b(k, ε) exp(−iHast) = exp(iεω(k)t)b(k, ε),

exp(iPasα)b(k, ε, t) exp(−iPasα) = exp(iεαk)b(k, ε, t).

Using this relation and the properties (1) and (2) of the operators

aex, we can show that

HS± = S±Has , PS± = S±Pas (9.15)

(i.e. the operators S+ and S− specify unitary equivalences between

the operators H and Has , P and Pas). From (9.15), it follows that

the scattering matrix commutes with the operators Has and Pas :

SHas = HasS,

SPas = PasS.

Let us now prove that the scattering matrix defined above has the

following properties:

(1) Sθ = θ (the vacuum is stable),

(2) Sb+(k)θ = c(k)b+(k)θ,

where |c(k)| = 1 (single-particle states are stable). The second of

these statements will be proved only under certain restrictions on the

function ω(k); it is sufficient to assume that it is strongly convex.

The stability of the vacuum follows immediately from the relations

S−θ = Φ, S+θ = Φ. To prove that single-particle states are stable,

we note that

HasSb+(k)θ = SHasb

+(k)θ = ω(k)Sb+(k)θ,

PasSb+(k)θ = SPasb

+(k)θ = kSb+(k)θ.

Page 206: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 159

The Scattering Matrix for Translation-Invariant Hamiltonians 159

Hence, the generalized vector function Ψ(k) = Sb+(k)θ satisfies the

condition HasΨ(k) = ω(k)Ψ(k), PasΨ(k) = kΨ(k); it is easy to see

that this generalized vector function is δ-normalized. It follows that

Ψ(k) = c(k)b+(k)θ, where |c(k)| = 1 (see Section 5.3).

We will now consider an ambiguity in our definition of in- and out-

operators. It is not difficult to check that the functions Λ−(k) and

Λ+(k) are not specified uniquely by the requirement the operators

aex(k, t) obey CR. Namely, we can replace the functions Λ−(k) and

Λ+(k) by functions Λ′−(k) and Λ′+(k) that have the same absolute

value as the functions Λ−(k) and Λ+(k) and obtain the new in- and

out-operators a′in(k, t) and a′out(k, t):

a′in(k, t) = exp(iφ−(k))ain(k, t),

a′out(k, t) = exp(iφ+(k))aout(k, t) (9.16)

(here, exp(iφ∓(k)) = Λ′∓(k)Λ−1± (k) and φ∓(k) are real-valued

functions).

One can check that under the above assumptions, all possible in-

and out-operators can be represented as operators a′ex(k, t).

It is easy to check that the Møller matrices S′− and S′+ constructed

by means of the operators a′ex(k, t) are related to the Møller matrices

S− and S+ corresponding to the operators aex(k, t) by the formula

S′∓ = S∓U∓,

where U∓ = exp(i∫φ∓(k)b+(k)b(k)dk) (this follows from the

relation U∓b(k)U−1∓ = exp(iφ∓(k))b(k)).

Hence, the scattering matrix S′ = S′∗+S′− corresponding to the

new ex-operators is related to the old scattering matrix S = S∗+S−by the relation S′ = U−1

+ SU−.

Using this ambiguity in the definition of in- and out-operators,

one can strengthen the stability condition of single-particle states,

namely, we can require

Sb+(k)θ = b+(k)θ. (9.17)

In what follows, we always assume that the in- and out-operators are

chosen in such a way that condition (9.17) is satisfied.

Page 207: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 160

160 Mathematical Foundations of Quantum Field Theory

There is less ambiguity in the definition of in- and out-operators

if we make the following assumption: the relation between the

old and new ex-operators is given by the formulas (9.16), where

φ−(k) = φ+(k); correspondingly, in the formulas relating the new

and old Møller matrices and new and old scattering matrices, we

have U− = U+.

Let us check that the condition (9.17) is satisfied if and only

if the functions Λ−(k) and Λ+(k) are equal: Λ−(k) = Λ+(k) =

Λ(k). To check this, we consider the decomposition of gener-

alized vector functions a+(k)Φ with respect to the generalized

basis a+in(k1) . . . a+

in(kn)Φ = Φ−(k1, . . . ,kn). This decomposition has

the form

a+(k)Φ = ρ(1,k)Φ−(k) +∞∑n=2

1√n!

∫ρ(1,k1, . . . ,kn)

× δ

(k−

n∑i=1

ki

)Φ−(k1, . . . ,kn)dnk, (9.18)

where the function ρ(ε,k1, . . . ,kn) is defined by the equation

1√n!〈a(k, ε)Φ,Φ−(k1, . . . ,kn)〉 = ρ(ε,k1, . . . ,kn)δ

(εk−

∑i

ki

).

From (9.18), it follows that∫f(k) exp(−iω(k)t)a+(k, t)Φdk

=

∫f(k) exp(−iω(k)t) exp(iHt)a+(k)Φdk

=

∫f(k)ρ(1,k)Φ−(k)dk +

∑n≥2

1√n!

∫ρ(1,k1, . . . ,kn)

× δ

(k−

∑i

ki

)exp(i(ω(k1) + · · ·+ ω(kn)

− ω(k))t)Φ−(k1, . . . ,kn)dnk.

Page 208: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 161

The Scattering Matrix for Translation-Invariant Hamiltonians 161

For large t, the second summand contains a fast-oscillating factor

and therefore it weakly tends to zero as t→ ±∞. Hence, we see that

wlimt→±∞

∫f(k) exp(−iω(k)t)a+(k, t)Φdk

=

∫f(k)ρ(1,k)Φ+(k)dk. (9.19)

From the other side, from (9.11), we obtain

wlimt→∓∞

∫f(k) exp(−iω(k)t)Λ∓(k)a+(k, t)Φdk

=

∫f(k)a+

inout

(k)Φdk =

∫f(k)Φ∓(k)dk. (9.20)

Comparing Equations (9.19) and (9.20) and noting that by (9.17),

Φ−(k) = Φ+(k), we obtain

Λ−(k) = Λ+(k) = Λ(k) = (ρ(1,k))−1. (9.21)

Hence, we have proven the equality Λ−(k) = Λ+(k) and also related

these quantities with the function ρ(1,k). Using this relation and

the Kallen–Lehmann representation, we will prove that the functions

ω(k) and |Λ(k)| can be expressed in terms of the Green function

G2(p1, 1, ω1,p2,−1, ω2) = G(p1, ω1)δ(ω1 − ω2)δ(p1 − p2).

Namely, the function ω(p) specifies the location of poles of the

functions G(p, ω) (i.e. the poles of the function G(p, ω) with respect

to the variable ω, for fixed p, are located at the points ω(p) and

−ω(p)). The function |Λ(p)| is equal to Z(p)−1/2, where iZ(p) is

the residue of the function G(p, ω) at the pole ω(p).

To prove this statement, we note that we can take as a

generalized eigenbasis of the operators H and P the basis of

vectors a+in(k1) . . . a+

in(kn)Φ = Φ−(k1, . . . ,kn). Let us write down

the Kallen–Lehmann representation for G(p, w) (8.10) in this basis.

Page 209: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 162

162 Mathematical Foundations of Quantum Field Theory

We obtain

G(p, ω) =i|ρ(−1,−p)|2

ω − ω(−p) + i0∓ i|ρ(+1,p)|2

ω + ω(p)− i0

+ i∑n≥2

∫|ρ(−1,k1, . . . ,kn)|2

ω − (ω(k1) + · · ·+ ω(kn)) + i0

× δ(p + k1 + · · ·+ kn)dk1 . . . dkn

∓i∑n≥2

∫|ρ(+1,k1, . . . ,kn)|2

ω + ω(k1) + · · ·+ ω(kn)− i0

× δ(−p + k1 + · · ·+ kn)dk1 . . . dkn.

In combination with (9.21), this relation gives a proof of the

statements we need.

The generalized functions

Sm,n(p1, . . . ,pm|q1, . . . ,qn)

=⟨Sb+(q1) . . . b+(qn)θ, b+(p1) . . . b+(pm)θ

⟩(the matrix entries of the operator S in the generalized basis

b+(p1) . . . b+(pm)θ) are called the scattering amplitudes. They can

be easily expressed in terms of the generalized vector functions

a+ex(p1) . . . a+

ex(pm)θ (that are called the in- and out-states), namely,

Sm,n(p1, . . . ,pm|q1, . . . ,qn)

=⟨S−b

+(q1) . . . b+(qn)θ, S+b+(p1) . . . b+(pm)θ

⟩=⟨a+

in(q1) . . . a+in(qn)Φ, a+

out(p1) . . . a+out(pn)Φ

⟩.

Knowing the scattering amplitudes Sm,n, we can calculate the

differential collision cross-section (see Section 10.5). It is often

convenient to represent the operator S in normal form

S =∑m,n

1

m!n!

∫σm,n(p1, . . . ,pm|q1, . . . ,qn)b+(p1) . . . b(qn)dmpdnq.

The functions σm,n are closely related to the scattering amplitudes.

This can be easily checked with Wick’s theorem (see Section 6.3); we

Page 210: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 163

The Scattering Matrix for Translation-Invariant Hamiltonians 163

note here that under the condition pi 6= qj(1 ≤ i ≤ m, 1 ≤ j ≤ n),

we have

Sm,n(p1, . . . ,pm|q1, . . . ,qn) = σm,n(p1, . . . ,pm|q1, . . . ,qn).

We will now show how to express the functions σm,n in terms of the

Green functions of a translation-invariant Hamiltonian H.

We will prove the following Lehmann–Symanzik–Zimmermann

formula (LSZ):

σm,n(p1, . . . ,pm|q1, . . . ,qn)

= (i√

2π)m+nm∏i=1

Λ(pi)

n∏j=1

Λ(qj) limωi→ω(pi)

limσj→ω(qj)

m∏i=1

(ωi − ω(pi))

×n∏j=1

(σj − ω(qj))Gm+n(q1, 1, σ1, . . . ,qn, 1, σn|p1,−1,

ω1, . . . ,pm,−1, ωm), (9.22)

where Gm+n are the Green functions of the translation-invariant

Hamiltonian H in the (k, ω) representation.

We will give the proof of the LSZ formula in the CCR case (the

case of CAR differs only by signs). First, let us note that

aout(k1, ε1)T (a(k2, ε2, t2) . . . a(kn, εn, tn))

= wlimt1→+∞

a(k1, ε1, t1) exp(−iε1ω(k1)t1)Λ(k1, ε1)

× T (a(k2, ε2, t2) . . . a(kn, εn, tn))

= wlimt1→+∞

exp(−iε2w(k1)t1)

× Λ(k1, ε1)T (a(k1, ε1, t1) . . . a(kn, εn, tn)). (9.23)

Indeed, this follows from the remark that if t1 > t2, . . . , > tn, then

a(k1, ε1, t1)T (a(k2, ε2, t2) . . . a(kn, εn, tn))

= T (a(k1, ε1, t1) . . . a(kn, εn, tn))).

Page 211: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 164

164 Mathematical Foundations of Quantum Field Theory

Similarly,

T (a(k2, ε2, t2) . . . a(kn, εn, tn))ain(k1, ε1)

= limt1→−∞

Λ(k1, ε1) exp(−iε1ω(k1)t1)

× T (a(k1, ε1, t1) . . . a(kn, εn, tn)). (9.24)

Combining the relations (9.23) and (9.24), we obtain

aout(k1, ε1)T (a(k2, ε2, t2) . . . a(kn, εn, tn))

− T (a(k2, ε2, t2) . . . a(kn, εn, tn))ain(k1, ε1)

= Λ(k1, ε1)

∫ ∞−∞

∂t1(exp(−iε1ω(k1)t1)T (a(k1, ε1, t1)

. . . a(kn, εn, tn)))dt1

=

∫ ∞−∞

L1T (a(k1, ε1, t1) . . . a(kn, εn, tn))dt1. (9.25)

Here, we used the notation

Lif(k1, ε1, t1, . . . ,kn, εn, tn) = Λ(ki, εi)∂

∂ti(exp(−iεiω(ki)t)

× f(k1, ε1, t1, . . . ,kn, εn, tn)).

It is useful to note the simplest form of (9.25)

aout(k, ε)− ain(k, ε) =

∫ ∞−∞

Λ(k, ε)∂

∂t(exp(−iεω(k)t)a(k, ε, t))dt

=

∫ ∞−∞

La(k, ε, t)dt

(Yang–Feldman equation).

In the case of CAR, we analogously have

aout(k1, ε1)T (a(k2, ε2, t2) . . . a(kn, εn, tn))

−(−1)n−1T (a(k2, ε2, t2) . . . a(kn, εn, tn))ain(k1, ε1)

=

∫ ∞−∞

Λ(k1, ε1)∂

∂t1(exp(−iε1ω(k1)t1)T (a(k1, ε1, t1)

. . . a(kn, εn, tn)))dt1

=

∫ ∞−∞

L1T (a(k1, ε1, t1) . . . a(kn, εn, tn))dt1.

Page 212: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 165

The Scattering Matrix for Translation-Invariant Hamiltonians 165

Applying the relation (9.25) several times, we can express the matrix

entries of the scattering matrix in terms of Green functions.

We will give another, shorter, proof. Let us introduce the oper-

ators c(k, ε, t) = S∗−a(k, ε, t)S−. We can easily express the Green

function of the Hamiltonian H in terms of these operators:

Gn(k1, ε1, t1, . . . ,kn, εn, tn)

= 〈T (a(k1, ε1, t1) . . . a(kn, εn, tn))S−θ, S−θ〉

= 〈T (c(k1, ε1, t1) . . . c(kn, εn, tn))θ, θ〉 .

Multiplying (9.25) on the left by S∗+ = SS∗− and on the right by S−and using the relations (9.13) and (9.14), we obtain

[S · T (c(k2, ε2, t2) . . . c(kn, εn, tn)), b(k1, ε1)]

= −∫ ∞−∞

L1S · T (c(k1, ε1, t1) . . . c(kn, εn, tn))dt1. (9.26)

Applying this formula several times, we see that

[. . . [S, b(k1, ε1)] . . . b(kn, εn)]

= (−1)n∫Ln . . . L1(S · T (c(k1, ε1, t1) . . . c(kn, εn, tn)))dt1 . . . dtn.

(9.27)

The expression for the functions σm,n can be obtained from the

remark that

σm,n(p1, . . . ,pm|q1, . . . ,qn)

= (−1)n⟨[. . . [[. . . [S, b+(q1)] . . . b+(qn)]b(p1)] . . . b(pm)]θ, θ

⟩.

(9.28)

Namely, from (9.27) and (9.28) and the relation Sθ = θ, it follows

that

σm,n(p1, . . . ,pm|q1, . . . ,qn)

= (−1)m∫Lm+n . . . L1Gm+n(q1, 1, t1, . . . ,qn, 1, tn,p1,−1,

tn+1, . . . ,pm,−1, tm+n)dt1 . . . dtm+n.

Page 213: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 166

166 Mathematical Foundations of Quantum Field Theory

In order to obtain (9.22) from the above formula, we should take the

Fourier transform over the variables t1, . . . , tm+n, using that∫Lf(k, ε, t)dt =

∫ ∞−∞

Λ(k, ε)d

dt(exp(−iεω(k)t)f(k, ε, t))dt

= iεΛ(k, ε) limω→ω(k)

(ω − ω(k))

×∫ ∞−∞

exp(−iεωt)f(k, ε, t)dt.

For a more rigorous proof of (9.22), see, for example, Robertson

et al. (1980). Such a proof can be obtained also from considerations

of Chapter 13.

9.3 The adiabatic definition of scattering matrix

Let us now provide a definition of scattering matrix of a translation-

invariant Hamiltonian that does not use the notion of operator

realization. Let H be a translation-invariant Hamiltonian of the form

H = H0 + V , where H0 =∫ε(k)a+(k)a(k)dk.

We will consider the adiabatic S-matrix SΩα = SΩ

α (∞,−∞) corre-

sponding to the pair of operators (HΩ, H0Ω) (the Hamiltonian HΩ is

defined as the Hamiltonian H with volume cutoff, see Section 8.1).

Definition 9.1. The scattering matrix for a Hamiltonian H is an

operator in the space Has that has the matrix entries

〈p1, . . . ,pm|S|q1, . . . ,qm〉 = limα→0

limΩ→∞

(L

) 32

(m+n)

×⟨p1, . . . ,pm|SΩ

α |q1, . . . ,qm⟩ ⟨θ|SΩ

α |θ⟩m+n

2−1√∏m

i=1 〈pi|SΩα |pi〉

∏nj=1 〈qj |SΩ

α |qj〉(9.29)

in the generalized basis b+(p1) . . . b+(pn)θ. Here Has, as in Sec-

tion 9.2, denotes the space of Fock representations of CR (the

operator generalized functions b+(k), b(k) acting on Has satisfy CR

Page 214: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 167

The Scattering Matrix for Translation-Invariant Hamiltonians 167

and k runs over E3). We use the notation

〈p1, . . . ,pm|S|q1, . . . ,qn〉 =⟨Sb+(q1) . . . b+(qn)θ, b+(p1)

. . . b+(pm)θ⟩,⟨

p1, . . . ,pm|SΩα |q1, . . . ,qn

⟩=⟨SΩα a

+q1. . . a+

bqnθ, a+

p1. . . a+

pmθ⟩

(recall that the operator SΩα , like the operator HΩ, acts on the Fock

space FΩ and the symbols a+p , ap denote the operators a+(φp), a(φp),

where φp(x) = L−3/2 exp(ipx) and p runs over the lattice TΩ,

consisting of the vectors 2πL n).

We should explain the meaning of convergence in (9.29), since the

functions of continuous argument are defined as limits of functions

with arguments in a lattice. We dealt with a similar situation

in Section 8.2; here and in the rest of the book, we will deal

with similar limits analogously (i.e. in the sense of generalized

functions). Namely, the relation (9.29) means that for any test

function φ(p1, . . . ,pm|q1, . . . ,qn)∫φ(p1, . . . ,pm|q1, . . . ,qn) 〈p1, . . . ,pm|S|q1, . . . ,qn〉 dmpdnq

= limα→∞

limΩ→∞

(2π

L

) 32

(m+n) ∑pi,qj∈TΩ

φ(p1, . . . ,pm|q1, . . . ,qn)

×⟨p1, . . . ,pm|SΩ

α |q1, . . . ,qm⟩ ⟨θ|SΩ

α |θ⟩m+n

2−1√∏m

i=1 〈pi|SΩα |pi〉

∏nj=1 〈qj |SΩ

α |qj〉.

We can represent the definition of scattering matrix in a different

form, noting that⟨p1, . . . ,pm|SΩ

α |q1, . . . ,qm⟩ ⟨θ|SΩ

α |θ⟩m+n

2−1√∏m

i=1 〈pi|SΩα |pi〉

∏nj=1 〈qj |SΩ

α |qj〉

=⟨p1, . . . ,pm|UΩ

α SΩαU

Ωα |q1, . . . ,qm

⟩, (9.30)

where UΩα = exp[i(C +

∑k r

Ωα (k)a+

k ak)] (the sum is taken over

the lattice TΩ), C = i2 ln

⟨θ|SΩ

α |θ⟩, rΩα (k) = i

2 ln〈k|SΩ

α |k〉〈θ|SΩ

α |θ〉(the

Page 215: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 168

168 Mathematical Foundations of Quantum Field Theory

relation (9.30) follows from UΩα a

+k1. . . a+

knθ = exp[i(C+rΩ

α (k1)+· · ·+rΩα (kn))]a+

k1. . . a+

knθ). The relation (9.30) follows from the fact that

S = limα→0

limΩ→∞

UΩα S

ΩαU

Ωα ,

where the convergence of operators is understood in the sense of

the convergence of the matrix entries in the basis a+k1. . . a+

knθ to the

matrix elements in the generalized basis b+(k1) . . . b+(kn)θ (further

in the book, we will understand analogous situations concerning the

convergence of operators acting on the space FΩ to operators on the

space Has in the same way).

Using the results of Chapter 4 (see Sections 4.2 and 4.3), one can

prove that the S-matrix defined by the relation (9.29) satisfies the

conditions Sθ = θ, Sb+(k)θ = b+(k)θ (the stability of the vacuum

and single-particle states).

Indeed, let φΩ(λ) (correspondingly, φΩk (λ)) be a stationary state

of the Hamiltonian HΩ(λ) = H0Ω + λVΩ continuously depending on

the parameter λ and converging to θ for λ = 0 (correspondingly

a+k θ). Furthermore, we let εΩ(λ) and εΩ(k, λ) be the energy levels of

the Hamiltonian HΩ(λ)

ρΩ =1

α

∫ 1

0

1

µ(εΩ(µ)− εΩ(0))dµ,

ρΩ(k) =1

α

∫ 1

0

1

λ(εΩ(k, λ)− εΩ(k, 0))dλ.

It follows from (4.11) that

θ = limα→0

exp(2iρΩ)SΩα θ,

a+k θ = lim

α→0exp(2iρΩ(k))SΩ

α a+k θ.

We conclude that for α→ 0,⟨k|SΩ

α |k′⟩≈ exp(2iρΩ(k))δkk′ , (9.31)⟨

θ|SΩα |θ⟩≈ exp(2iρΩ), (9.32)

and therefore,

UΩα ≈ UΩ

α = exp

(−iρα − i

∑k

(ρΩ(k)− ρΩ)a+k ak

). (9.33)

Page 216: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 169

The Scattering Matrix for Translation-Invariant Hamiltonians 169

Using Equations (9.31) and (9.32), we see that

limα→0

UΩα S

ΩαU

Ωα θ = θ, (9.34)

limα→0

UΩα S

ΩαU

Ωα a

+k θ = a+

k θ. (9.35)

Multiplying the relations (9.34) and (9.35) by the vectors a+k1. . . a+

knθ

and taking the limit Ω→∞ in the resulting matrix elements of the

operator UΩα S

ΩαU

Ωα , we obtain a proof of the stability of the vacuum

and single-particle states.

Let us also note that from the previous reasoning, we can obtain

the “almost unitarity” of the operators UΩα when α → 0. More

precisely, the scattering matrix can be represented in the form

S = limα→0

limΩ→∞

UΩα S

Ωα U

Ωα , (9.36)

where UΩα = exp(−iρΩ − i

∑k(ρΩ(k) − ρΩ)a+

k ak) are unitary

operators.

Note that the proofs above are clearly not rigorous.

We can modify the scattering matrix definition given above by

introducing the notion of N -equivalent and S-equivalent operators,

acting in Fock space.

Two operators S and S′ on Fock space Tas are N -

equivalent if there exist unitary operators U and V of the form

exp[i(∫ν(k)b+(k)b(k)dk + C)], such that S′ = USV . If U = V −1,

then we say that the operators S and S′ are S-equivalent.

The definitions of N -equivalence and S-equivalence for operators

acting on the Fock space FΩ only differ by assuming that the

operators U and V take the form exp[i(∑

k νka+k ak + C)].

Definition 9.2. The operator S on the space Has is called the

scattering matrix of the Hamiltonian H if it satisfies the conditions

Sθ = θ, Sb+(k)θ = b+(k)θ and it can be represented in the form

S = limα→0

limΩ→∞

SΩα , (9.37)

where SΩα are operators that are N -equivalent to the operators SΩ

α .

Page 217: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 170

170 Mathematical Foundations of Quantum Field Theory

Let us recall that the limit in the expression of the form (9.37) is

understood in the sense of the convergence of the matrix entries of

SΩα in the basis a+

k1. . . a+

knθ to the matrix entries of the operator S

in the generalized basis b+(k1) . . . b+(kn)θ.

It is clear from the relation (9.36) that a scattering matrix

specified by Definition 7.1 is also a scattering matrix in the sense of

Definition 9.1. However, unlike Definition 7.1, Definition 9.1 does not

specify a scattering matrix unambiguously: using the ambiguity in

the choice of the operators SΩα , it is easy to see that an operator that

is S-equivalent to the scattering matrix is also a scattering matrix in

the sense of Definition 9.1. The converse of this observation holds as

well: two scattering matrices of a Hamiltonian H are S-equivalent.

Note that in Section 9.2 the scattering matrix was also defined up to

S-equivalence.

In the framework of perturbation theory, one can prove that

Definition 9.1 is equivalent to the definition of the scattering matrix

given in Section 9.2.2 However, to show this equivalence, we need to

make a few assumptions about the Hamiltonian H (it is sufficient

to require that the Hamiltonian belongs to the class M defined in

Section 11.1 and the function ε(k), entering the definition of the

Hamiltonian H0, satisfies the condition ε(k1 + k2) < ε(k1) + ε(k2)).

We consider the connection between adiabatic S-matrices and

scattering matrices in Section 10.6 (in the framework of axiomatic

scattering theory) and in Section 11.5 (for Hamiltonians that do

not generate vacuum polarization). The proof of the equivalence of

Definition 9.1 with other definitions of scattering matrices in the

framework of perturbation theory with partial summation can be

obtained by modifying the considerations in Section 10.6. Another

less rigorous proof can be found in Likhachev et al. (1972).

9.4 Faddeev’s transformation and equivalence

theorems

Let us consider a Hamiltonian H = H0 +V of the form (9.29). In the

case of vm,0 ≡ vm,1 ≡ 0, the scattering matrix of H can be defined

2More precisely, to show this equivalence, we should perform a partial summa-tion of the series in perturbation theory for SΩ

α .

Page 218: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 171

The Scattering Matrix for Translation-Invariant Hamiltonians 171

by using the standard relations of formal scattering theory (5.3) (see

Section 9.1). If the Hamiltonian H is arbitrary, then under certain

conditions, we may replace H with an equivalent, in a certain sense,

Hamiltonian H ′ = H ′0 + V ′ for which the scattering matrix can be

defined as

S = slimt→∞t0→−∞

exp(iH ′0t) exp(−iH ′(t− t0)) exp(−iH ′0t0). (9.38)

This scattering matrix can be regarded as the scattering matrix for

the original Hamiltonian H.

The following discussion will proceed in the framework of per-

turbation theory. This means that the Hamiltonian H0 + V will be

included in the family of Hamiltonians H0 + gV depending on a

parameter g (coupling constant) and the Hamiltonian H ′ will be

constructed as a series in powers of g.

To begin, let us prove a few helpful statements. Consider the

equation

[h0, x] = r, (9.39)

where h0, x, r are operators on the Fock space FΩ = F (L2(Ω))

corresponding to the finite-volume Ω:

h0 =∑k∈TΩ

ε(k)a+k ak, (9.40)

where the positive function ε(k) satisfies

ε(k1 + k2) < ε(k1) + ε(k2). (9.41)

Let us write down the unknown operator x and the known operator

r in normal form:

x =∑m,n

∑ki,pj

ξm,n(k1, . . . ,km|p1, . . . ,pn)a+k1

. . . a+kmap1 . . . apn ; (9.42)

r =∑m,n

∑ki,pj

ρm,n(k1, . . . ,km|p1, . . . ,pn)a+k1

. . . a+kmap1 . . . apn

(in these formulas, just as in the following analogous formulas, the

sums are taken over ki,pj ∈ TΩ). If we calculate the commutator

Page 219: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 172

172 Mathematical Foundations of Quantum Field Theory

[h0, x], we obtain the relation between the unknown functions ξm,nand the known functions ρm,n:

(ε(k1) + · · ·+ ε(km)− ε(p1)− · · · − ε(pn))

× ξm,n(k1, . . . ,km|p1, . . . ,pn)

= ρm,n(k1, . . . ,km|p1, . . . ,pn). (9.43)

If the operator is presented in normal form, then the sum-

mands containing m creation operators and n annihilation oper-

ators will be called summands of type (m,n). Summands of type

(m, 0), (0,m), (n, 1), and (1, n), where m ≥ 1, n ≥ 2, will be called

bad summands.

Let us prove the following lemma.

Lemma 9.1. If an operator r commutes with the momentum

operator PΩ =∑

ka+k ak and only contains bad summands, then

equation (9.39) is solvable. If we also assume that the operator

x commutes with the momentum operator and contains only bad

summands, then equation (9.39) has a unique solution which will

be denoted by Γ(r). If the operator r is Hermitian, then so is the

operator Γ(r).

In order to prove this lemma, we first note that from the

condition (9.43), we can find the functions ξm,n in terms of the

functions ρm,n if and only if for all values of k1, . . . ,km,p1, . . . ,pnsatisfying ε(k1) + · · · + ε(km) = ε(p1) + · · · + ε(pn), we have

ρm,n(k1, . . . ,km|p1, . . . ,pn) = 0. If we assume ε(k1) + . . . +

ε(km) > 0, then for the functions ρm,0 and ρ0,m, we can find the

functions ξm,0 and ξ0,m. Furthermore, from the assumption that r

commutes with the momentum operator, it follows that the quantity

ρm,n(k1, . . . ,km|p1, . . . ,pn) is non-zero only if k1 + · · · + km =

p1 + · · ·+ pn. This allows us to say that the inequality

ε(k1) + · · ·+ ε(km) > ε(p),

which holds in the case of k1 + · · · + km = p,m > 1, guarantees

the solvability of (9.43) for m > 1, n = 1 and n > 1,m = 1 (this

Page 220: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 173

The Scattering Matrix for Translation-Invariant Hamiltonians 173

follows from (9.41)). Therefore, under the conditions of the lemma,

equation (9.39) is solvable. The proofs of the rest of the statements

in the lemma are trivial (for example, the uniqueness follows from

the fact that under the condition ε(k1) + · · · + ε(km) 6= ε(p1) +

· · ·+ ε(pn) we can find a unique ξm,n(k1, . . . ,km|p1, . . . ,pn) for each

ρm,n(k1, . . . ,km|p1, . . . ,pn)).

Let us take a Hermitian operator h, acting on the Fock space

FΩ, that commutes with the momentum operator PΩ, and consider

the question of finding a unitary operator w such that the operator

h′ = whw−1 contains no bad summands. We will show that in the

framework of perturbation theory, the operator w can be constructed.

More precisely, we will show that for an operator of the form

h = h0 + gv, where h0 takes the form (9.40) considered above

and g is a scalar parameter (coupling constant), the operators h′

and w can be constructed as formal series in powers of g. The

constructed operators h′ and w will commute with the momentum

operator PΩ; this implies that the operator h′ can be represented

in the form h′ = c + h′0 + v′, where c is a constant [a summand

of type (0, 0)], h′0 =∑ωka

+k ak [a summand of type (1, 1)],

v′ =∑

m,n≥2

∑v′m,n(k1, . . . ,km|p1, . . . ,pn)δp1+···+pn

k1+···+kma+k1. . . a+

kmap1

. . . apn (a sum of summands of type (m,n) with m ≥ 2, n ≥ 2). We

will assume that the operator w takes the form

w = exp(−iα),

where α is a Hermitian operator that can be represented as a series

in powers of g:

α =∞∑n=1

gnαn.

We will use the formula3

h′ = exp(−iα)h exp(iα) =

∞∑n=0

in

n![. . . [h, α] . . . , α].

3It is easy to see that the operators C(t) = exp(−itα)b exp(itα) and D(t) =∑∞n=0

(it)n

n![. . . [b, α] . . . , α] satisfy identical equations i dC(t)

dt= [α,C(t)], i dD(t)

dt=

[α,D(t)] with the same initial condition C(0) = D(0) = b; henceC(t) = D(t).

Page 221: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 174

174 Mathematical Foundations of Quantum Field Theory

Equating the terms of the same degree with respect to g, we obtain

the equation

h′n = i[h0, αn] + qn,

where qn denotes the operator that can be expressed in terms of

h0, v, and αk with k ≤ n − 1, and h′ngn denotes an order-n term

in the expansion of the operator h′ in powers of g. We will show

by induction with respect to n that we can find operators αn and

h′n such that the operator h′n contains no bad summands. This can

be done if we take as h′n the sum of good summands in the normal

form of the operator qn. Then, the operator h′n − qn contains only

bad terms and satisfies the conditions imposed on the operator r in

the lemma. Therefore, the operator αn satisfying the above equation

exists; let us define it by the formula αn = Γ(h′n − qn). It is easy to

check that the operators qn are Hermitian; it follows that h′n and αnare Hermitian as well.

By this method, we can inductively construct the operators h′nand αn such that the operator h′ =

∑h′ng

n does not contain bad

summands and the operator w = exp(−iα) = exp(−i∑αng

n)

establishes a unitary equivalence between the operators h and h′

(i.e. h′ = whw−1).

Note that our construction for the operators h′ and w is unam-

biguous; we will call this construction the Faddeev construction and

the unitary operator w the Faddeev transformation.

It is worth noting that the operators h′ and w were constructed in

the framework of perturbation theory; in other words, these operators

were constructed in the form of formal series in powers of g. It is not

clear whether these series converge.

Moreover, as was done in the lemma and in the construction

of the Faddeev transformation, we used the term “operator” for

expression (9.42), but we did not prove that this formal expression

specifies an operator on Fock space. In the situations where we use

Faddeev construction, it is easy to check that the formal expressions

hn and αn specify Hermitian operators; the proof is based on the

considerations in Section 6.2.

Page 222: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 175

The Scattering Matrix for Translation-Invariant Hamiltonians 175

Let us consider the Faddeev transformation from another point

of view.

Consider the operators

b+k = w−1a+kw; bk = w−1akw. (9.44)

It is clear that these operators satisfy the same commutation relations

as the operators a+k , ak:

[bk, bk′ ]∓ = [b+k , b+k′ ]∓ = 0, [bk, b

+k′ ]∓ = δk,k′ .

A transformation from the operators ak, a+k to the operators bk, b

+k

satisfying the same commutation relations is called a canonical

transformation. From the relation h = w−1h′w, it follows that the

operator h can be expressed in terms of the operators b+k , bk in

the same way that h′ can be expressed with a+k , ak; in this way,

representing the operator h with the operators b+k , bk in normal form,

we obtain an expression without bad summands.

Let us now return to the translation-invariant Hamiltonian H =

H0 + gV , where H0 and V satisfy the formulas (9.1) and (9.2).

Suppose that ε(k) is a smooth function, whose derivatives do not

grow faster than a power, and that wm,n belongs to the space S. Let

us also assume that the function ε(k) satisfies equation (9.41).

Let us now apply the Faddeev construction to the Hamiltonian

HΩ obtained from H by means of a volume cutoff. The Hamiltonian

obtained through this construction contains no bad summands and

will be denoted by H ′Ω. The operator establishing a unitary equiv-

alence between H ′Ω and HΩ will be denoted by WΩ = exp(−iAΩ).

It is easy to understand the behavior of the operators H ′Ω and AΩ

as Ω → ∞. Namely, the operator H ′Ω can be written in the form

H ′Ω = CΩ +H ′Ω0 + V ′Ω, where CΩ is a constant that grows linearly

as Ω→∞. We obtain

H ′Ω0 =∑

ωΩk a

+k ak,

V ′Ω =∑m,n

∑(L

)−3(m+n2−1)

v′Ωm,n(k1, . . . ,km|p1, . . . ,pn)

× δp1+···+pnk1+···+km

a+k1. . . a+

kmap1 . . . apn ,

Page 223: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 176

176 Mathematical Foundations of Quantum Field Theory

the functions ωΩk , v

′Ωm,n converge to a limit as Ω → ∞.4 Let us

introduce the notation

limωΩk = ω(k);

lim v′Ωm,n(k1, . . . ,km|p1, . . . ,pn)

= v′m,n(k1, . . . ,km|p1, . . . ,pn), (9.45)

H ′0 =

∫ω(k)a+(k)a(k)dk;

V ′ =∑m,n

∫v′m,n(k1, . . . ,km|p1, . . . ,pn)

× δ(k1 + · · ·+ km − p1 − · · · − pn)a+(k1)

. . . a+(km)a(p1) . . . a(pn)dmkdnp. (9.46)

The translation-invariant Hamiltonian H ′ = H ′0+V ′ does not contain

bad summands; we will say that this Hamiltonian is obtained from

the Hamiltonian H by means of Faddeev transformation. One can

say that

H ′ = limΩ→∞

(H ′Ω − CΩ) (9.47)

(the relation (9.47) can be understood as a shorthand version

of (9.46)).

The operator AΩ behaves the same way in the limit Ω → ∞,

namely, it can be written in the form

AΩ =∑m,n

(L

)−3(m+n2−1)∑

αΩm,n(k1, . . . ,km|p1, . . . ,pn)

× δp1+···+pnk1+···+km

a+k1. . . a+

kmap1 . . . apn ,

4The functions ωΩk , v

′Ωm,n are expressed in the form of a series in powers of g;

we understand the convergence as Ω → ∞ in the sense of convergence for everypower of g. It is useful to note that we can find such functions νrm,n ∈ S and σr(k)that do not grow faster than a power function and that satisfy the conditions|(v′Ωm,n)r| ≤ νrm,n, |(ωΩ

k )r| ≤ σr(k) (by the symbol fr here, we denote a term oforder r in g in the expansion of the function f).

Page 224: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 177

The Scattering Matrix for Translation-Invariant Hamiltonians 177

where the functions αΩm,n(k1, . . . ,km|p1, . . . ,pn) have a limit as

Ω→∞,

αm,n(k1, . . . ,km|p1, . . . ,pn) = limΩ→∞

αΩm,n(k1, . . . ,km|p1, . . . ,pn)

(9.48)

in each term in the series with respect to g. Using the formal

translation-invariant expression

A =∑m,n

∫αm,n(k1, . . . ,km|p1, . . . ,pn)

× δ(k1 + · · ·+ km − p1 − · · · − pn)a+(k1)

. . . a+(km)a(p1) . . . a(pn)dmkdnp,

we can write the relation (9.49) in the form A = limΩ→∞AΩ.

The behavior of the operator WΩ for large Ω is more complicated;

one cannot find a formal translation-invariant expression for this

operator in the limit Ω → ∞. By construction, the expression for

V ′ has no summands of type (m, 0) or (m, 1) and therefore the

formal expression for H ′ can specify an operator H ′ in Fock space.

In this case, the scattering matrix for the Hamiltonian H ′ can be

constructed like an S-matrix for the pair of operators (H ′, H ′0) (see

Section 9.1). In Chapter 11, we will show that the S-matrix for the

pair of operators (H ′, H ′0) can be constructed in the framework of

perturbation theory.

Faddeev proposed to define the scattering matrix for the Hamil-

tonian H as the S-matrix constructed for the pair of operators

(H ′, H ′0). It can be proved that Faddeev’s definition of the scattering

matrix is equivalent to the other definitions we have considered (the

proof can only be given in perturbation theory, since Faddeev’s

construction is based on perturbation theory). The proof of this

statement is provided in Section 11.3, while here we will instead

turn our attention to some generalizations of Faddeev’s construction.

First, we will represent this construction in a slightly different form.

Page 225: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 178

178 Mathematical Foundations of Quantum Field Theory

It is easy to check that the S-matrix (9.38) can be written in the form

S = limt→∞,t0→−∞

limΩ→∞

exp(iH ′Ω0 t) exp(−i(H ′Ω − CΩ)(t− t0))

× exp(−iH ′Ω0 t0)

(the limit should be understood in the sense explained in Section 9.3).

Using the relation H ′Ω = WΩHΩ(WΩ)−1, we can write that

S = limt→∞,t0→−∞

limΩ→∞

exp(i(H ′Ω0 + CΩ)t)WΩ exp(−iHΩ(t− t0))

× (WΩ)−1 exp(−i(H ′Ω0 + CΩ)t0). (9.49)

The operator (WΩ)−1 is called a dressing operator, since it trans-

forms the bare vacuum θ and the bare single-particle state a+k θ into

stationary states of the Hamiltonian HΩ that can be described as

the dressed vacuum and dressed single-particle states (in fact, the

operator (WΩ)−1 transforms stationary states of the Hamiltonian

H ′Ω into stationary states of the Hamiltonian HΩ and θ and a+k θ are

clearly stationary states of H ′Ω).

The formula (9.49) suggests the following generalization of a

dressing operator (more precisely, a family of dressing operators DΩ

depending on the parameter Ω).

The family of operators DΩ acting on the space FΩ is called a

family of dressing operators for the Hamiltonian H if there exists a

Hamiltonian HΩkb of the form γΩ+

∑k ν

Ωk a

+k ak such that the S-matrix

for the Hamiltonian H can be written in the form

S = limt→∞,t0→−∞

limΩ→∞

exp(iHΩkbt)(D

Ω)−1

× exp(−iHΩ(t− t−))DΩ exp(−iHΩkbt0).

Formula (9.49) shows that the family of operators (WΩ)−1 is a family

of dressing operators in the sense of the above definition.

In Section 11.4, we will describe a broad class of families of

dressing operators that includes the operators (WΩ)−1.

The theorem establishing the equivalence of Faddeev’s definition

with other definitions of scattering matrices is a particular case

of a theorem of invariance of scattering matrices under canonical

Page 226: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 179

The Scattering Matrix for Translation-Invariant Hamiltonians 179

transformations (theorems of this kind are called equivalence theo-

rems).

Let us give a general definition of a canonical transformation.

Let us express the symbols b+(k), b(k) in terms of the symbols

a+(k), a(k) by using the formulas

b(k) =∑m,n

∫σm,n(k1, . . . ,km|p1, . . . ,pn)

× δ(k + k1 + · · ·+ km − p1 − · · · − pn)a+(k1)

. . . a+(km)a(p1) . . . a(pn)dmkdnp; (9.50)

b+(k) =∑m,n

∫σm,n(k1, . . . ,km|p1, . . . ,pn)

× δ(k + k1 + · · ·+ km − p1 − · · · − pn)a+(p1)

. . . a+(p1)a(km) . . . a(k1)dmkdnp, (9.51)

where σm,n are smooth functions whose derivatives do not grow faster

than a power.

The transformation from the symbols a+(k), a(k) into the sym-

bols b+(k), b(k) is called a canonical transformation if the sym-

bols b+(k), b(k) obey the same commutation (or anticommutation)

relations as the symbols a+(k), a(k) (i.e. obey CR). In the case of

CCR, we can calculate the commutators for the symbols b+(k), b(k)

by using CCR for the operators a+(k), a(k), distributivity, and

the relation [A,B,C] = [A,C]B + A[B,C] (we can calculate the

anticommutator in the case of CAR analogously).

The above equivalence theorem can be reformulated (not quite

rigorously) in the following way.

Theorem 9.1. The canonical transformation (9.50) transforms the

translation-invariant Hamitlonian H of the form (9.1) into the

translation-invariant Hamiltonian H having the same scattering

matrix.

(The Hamiltonian H can be constructed in the following way:

in the expression of the Hamiltonian H with respect to the symbols

Page 227: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 180

180 Mathematical Foundations of Quantum Field Theory

a+(k), a(k), we substitute against the symbols b+(k), b(k), expressed

in terms of a+(k), a(k) using the formula (9.50); we then write

the obtained expression into normal form using CR, discarding any

infinite constants that arise.)

It is easy to see that the Hamiltonian H ′ obtained from the

translation-invariant Hamiltonian using Faddeev’s construction is

connected with the Hamiltonian H through the canonical transfor-

mation (this canonical transformation is obtained from the canonical

transformation (9.44) relating the operators b+k , bk with the operators

a+k , ak in the limit Ω→∞).

It is therefore clear from the equivalence theorem that the

scattering matrices for the operators H and H ′ coincide; this

demonstrates that Faddeev’s construction leads to a conventional

scattering matrix.

We will prove the equivalence theorem in the framework of

perturbation theory in Section 11.3.

Note that for linear canonical transformations, the considerations

in Section 11.3 prove the equivalence theorem outside the framework

of perturbation theory.

9.5 Semiclassical approximation

We have already seen that the Hamiltonians of quantum field theory

can be obtained through the quantization of classical systems with

an infinite number of degrees of freedom. We will leverage this fact

to obtain approximate solutions to problems in quantum field theory.

As in Section 8.3, we will not make the assumption that ~ = 1, as

we do in the rest of this book.

Let us consider a quantum mechanical system with a finite num-

ber of degrees of freedom that is described by the Hamiltonian (8.22).

In order to find weakly excited states (stationary states whose

energy is close to the energy of the ground state), we can apply

the following method. Let us find the minimum of the function

U(q1, . . . , qn) (potential energy); suppose the minimum occurs at the

point (q01, . . . , q

0n). Let us introduce the new variables xi = qi − q0

i

(these should be understood as the deviation from the classical

Page 228: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 181

The Scattering Matrix for Translation-Invariant Hamiltonians 181

equilibrium state) and let us decompose the function U as a series in

powers of xi:

U(q1, . . . , qn) = U(q01, . . . , q

0n) +

1

2

∑i,j

cijxixj

+1

3!

∑i,j,k

cijkxixjxk + · · · .

If the matrix cij is positive definite, then the energy of a weakly

excited state can be determined by the following approximate

expression:

U(q01, . . . , q

0n) +

~2

∑ωi + ~

∑niωi, (9.52)

where ω21, . . . , ω

2n are eigenvalues of the matrix cij and ni = 0, 1, 2, . . .

are non-negative integers.

To check this, we will write the Hamiltonian (8.22) in terms of

the operators pi and xi = qi − q0i , satisfying the relations [pi, pj ] =

[xi, xj ] = 0, [pi, xj ] = ~i δij ; we obtain

H = H0 + V ;

H0 =1

2

∑p2i +

1

2

∑cij xixj + U(q0

1, . . . , q0n);

V =1

3!

∑cijkxixj xk + · · · .

The operators pi, xi satisfy the same commutation relations as pi, qi,

i.e. the transformation between these operators can be considered to

be a canonical transformation.

The eigenvalues of the operator H0 are specified by the expres-

sion (9.52) (see Section 2.6). The eigenvalues of the operator H =

H0 + V can be calculated in the framework of perturbation theory

by considering the operator V as a perturbation. It is easy to see

that the corrections to the energies of a weakly excited state are

small. This becomes evident if we express H0 and V in terms of the

operators a+i , ai, satisfying CCR (we rely on the considerations in

Page 229: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 182

182 Mathematical Foundations of Quantum Field Theory

Section 2.6):

H0 = U(q01, . . . , q

0n) +

1

2~∑

ωi + ~∑

ωia+i ai;

V =∑

m+n≥3

~( 12

)(m+n)∑

i1,...,im,j1,...,jn

Γm,ni1,...,im,j1,...,jna+i1. . . ajn

(the perturbation series here can be rewritten as a series in powers

of ~1/2, hence the correction to the energy has order at least ~3/2).

We see that the analysis of weakly excited states of Hamiltonian

H is based on the canonical transformation (transforming to the

operators pi, xi) where we single out the quadratic part of H0 and

consider V as a perturbation.

Let us now consider a classical system specified by the

Hamiltonian

H(π, φ) =1

2

∑i

∫π2i (x)dx +

∑m

∑i1,...,im

∫Vi1,...,im(x1, . . . ,xm)

×φi1(x1) . . . φim(xm)dmx; (9.53)

Vi1,...,im(x1, . . . ,xm) = vi1,...,im(x1−xm, . . . ,xm−1−xm). Quantizing

this system leads to the Hamiltonian

H =1

2

∑i

∫π2i (x)dx

+∑m

∑i1,...,im

∫Vi1,...,im(x1, . . . ,xm)φi1(x1)

. . . φim(xm)dmx, (9.54)

where

[πi(x), πj(x′)] = [φi(x), φj(x

′)] = 0,

[πi(x), φj(x′)] =

~iδi,jδ(x− x′)

Page 230: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 183

The Scattering Matrix for Translation-Invariant Hamiltonians 183

(for a slightly less general Hamiltonian, see (8.23) discussed in

Section 8.3; everything that is known for the Hamiltonian (8.23) can

be easily proven for the Hamiltonians (9.54)).

As in the case of a finite number of degrees of freedom, the analysis

of the Hamiltonian (9.54) can be performed with the canonical

transformation: instead of the symbols φi(x), we should introduce the

symbols ξi(x) = φi(x)− αi, where αi are real numbers; the symbols

πi(x), ξi(x) satisfy the same commutation relations as πi(x), φj(x)

(we could introduce the symbol ξi(x) by the more general formula

ξi(x) = φi(x) − αi(x), but since our Hamiltonian is translation-

invariant it is natural to assume that αi(x) do not depend on x).

Let us consider a function of n variables

ν(φ1, . . . , φn) =∑m

∑i1,...,im

νi1,...,imφi1,...,im ,

where

νi1,...,im =

∫vi1,...,im(x1, . . . ,xm−1)dm−1(x)

(the function ν has the physical meaning of energy density of the

classical system in the state π1(x) = · · · = πn(x) ≡ 0, φ1(x) ≡φ1, . . . , φn(x) ≡ φn). The numbers α1, . . . , αn specifying the canoni-

cal transformation will be found from the condition

minφ1,...,φn

ν(φ1, . . . , φn) = ν(α1, . . . , αn).

The point in n-dimensional space α = (α1, . . . , αn), where the

function ν(φ1, . . . , φn) achieves its minimum, will be called the

classical vacuum.

It is easy to check, under certain conditions, that one can justify

the choice of the numbers α1, . . . , αn by considering a limit of systems

with finite number of degrees of freedom.

Replacing φi(x) with ξi(x) +αi in (9.54), we obtain the Hamilto-

nian H ′ with the same scattering matrix by the equivalence theorem.

Extracting from H ′ the quadratic terms H0, we can represent the

Page 231: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 184

184 Mathematical Foundations of Quantum Field Theory

Hamiltonian H ′ in the form H0 + V , where

H0 =1

2

∑i

∫π2i (x)dx +

∑i,j

1

2

∫cij(x− y)

× ξi(x)ξj(y)dxdy,

V =∑m≥3

∑i1,...,im

∫ci1,...,im(x1 − xm, . . . ,xm−1 − xm)

× ξi1(x1) . . . ξim(xm)dmx.

Using the ideas of Section 8.3, we can construct an operator

realization of the Hamiltonian H0. We see that H0 describes n types

of particles with energy dependence on the momentum, defined by

the formula Ej(p) = ~ωj(p~ ), where ω21(k), . . . , ω2

n(k) are eigenvalues

of the matrix cij(k) = (2π)−3∫cij(x) exp(ikx)dx. Understanding

V as a perturbation, we obtain the corrections to the energy of

single-particle states. By perturbation theory, we can also calculate

the scattering matrix corresponding to the Hamiltonian H ′ and,

therefore, to the Hamiltonian H.

Hence, the semiclassical approximation, as it is described in this

section, prompts a reasonable choice of an initial approximation in

the perturbation theory. However, one can show that semiclassical

approximation leads to important qualitative conclusions. Let us

consider, for example, the Hamiltonian

H =1

2

∫π2(x)dx +

1

2

∫(∇φ)2dx +

∫U(φ(x))dx, (9.55)

where U(φ) = a1φ+a2φ2 +a3φ

3 +a4φ4. (This Hamiltonian describes

a Lorentz-invariant theory (see Section 12.3). The Hamiltonian H

leads to ultraviolet divergences that should be treated by means of

renormalization; we do not analyze these divergences here.) It is easy

to see that for the Hamiltonian (9.55), ν(φ) = U(φ). If U(φ) =

a2φ2 + a4φ

4, a2 > 0, a4 > 0, then the minimum of ν(φ) is achieved

at φ = 0. This means that in the analysis of the Hamiltonian (9.55),

we should use the perturbation theory where H0 is specified in the

Page 232: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 185

The Scattering Matrix for Translation-Invariant Hamiltonians 185

usual way as follows:

H0 =1

2

∫π2(x)dx +

1

2

∫(∇φ)2dx +

∫a2φ

2(x)dx. (9.56)

In general, it is unreasonable to specify H0 by (9.56) (this is clear

since the Hamiltonian (9.56), in the case a2 < 0, describes particles

with imaginary mass; it is impossible to construct its operator

realization). In order to make the correct choice for an initial

approximation, we should find

min ν(φ) = minU(φ) = min(a1φ+ a2φ2 + a3φ

3 + a4φ4).

If this minimum is achieved for φ = α, then, using the canonical

transformation, we get the Hamiltonian H ′ = H0 + V , where

H0 =1

2

∫π2(x)dx +

1

2

∫(∇ξ)2dx +

1

2µ2

∫ξ2(x)dx,

µ2 = U ′′(α).

Now, we can apply the perturbation theory to the Hamiltonian H ′

considering V as a perturbation. In the case when U(φ) = a2φ2 +

a4φ4, a2 < 0, a4 > 0, the minimum of the function U(φ) is achieved

at two points α = ±√

a22a4

(there are two different classical vacua).

Hence, the necessary canonical transformation and the Hamiltonian

H ′ can be constructed in two ways: H ′± = H0 + V±, where

V± = a4

∫ξ4(x)dx± 2

√2a2a4

∫ξ3(x)dx.

This allows us to conjecture that in the case at hand, one can

construct different operator realizations of the Hamiltonian H. In

fact, let us take an operator realization of the Hamiltonian H ′+assuming that V+ is a perturbation. This means that

lim~→0

⟨ξ(x, t)Φ,Φ

⟩= 0.

Page 233: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 186

186 Mathematical Foundations of Quantum Field Theory

Starting with the operator realization of the Hamiltonian H ′+, we get

an equivalent operator realization of the Hamiltonian H that gives

lim~→0

⟨φ(x, t)Φ,Φ

⟩=

√a2

2a4+ lim

~→0〈ξ(x, t)Φ,Φ〉

=

√a2

2a4> 0.

For an operator realization of the Hamiltonian H, constructed in a

similar way by means of H ′−, we get that

lim~→0〈φ(x, t)Φ,Φ〉 = −

√a2

2a4< 0,

and hence, we have constructed two essentially different operator

realizations.

Note that we have encountered an important phenomenon called

symmetry breaking. The initial Hamiltonian function is invariant

with respect to the transformation φ(x) → −φ(x); however, the

classical vacuum is not invariant with respect to this transformation

(it transforms one classical vacuum into another). A similar situation

arises in the quantum case: the Hamiltonian H is invariant with

respect to the transformation φ(x) → −φ(x) and therefore, from

one operator realization of the Hamiltonian H, we can get another

by replacing the operators φ(x, t) by the operators −φ(x, t). If

〈φ(x, t)Φ,Φ〉 6= 0, then this replacement gives an operator realization

that is not equivalent to the original.

Let us consider one more interesting example, the Hamiltonian

H =1

2

∑∫π2i (x)dx +

∫ (∑φ2i (x)− a2

)2dx. (9.57)

Here,

ν(φ1, . . . , φn) = (φ21 + · · ·+ φ2

n − a2)2

and hence, the minimum of the function ν is achieved at the points

(α1, . . . , αn) satisfying the condition∑α2i = a2 (the classical vacua

fill an (n−1)-dimensional sphere). The Hamiltonian (9.57), as well as

Page 234: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 187

The Scattering Matrix for Translation-Invariant Hamiltonians 187

the corresponding classical Hamiltonian functional, is invariant with

respect to the transformations

φ′i(x) =∑

aijφj(x), π′i(x) =∑

aij πj(x),

where aij ∈ O(n) (here, O(n) is the group of orthogonal matrices).

However, in the operator realization of the Hamiltonian H, we have

a smaller symmetry group — the group O(n − 1). To check this

fact, we note that the operator realization of the Hamiltonian H

corresponds to a classical vacuum and, for every classical vacuum, the

transformations g ∈ O(n) that sends this vacuum to itself (satisfying

the condition gα = α) constitute a subgroup isomorphic to O(n−1).

Then, transformations that do not belong to this subgroup send an

operator realization to another that is not equivalent to the original

operator realization.

In this example, we obtain a symmetry breaking of a continuous

symmetry group. In this situation, we necessarily have particles with

energy tending to zero when the momentum tends to zero (Goldstone

particles). It is easy to verify this statement when one can apply

the semiclassical considerations of the present section. Indeed, in the

initial approximation, the energy of the particles is determined by the

eigenvalues of the matrix cij(k). This matrix continuously depends

on k, and therefore, it is sufficient to analyze the matrix cij(0) that

coincides with the matrix of second derivatives ∂2ν∂φi∂φj

of the function

ν at the minimum point. This matrix is necessarily degenerate,

since in our assumption the minimum cannot be achieved at one

isolated point. Hence, we conclude that in the zeroth approximation,

the Goldstone particles exist. It is easy to check that taking into

account the higher approximations of the perturbation theory with

respect to V does not change this statement. In Lorentz-invariant

theory, the energy of a particle with momentum p is equal to√p2 +m2, where m is the mass of the particle, and hence Goldstone

particles have zero mass.

We can conclude that the Hamiltonian (9.54) describes particles

with energy having the form ~ωj(~−1p) when ~ → 0. However,

aside from these particles, that can be called elementary, the

Hamiltonian (9.54) also describes the particles with energies having

Page 235: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 188

188 Mathematical Foundations of Quantum Field Theory

a finite limit as ~ → 0. These particles correspond to the so-

called “particle-like” solutions of classical equations. The simplest

particle-like solutions are solitons. A solution πi(x, t), φi(x, t) of the

Hamiltonian equations is called a soliton if the function φi(x, t)

can be represented in the form si(x − vt) (we assume that the

energy of this solution is finite). (Here, we assume that the energy is

measured as a deviation from the classical vacuum, i.e. H(π, φ) = 0

if πi(x) ≡ 0, φi(x) = αi, where (α1, . . . , αn) is a classical vacuum. If

the condition H(π, φ) = 0 is not satisfied, then we should replace the

Hamiltonian functional (9.54) with the functional

H(π, φ) =1

2

∑i

∫π2i (x)dx +

∑m

∑i1,...,im

∫Vi1,...,im(x1, . . . ,xm)

× (φi1(x1) . . . φim(xm)− αi1 . . . αim)dmx,

which leads to the same equations of motion.)

A soliton with a zero velocity (v = 0) is a solution of the equations

of motion that does not depend on time. Note that it follows

from translation-invariance of the Hamiltonian functional (9.53)

that the momentum P =∑

i

∫πi∂φi∂x dx is an integral of motion

of the Hamilton equations; the soliton with zero velocity has zero

momentum. In the Lorentz-invariant case, we can obtain solitons

with arbitrary momentum (and with aribtrary velocity that is less

than the speed of light) from the solitons of zero velocity by means

of Lorentz transformation.

One can check that there exists a quantum particle that corre-

sponds to a stable soliton (the solution φi(x, t), πi(x, t) is called stable

if every solution φ′i(x, t), π′i(x, t) that is close to the original at t = 0

stays close to it at any other moment in time t). More precisely,

if for every three-dimensional vector p, we have a stable soliton

of the Hamilton equations, having the momentum p and energy

ε(p), then under certain conditions, we can prove that among the

particles described by Hamiltonian (9.54) are particles with energy

tending to ε(p) as ~ → 0 (in other words, in the space of operator

realizations of the Hamiltonian (9.54), we have vector generalized

functions Φ~(p) that are δ-normalized and obey the conditions

HΦ~(p) = E~(p)Φ~(p),PΦ~(p) = pΦ~(p), where limE~(p) = ε(p);

Page 236: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 189

The Scattering Matrix for Translation-Invariant Hamiltonians 189

for more, see Takhtadzhyan and Faddeev (1974), Faddeev (1975),

Tiupkin et al. (1975b), Polyakov (1974) and Tyupkin et al. (1975).

As an example, we can consider a classical system with the

Hamiltonian

H(π, φ) =1

2

∫π2(x)dx +

1

2

∫ (dφ

dx

)2

dx

+

∫(φ2(x)− a2)2dx,

where x ∈ E1. The Hamiltonian equations take the form

∂tφ(x, t) = π(x, t),

∂tφ(x, t) =

∂2

∂x2φ(x, t)− 4φ(x, t)(φ2(x, t)− a2).

If φ(x, t) = s(x), then the function s(x) satisfies the equation

d2s

dx2= 4s(s2 − a2). (9.58)

From the condition of finite energy, we can conclude that

limx→±∞ |s(x)| = a; using this relation and (9.58), we obtain, in

the case at hand, the solitons with zero velocity

s(x) = ±ath√

2ax,

they have energy µ = 4√

23 a3. The soliton with momentum p can be

obtained directly or by means of Lorentz transformation; the energy

of this soliton is equal to√

p2 + µ2. One can prove that there exist

quantum particles corresponding to these solitons.

Let us note in conclusion that one can use the semiclassical

approximation to analyze more general Hamiltonians (8.2). Namely,

we should introduce the complex functions ψ(x) = 1√2(φ(x) +

iπ(x)), ψ(x) = 1√2(φ(x)− iπ(x)) in place of the generalized momenta

π(x) and coordinates φ(x). An arbitrary Hamiltonian H(π, φ) can be

expressed in terms of ψ(x), ψ(x); let us assume that

H(ψ,ψ) =∑m,n

∫Hm,n(x1, . . . ,xm|y1, . . . ,yn)

× ψ(x1) . . . ψ(xm)ψ(y1) . . . ψ(yn)dmxdny.

Page 237: Mathematical Foundations of Quantum Field Theory

March 27, 2020 9:30 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch09 page 190

190 Mathematical Foundations of Quantum Field Theory

By the quantization of the functions ψ(x), ψ(x), we obtain the

operator generalized functions ψ+(x), ψ(x) obeying the commutation

relations

[ψ(x), (x′)] = [ψ+(x), ψ+(x′)] = 0,

[ψ(x), ψ+(x′)] = ~δ(x− x′),

and the Hamiltonian functionalH specifies the quantum Hamiltonian

H =∑m,n

∫Hm,n(x1, . . . ,xm|y1, . . . ,yn)

× ψ+(x1) . . . ψ+(xm)ψ(y1) . . . ψ(yn)dmxdny

(in the transition from the Hamiltonian functional H to the quantum

Hamiltonian H, we should answer the question about the ordering

of the operators ψ+, ψ in the expression of H; the choice we have

made corresponds to the so-called “Wick quantization”). If the

Hamiltonian functional H is translation-invariant, then using the

formulas

ψ+(x) = (2π)−3/2~1/2

∫exp(ikx)a+(k)dk,

ψ(x) = (2π)−3/2~1/2

∫exp(−ikx)a(k)dk

we can write down the Hamiltonian (9.35) in the form (8.2). This

allows us to apply the considerations of this section to the analysis

of the Hamiltonian (8.2).

Page 238: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 191

Chapter 10

Axiomatic Scattering Theory

10.1 Main assumptions and the construction of the

scattering matrix

Let us suppose that we have four commuting self-adjoint operators

H and P = (P1, P2, P3) defined on the Hilbert space H. The

operator H has the physical meaning of an energy operator and the

(vector) operator P has the meaning of the momentum operator.

The operator exp(iPa) is called the spatial shift operator or spatial

translation and the operator exp(−iHt) is the time shift operator

(time translation). Let us assume that the energy operator has the

unique ground state Φ and the state is invariant with respect to

space shifts and time shifts: HΦ = PΦ = 0. The vector Φ is called

the physical vacuum.

In the situation at hand, we can apply the definition of a particle

given in Section 5.3. We will formulate this definition in a slightly

different way.

Let us suppose that for every function f ∈ L2(E3) we have

a corresponding vector Φ(f) ∈ H that linearly depends on the

function f . This relationship can be understood as a linear operator

mapping from L2(E3) to H; however, we will prefer to view it as

a vector generalized function Φ(k), where the vector Φ(f) can be

written in the form Φ(f) =∫f(k)Φ(k)dk. The vector generalized

function Φ(k) is called a particle (a single-particle state), if for every

191

Page 239: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 192

192 Mathematical Foundations of Quantum Field Theory

pair of functions f, g ∈ L2(E3) we have the relations

HΦ(f) = Φ(hf),

PΦ(f) = Φ(pf),

〈Φ(f),Φ(g)〉 = 〈f, g〉 ,

where

(hf)(k) = ω(k)f(k),

(pf)(k) = kf(k).

The vector Φ(f) represents the state of a particle with the wave

function f and the function ω(k) has the meaning of the energy

of the single-particle state (dispersion law). The relations above

are equivalent to the relations (5.20)–(5.22). The last one implies

that the operator taking the function f to the vector Φ(f) is an

isometry; this is equivalent to the relation (5.22), which shows

that Φ(k) is δ-function normalized. It is possible that there are

several types of particles (several types of generalized functions Φ(k))

that satisfy (5.20)–(5.22). A system of particles Φ1(k), . . . ,Φs(k)

is called complete, if every pair of particles is orthogonal (i.e.

〈Φi(k),Φj(k′)〉 = 0 for i 6= j) and there does not exist another

particle that is orthogonal to all the particles in the system.

One can show that every particle Φ(k) can be decomposed into

the particles of a complete system (i.e. we can find functions fi(k) so

that Φ(k) =∑s

i=1 fi(k)Φi(k)).

Let us fix a complete particle system Φ1(k), . . . ,Φs(k) with the

dispersion laws ω1(k), . . . , ωs(k), assuming that the functions ωi(k)

are strongly convex. (We suppose for definiteness that we have a finite

number s of particles, even though the case of an infinite number of

particles is not much more difficult.) The single-particle subspace H1

of the space H will be defined as the smallest subspace that contains

all vectors of the form Φi(f) =∫f(k)Φi(k)dk. The multi-particle

subspace M will be defined as the orthogonal complement of the

direct sum H0 +H1 in H, where H0 is the one-dimensional subspace

generated by the physical vacuum Φ.

Page 240: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 193

Axiomatic Scattering Theory 193

The spaces H0,H1,M are clearly invariant with respect to the

operators H,P. The joint spectrum of the system of commuting

operators H,P in the space H1 (in the space M) will be called the

single-particle (correspondingly, the multi-particle) spectrum and we

will denote it by∑

1 (correspondingly,∑M). The set

∑1 clearly

coincides with the union of the sets∑(i)

1 , where∑(i)

1 denotes the set

of points of the form (k, ωi(k)).

Let us assume that the single-particle spectrum does not intersect

the multi-particle spectrum and that there exists δ > 0, such that

the spectrum of the operator H, with the exception of the point 0,

corresponding to the physical vacuum, is contained in the ray [δ,+∞).

We will later show how this assumption can be relaxed.

Let us consider, along with the spaceH, the asymptotic state space

Has, which we will define as the Fock space F (L2(E3×N)), where N

is a finite set whose elements are in one-to-one correspondence with

the complete system of particles we have assumed previously. Let us

suppose that the operator generalized functions a+i (k), ai(k) obey

CCR (see Section 3.2) and act on the space Has; these functions

can be considered as the creation and annihilation operators of

particles of the ith type with momentum k ∈ E3. We will define

the asymptotic Hamiltonian Has and the momentum operator Pas

by the formulas

Has =s∑i=1

∫ωi(k)a+

i (k)ai(k)dk,

Pas =

s∑i=1

∫ka+

i (k)ai(k)dk.

If A is an operator on the space H, then A(x, t) will denote the

operator exp(i(Ht − Px))A exp(−i(Ht − Px)). Bounded operators

A and B, acting on the space H, will be said to be asymptotically

commuting if for every natural number n we can find real numbers

C and r such that

‖[A,B(x, t)]‖ ≤ C 1 + |t|r

1 + |x|n.

Page 241: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 194

194 Mathematical Foundations of Quantum Field Theory

A family of bounded operators A on the space H that contains

the conjugate operator A∗ for every operator A in A will be

called asymptotically Abelian if two arbitrary operators in A are

asymptotically commuting.

An asymptotically Abelian family that contains the identity

operator will be called an asymptotically Abelian algebra (or asymp-

totically commutative algebra) if for every pair of operators A, B in

M, the operators

λA+ µB, AB,

∫f(x, t)A(x, t)dxdt

(here, λ, µ are complex numbers, x ∈ E3, −∞ < t <∞, and f(x, t)

is a function in the space S of smooth, fast-decaying functions) are

in M as well. It is easy to show that every asymptotically Abelian

family is contained in an asymptotically Abelian algebra. (The proof

is based on the fact that adding any of the listed operators along with

their adjoints to an asymptotically Abelian family preserves asymp-

totic commutativity.) Therefore, without loss of generality, we will

consider asymptotically Abelian algebras instead of asymptotically

Abelian families.

Let us fix an asymptotically Abelian algebra A of operators on the

space H; let us assume that the physical vacuum Φ is a cyclic vector

of the algebra A.

With the definitions and assumptions above, we will now show

how we can construct a scattering matrix (S-matrix) corresponding

to an asymptotically Abelian algebra A (we can then define a scat-

tering matrix of an asymptotically Abelian family as the scattering

matrix corresponding to the asymptotically Abelian algebra that

contains this family). For the sake of simplicity of definition, we will

assume that the full particle system consists of a single particle Φ(k).

Let us first make several preliminary definitions.

An operatorB is called smooth if it can be written in the formB =∫f(x, t)A(x, t)dxdt, where A ∈ A and the function f belongs to the

space S. (This name comes from the observation that the operator

B(x, t), where B is a smooth operator, is infinitely differentiable in x

Page 242: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 195

Axiomatic Scattering Theory 195

and t in the sense of differentiation in norm.) The smooth operator

B is called good, if (1) B∗Φ = 0; (2) there exists a function φ such

that BΦ =∫φ(k)Φ(k)dk.

Definition 10.1. An isometric operator S− (S+), transforming the

space Has into the space H, is called a Møller matrix if for every

collection of good operators B1, . . . , Bn and for every collection of

smooth functions with compact support f1(p), . . . , fn(p)

limt→ −∞(+∞)

B1(f1, t) · · ·Bn(fn, t)Φ = S −(+)a+(f1φ1) · · · a+(fnφn)θ.

(10.1)

[Here, θ is a vacuum vector in the Fock space Has, φj(k) =

〈BjΦ,Φ(k)〉, the functions fj(x|t) are determined by the relation

fj(x|t) =

∫exp(−iω(p)t+ ipx)fj(p)

dp

(2π)3, (10.2)

and the operators Bj(fj , t) are given by the formula

Bj(fj , t) =

∫fj(x|t)Bj(x, t)dx.

](10.3)

The vector B1(f1, t) · · ·Bn(fn, t)Φ, constructed with good oper-

ators B1, . . . , Bn and the smooth functions with compact support

f1, . . . , fn, will be denoted by Ψ(B1, . . . , Bn|f1, . . . , fn|t). The limit of

this vector as t→ ±∞ will be denoted by Ψ±(B1, . . . , Bn|f1, . . . , fn).

Definition 10.2. The operator S = S∗+S−, acting on the space Has,

is called the scattering matrix, corresponding to the asymptotically

Abelian algebra A.

Let us prove the following statement.

With the previous assumptions, the operators S− and S+ satisfy-

ing the conditions of Definition 10.1, exist, and are uniquely specified

Page 243: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 196

196 Mathematical Foundations of Quantum Field Theory

by these conditions. The following relations also hold :

S∓θ = Φ, S±a+(k)θ = Φ(k), (10.4)

HS∓ = S∓Has, PS∓ = S∓Pas. (10.5)

It follows from relation (10.5) that the scattering matrix S =

S∗+S− commutes with the Hamiltonian and the momentum operator

SHas = HasS, SPas = PasS,

and the relation (10.4) implies the stability of the vacuum and single-

particle states:

Sθ = θ,

Sa+(k)θ = a+(k)θ.

If the images of the operators S− and S+ coincide (S−Has = S+Has),

then the scattering matrix is unitary and can be written in the form

S = S−1+ S−.

If, in addition to the algebra A, we have a second asymptotically

Abelian algebra A′, for which the vector Φ is cyclic and the algebras

A and A′ asymptotically commute,1 then the Møller matrices S∓and the scattering matrix S, constructed with the algebra A, coincide

with the Møller matrix S′∓ and the scattering matrix S′, constructed

with the algebra A′.The proof of the listed statements is based on a sequence of

lemmas. Let us now formulate the necessary lemmas and then

prove the statements. The proofs of the lemmas are relegated to

Section 10.2.

Lemma 10.1. The set of vectors of the form BΦ, where B runs over

all good operators in the algebra A, is dense in the single-particle

space H1.

1The algebras A and A′ asymptotically commute if every operator in Aasymptotically commutes with every operator in A′.

Page 244: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 197

Axiomatic Scattering Theory 197

Lemma 10.2. Let ω(p) be a smooth, strongly convex function, f(p)

be a smooth function with compact support, and

f(x|t) =

∫exp(−iω(p)t+ ipx)f(p)

dp

(2π)3, (10.6)

then

supx|f(x|t)| ≤ C1|t|−3/2, (10.7)∫|f(x|t)|dx ≤ C2(|t|3/2 + 1). (10.8)

If U is a set of vectors of the form v(p) = ∇ω(p), where p ∈ suppf

and Uε is an ε-neighborhood of the set U , then for any n and ε, we

can find a constant D such that

|f(x|t)| ≤ D(1 + x2 + t2)−n (10.9)

whenever xt 6∈ Uε.

[The symbol supp f , as usual, denotes the closure of the set of

points where f(p) 6= 0.]

Before we formulate Lemma 10.3, we need to introduce the

concept of a truncated vacuum expectation value.

Let us define the vacuum expectation value 〈A〉 of the operator A

as the number 〈AΦ,Φ〉. The truncated vacuum expectation value

〈A1 . . . An〉T of the product of n operators A1, . . . , An is defined by

the following formulas for n = 1, 2, 3:

〈A1〉T = 〈A1〉,

〈A1A2〉T = 〈A1A2〉 − 〈A1〉T 〈A2〉T = 〈A1A2〉 − 〈A1〉 〈A2〉,

〈A1A2A3〉T = 〈A1A2A3〉 − 〈A1A2〉T 〈A3〉T − 〈A1〉T 〈A2A3〉T

− 〈A2〉T 〈A1A3〉T − 〈A1〉T 〈A2〉T 〈A3〉T .

For arbitrary n, the truncated vacuum expectation value is defined by

the recurrence relation

〈A1 . . . An〉 =

n∑k=1

∑ρ∈Rk

〈A(π1)〉T · · · 〈A(πk)〉T , (10.10)

Page 245: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 198

198 Mathematical Foundations of Quantum Field Theory

where Rk is the collection of all partitions of the set 1, . . . , n into

k subsets; π1, . . . , πk are the subsets that constitute the partition

ρ ∈ Rk; 〈A(π)〉T is a truncated vacuum expectation value of the

product of the operators Ai with indices in the set π (the operators

are ordered in ascending order by the index i).

Let us suppose that for every λ in the (finite or infinite) set Λ we

have a corresponding operator Aλ from the algebra A.

The expectation value

wn(λ1,x1, t1, . . . , λn,xn, tn) = 〈Aλ1(x1, t1) · · ·Aλn(xn, tn)〉

of the product Aλ1(x1, t1) · · ·Aλn(xn, tn) in the ground state Φ will

be called the (n-point) Wightman function (see Section 7.1). Let us

define the truncated Wightman function wTn (λ1,x1, t1, . . . , λn,xn, tn)

as the truncated vacuum expectation value of this product:

wTn (λ1,x1, t1, . . . , λn,xn, tn) = 〈Aλ1(x1, t1) · · ·Aλn(xn, tn)〉T .

In other words,

wn(λ1,x1, t1, . . . , λn,xn, tn) =

n∑k=1

∑ρ∈Rk

wTα1(π1) · · ·wTαk(πk).

(10.11)

Here, αr denotes the number of elements in the subset πr; wTαr(πr) is

a truncated Wightman function with the arguments λi1 , xi1 , ti1 , . . . ,

λiαr ,xiαr , tiαr , where i1, . . . , iαr ∈ πr (the order of the arguments is in

order of increasing indices i). For n = 1, 2, 3, the relations connecting

Wightman functions and truncated Wightman functions have the

form

w1(λ1,x1, t1) = wT1 (λ1,x1, t1),

w2(λ1,x1, t1, λ2,x2, t2) = wT2 (λ1,x1, t1, λ2,x2, t2)

+wT1 (λ1,x1, t1)wT1 (λ2,x2, t2),

Page 246: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 199

Axiomatic Scattering Theory 199

w3(λ1,x1, t1, λ2,x2, t2, λ3,x3, t3)

= wT3 (λ1,x1, t1, λ2,x2, t2, λ3,x3, t3)

+wT1 (λ1,x1, t1)wT2 (λ2,x2, t2, λ3,x3, t3)

+wT1 (λ2,x2, t2)wT2 (λ1,x1, t1, λ3,x3, t3)

+wT1 (λ3,x3, t3)wT2 (λ1,x1, t1, λ2,x2, t2)

+wT1 (λ1,x1, t1)wT1 (λ2,x2, t2)w

T1 (λ3,x3, t3).

Lemma 10.3. If all the operators Aλi are smooth, then the truncated

Wightman function

wTn (λ1,x1, t1, . . . , λn,xn, tn) = 〈Aλ1(x1, t1) . . . Aλn(xn, tn)〉T ,

for fixed t1, . . . , tn, tends to zero faster than any power function of

D = max1≤i,j≤n |xi − xj | as D →∞.

Lemma 10.4. Let f0, . . . , fn be finite smooth functions and let

A0, . . . , An be smooth operators belonging to the algebra A. Then for

n ≥ 1, we have the inequality

| 〈A0(f0, t) . . . An(fn, t)〉T | ≤ C(1 + |t|)−32

(n−1), (10.12)

where Aj(fj , t) =∫fj(x|t)Aj(x|t)dx. This inequality remains valid if

in the expression for 〈A0(f0, t) . . . An(fn, t)〉T some of the operators

Aj(fj , t) are replaced by their time derivatives Aj(fj , t) and the

operators (Aj(fj , t))∗ are replaced by (Aj(fj , t))

∗.

Let us now turn to the derivation of the statements made in the

beginning of this chapter from the lemmas we have formulated. First,

let us derive the relations (10.4). The first of these is self-evident; to

prove the second relation, let us note that

Ψ(B|f |t) = B(f, t)Φ

=

∫f(x|t)B(x, t)Φdx

=

∫f(x|t) exp(itH − iPx)BΦdx

Page 247: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 200

200 Mathematical Foundations of Quantum Field Theory

=

∫f(x|t) exp(itω(p)− ipx)φ(p)Φ(p)dxdp

=

∫f(p)φ(p)Φ(p)dp, (10.13)

i.e. Ψ(B|f |t) does not depend on t:

d

dtΨ(B|f |t) =

d

dtB(f, t)Φ = 0. (10.14)

This implies that

Ψ±(B, f) =

∫f(p)φ(p)Φ(p)dp,

i.e.

S∓(

∫f(p)φ(p)dpθ) =

∫f(p)φ(p)Φ(p)dp. (10.15)

It follows from Lemma 10.1 that vectors of the form

BΦ =

∫φB(p)Φ(p)dp,

where B runs over a set of good operators, constitute a dense set

in the space of single-particle states. This implies that the functions

φB(p), corresponding to good operators B, constitute a dense set

in L2(E3), since the correspondence between the function φ(p) and

the vector∫φ(p)Φ(p)dp is an isometry between the space L2(E3)

and the space of single-particle states. It is clear that the functions

f(p)φB(p), where f(p) is a smooth function, also form a dense set

in L2(E3). The necessary relation then follows from (10.15).

Let us now show that the limit in the relation (10.1) exists. To

show this, we will prove the inequality⟨dΨ

dt,dΨ

dt

⟩≤ C|t|−3, (10.16)

Page 248: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 201

Axiomatic Scattering Theory 201

where Ψ(t) = Ψ(B1, . . . , Bn|f1, . . . , fn|t). This inequality implies the

convergence of the limit in (10.1) since

‖Ψ(t2)−Ψ(t1)‖ ≤∫ t2

t1

∥∥∥∥dΨ

dt

∥∥∥∥ dt ≤ ∫ t2

t1

C|t|−3/2dt→ 0,

for t1 ≤ t2 and t1 → +∞ or t2 → −∞.

In order to prove the inequality, let us write the expansion⟨dΨ

dt,dΨ

dt

⟩=∑i,j

〈B1(f1, t) . . . Bi(fi, t)

. . . Bn(fn, t)Φ, B1(f1, t) . . . Bj(fj , t) . . . Bn(fn, t)Φ〉

=∑i,j

〈(Bn(fn, t))∗ . . . (Bj(fj , t))

∗ . . . (B1(f1, t))∗

×B1(f1, t) . . . Bi(fi, t) . . . Bn(fn, t)Φ,Φ〉

as a sum of the products of truncated vacuum expectation values

using the relation (10.10).

Every factor in an arbitrary term of this expansion has the form

Ik,l(t) = 〈(Bi1(fi1 , t))∗ . . . (Bik(fik , t))

× Bj1(fj1 , t) . . . Bjl(fjl , t)〉T ,

where the dots above the operators denote differentiation with

respect to time, which can enter in one of the first k operators and in

one of the last l operators. We can suppose without loss of generality

that k ≥ 1, l ≥ 1 (otherwise, Ik,l = 0 by the relation B∗Φ = 0). If

each factor in the given term has k = l = 1, then the term is equal

to zero because one of the factors either has the form

〈(Bi(fi, t))∗Bj(fj , t)〉T = 〈(Bi(fi, t))∗Bj(fj , t)Φ,Φ〉

or the form

〈(Bi(fi, t))∗Bj(fj , t)〉T = 〈(Bi(fi, t))∗Bj(fj , t)Φ,Φ〉

and both expressions are zero by the relation (10.14). Hence, all

non-zero terms in the sum above contain either a factor Ik,l with

Page 249: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 202

202 Mathematical Foundations of Quantum Field Theory

k + l ≥ 4 or at least two factors Ik,l with k + l = 3 (a term where

all factors except one have k + l = 2 and the remaining factor has

k+ l = 3 is not possible since the total number of operators is even).

The inequality

|Ik,l(t)| < C|t|−32

(k+l−2),

follows from Lemma 10.4. In particular, we have

|I1,1(t)| < C,

|Ik,l(t)| < Ct−3/2,

for k + l = 3, and

|Ik,l(t)| < Ct−3

for k + l ≥ 4. From these inequalities and the statements shown

above, we obtain the inequality (10.16) for⟨dΨdt ,

dΨdt

⟩and therefore

the existence of the limit in (10.1) follows.

Since the limit in (10.1) exists, the relation (10.1) specifies the

operators S∓ on some subspace L of the space Has. However, it is

not clear if the action of the operators S∓ is uniquely defined, i.e.

whether the vector S∓ξ = Ψ∓(B1, . . . , Bn|f1, . . . , fn) remains the

same if we write ξ in two different forms

ξ = a+(φ1f1) . . . a+(φnfn)θ,

where the functions fi are smooth and finite and the functions φicorrespond to good operators.

Therefore, let us now prove that the relation (10.1) defines S∓uniquely and, furthermore, this operator can be uniquely extended

to a linear isometric operator defined on the whole space Has. To

show this, we will use the following general statement.

Let us suppose that on the total2 subset L of the Hilbert space

H, we have a multi-valued isometric mapping α taking values in H′(i.e. to every point ξ ∈ L corresponds a set of points α(ξ) ⊂ H′ in

such a way that for any selection of points ξ1, ξ2 ∈ L, x1, x2 ∈ H′satisfying x1 ∈ α(ξ1), x2 ∈ α(ξ2) we have 〈x1, x2〉 = 〈ξ1, ξ2〉). Then

2A subset is called total if the set of its linear combinations is dense in H.

Page 250: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 203

Axiomatic Scattering Theory 203

the mapping α is in fact single-valued (i.e. every set α(ξ) consists of

a single point) and can be uniquely extended to a linear isometric

operator defined on the whole space H.

To prove this statement, let us first extend the mapping α by

linearity to the multi-valued mapping α, defined on the set L of linear

combinations of points in the set L. More precisely, we will assume

that x ∈ α(ξ) if we can write ξ and x in the forms ξ =∑n

i=1 λiξi,

x =∑n

i=1 λixi with xi ∈ α(ξi). It is easy to see that the mapping

α inherits linearity: if ξ = λ1ξ1 + λ2ξ2, x1 ∈ α(ξ1), x2 ∈ α(ξ2), then

λ1x1 + λ2x2 ∈ α(ξ).

The mapping α is also a multi-valued isometric mapping (if x(1) ∈α(ξ(1)), x(2) ∈ α(ξ(2)) then ξ(1) =

∑λiξ

(1)i , x(1) =

∑λix

(1)i , ξ(2) =∑

µjξ(2)j , x(2) =

∑µjx

(2)j , where x

(1)i = α(ξ

(1)i ), x

(2)j = α(ξ

(2)j ), and

therefore⟨x(1), x(2)

⟩=∑

λiµj

⟨x

(1)i , x

(2)j

⟩=∑

λiµj

⟨ξ

(1)i , ξ

(2)j

⟩=⟨ξ(1), ξ(2)

⟩).

Let us now prove that the mapping α is in fact single-valued. Let

us suppose that x1 ∈ α(ξ), x2 ∈ α(ξ). It follows from the isometry of

the mapping that

〈x1, x1〉 = 〈x2, x2〉 = 〈x1, x2〉 = 〈ξ, ξ〉 .

From this equation, we obtain that x1 = x2.

Therefore, α is a well-defined isometric linear mapping defined on

the set L. Since the set L is total and the set L is dense in H, we can

continuously extend α to an isometric linear operator defined on all

of H.

To apply the statement above to the proof of the Møller matrix

properties, we should check that the set L of vectors of the form

a+(φ1f1) . . . a+(φnfn)θ, on which the operators S∓ are defined

by (10.1), is total in the space Has, and furthermore, we should

check that for good operators B1, . . . , Bn, B′1, . . . , B

′m and for smooth

functions with compact support f1, . . . , fn, f′1, . . . , f

′m we have⟨

Ψ∓(B1, . . . , Bn|f1, . . . , fn),Ψ∓(B′1, . . . , B′m|f ′1, . . . , f ′m)

⟩=⟨a+(φ1f1) · · · a+(φnfn)θ, a+(φ

′1f′1) · · · a+(φ

′nf′n)θ⟩. (10.17)

Page 251: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 204

204 Mathematical Foundations of Quantum Field Theory

The totality of the set L follows from the fact that functions of

the form φB(k)f(k), where f(k) is a smooth function with compact

support and B is a good operator, are dense in L2(E3) (as we have

noted above, this fact follows from Lemma 10.1).

To prove relation (10.17), let us represent the left side of this

relation in the form⟨Ψ∓(B1, . . . , Bn|f1, . . . , fn),Ψ∓(B′1, . . . , B

′m|f ′1, . . . , f ′m)

⟩= lim

t→∓∞

⟨Ψ(B1, . . . , Bn|f1, . . . , fn|t),Ψ(B′1, . . . , B

′m|f ′1, . . . , f ′m|t)

⟩= lim

t→∓∞

⟨(B′m(f ′m, t))

∗ · · · (B′1(f ′1, t))∗B1(f1, t) · · ·Bn(fn, t)

⟩(10.18)

and then expand the expression under the limit sign in terms

of truncated Wightman functions. Every term of the resulting

expansion consists of a product of factors of the form

Rk,l(t) =⟨(B′i1(fi1 , t))

∗ · · · (B′ik(fik , t))∗Bj1(fj1 , t) · · ·Bjl(fjl , t)

⟩T.

It follows from Lemma 10.4 that for k + l ≥ 3 the factors Rk,l(t)

converge to zero as t→ ±∞. On the other hand, the factors Rk,l are

equal to zero if k = 0 or l = 0, by the relation BΦ = 0. Therefore,

the terms that differ from zero in the limit t → ∓∞ are the ones

where all the factors have the form R1,1(t). It is easy to check that

R1,1(t) =⟨(B′i(f

′i , t))

∗Bj(fj , t)⟩T

=⟨(B′i(f

′i , t)

∗Bj(fj , t))⟩

=⟨Bj(fj , t)Φ, B

′i(f′i , t)Φ

⟩=

⟨∫fj(k)φj(k)Φ(k)dk,

∫f ′i(k)φ′i(k)Φ(k)dk

⟩=⟨fjφj , f

′iφ′i

⟩=⟨a+(f jφj)θ, a

+(f′iφ′i)θ⟩

and that R1,1 in fact does not depend on t [in the calculations we

have used the formula (10.13)].

The statements proven above imply that the expression (10.18)

is equal to δnm∑∏n

s=1 〈φsfs, φ′isf ′is〉 [the sum is taken over all

permutations (i1, . . . , in)]. This justifies equation (10.17).

Page 252: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 205

Axiomatic Scattering Theory 205

Thus, we have proven that relation (10.1) specifies S∓ as an

isometric mapping on a total set and we have shown that one can

uniquely extend this mapping to an isometric operator defined on

the space Has.

To prove relation (10.5), note that the equation

exp(−iHτ)B(f, t) exp(iHτ)

=

∫exp(−iHτ)f(x|t)B(x, t) exp(iHτ)dx

=

∫f(x|t)B(x, t− τ)dx

=

∫ (∫exp(−iω(p)t+ ipx)f(p)

dp

(2π)3

)B(x, t− τ)dx

=

∫ (∫exp(−iω(p)(t− τ) + ipx)f(p) exp(−iω(p)τ)

dp

(2π)3

)×B(x, t− τ)dx,

holds; this implies

exp(−iHτ)B(f, t) exp(iHτ) = B(f τ , t− τ), (10.19)

[f τ denotes the function f τ (p) = f(p) exp(−iω(p)τ)]. It follows from

equation (10.19) that

exp(−iHτ)Ψ(B1, . . . , Bn|f1, . . . , fn|t)

= Ψ(B1, . . . , Bn|f τ1 , . . . , f τn |t− τ). (10.20)

Taking the limit t→ ∓∞ in (10.20) and using the fact that

exp(−iHasτ)a+(φ1f1) . . . a+(φnfn)θ

= a+(φ1fτ1) . . . a+(φnf

τn)θ,

we obtain

exp(−iHτ)S∓ξ = S∓ exp(−iHasτ)ξ

Page 253: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 206

206 Mathematical Foundations of Quantum Field Theory

for any vector ξ ∈ L. By the totality of the set L, we have

exp(−iHτ)S∓ = S∓ exp(−iHasτ)

and, therefore, the first relation in (10.5) follows (the second can be

proven by the same method, but is even easier).

To prove the final statement, let us note that Møller matrices

do not change when we replace an asymptotically Abelian algebra

by a larger one. Indeed, if the algebra A, satisfying the necessary

conditions, is contained in an asymptotically Abelian algebra A, then

in the construction of the operators S∓ corresponding to the algebra

A, we can use the good operators that belong to the algebraA; hence,

it is clear that Møller matrices corresponding to the algebrasA and Aare identical. If the algebras A and A′ are asymptotically commuting,

then the family A∪A′ is asymptotically Abelian and can be included

in an asymptotically Abelian algebra A. Since A ⊂ A and A′ ⊂ A,

the Møller matrices S∓ and S′∓, constructed with the algebras A and

A′, coincide with the Møller matrices constructed with the algebra

A, and therefore coincide with each other. The coincidence of Møller

matrices clearly entails the coincidence of the scattering matrices

constructed with the algebras A and A′.In conclusion, let us note that the Møller matrices and the

scattering matrix do not depend on the choice of complete particle

system (recall that there exists an isomorphism between any two

choices of complete particle system for the space Has; therefore,

it makes sense to discuss the coincidence of scattering matrices

constructed under different particle systems). To prove this, note

that for any particle system with a single particle Φ(k), any other

particle can be written in the form

Φ′(k) = exp(iα(k))Φ(k),

where α(k) is a real-valued function.

The operators a+(k), a(k) in the space Has, corresponding to

the particle Φ(k), are related to the operators a′+(k), a′(k) on Has,

corresponding to Φ′(k), by the equations

a′(k) = exp(−iα(k))a(k),

a′+(k) = exp(iα(k))a+(k).

Page 254: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 207

Axiomatic Scattering Theory 207

It is easy to check, using these relations, that for any operator Bj

a′+(f jφ′j) = a+(f jφj), (10.21)

where

φ′j(k) =⟨BjΦ,Φ

′(k)⟩, φj(k) = 〈BjΦ,Φ(k)〉 .

Applying (10.21) and noting that the left part of (10.1) does not

depend on the choice of the function Φ(k) representing the particle,

we obtain that the Møller matrices S± do not depend on the choice

of the function Φ(k).

10.2 Proof of lemmas

Proof of Lemma 10.1. Let λ(ω,p) be a function of four parame-

ters with support in the set ∆. We will consider the operator λ(H,P)

as a function of the commuting self-adjoint operators H,P. It is easy

to check that if the set ∆ does not intersect the spectrum of the

operators (H,P), then the operator λ(H,P) = 0, and if the set ∆

does not contain the origin and points belonging to the multi-particle

spectrum∑M, then the set of values of the operator λ(H,P) belongs

to the single-particle subspace.

Using this remark, we can construct a good operator by the

following method.

Let us consider a function α(t,x) ∈ S(E4) with the property that

its Fourier transform

α(ω,p) =

∫α(t,x) exp(i(ωt− px))dxdt

has support that does not intersect the multi-particle spectrum and

the half-space ω ≤ 0. Then the operator

B =

∫α(t,x)A(x, t)dxdt,

where A ∈ A, is good.

Page 255: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 208

208 Mathematical Foundations of Quantum Field Theory

Indeed, the operator B is smooth. The vector

BΦ =

∫α(t,x)A(x, t)Φdxdt

=

∫α(t,x) exp(iHt− iPx)A exp(−iHt+ iPx)Φdxdt

=

∫α(t,x) exp(i(Ht−Px))AΦdxdt = α(H,P)AΦ

belongs to the single-particle subspace, by the remark above. The

vector

B∗Φ =

∫α(t, x) exp(i(Ht−Px))A∗Φdxdt

= α(−H,−P)A∗Φ = 0,

since the support of the function α(−ω,−p) does not intersect the

spectrum of the operators H, P.

Let us now show that vectors BΦ, where B is a good operator,

constructed by the method above, are dense in the single-particle

subspace.

To prove this, let us consider the vector Φ(λ) =∫λ(k)Φ(k)dk,

where λ(k) is a smooth function with compact support, in H1.

We will construct a sequence of good operators Bn such that

BnΦ → Φ(λ) (this is enough to prove the statement above since

smooth functions with compact support λ(k) are dense in L2(E3)

and therefore the corresponding vectors Φ(λ) =∫λ(k)Φ(k)dk are

dense in H1).

By the cyclicity of the vacuum vector there exists a sequence of

operators An ∈ A so that AnΦ → Φ(λ). The necessary sequence of

good operators can now be constructed by setting Bn =∫α(t,x)

An(x, t)dxdt, where

α(t,x) =

∫α(ω,p) exp[−i(ωt− px)]

dωdp

(2π)4,

the properties of the supports of the functions α ensure that the

operators Bn are good, and the function α satisfies the condition

Page 256: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 209

Axiomatic Scattering Theory 209

α(ω(k),k) = 1 for all k, belonging to the support of the function

λ(k). Indeed, we have

limn→∞

BnΦ = limn→∞

α(H,P)AnΦ

= α(H,P)Φ(λ)

=

∫λ(k)α(H,P)Φ(k)dk

=

∫λ(k)α(ω(k),k)Φ(k)dk

=

∫λ(k)Φ(k)dk = Φ(λ).

This completes the proof of Lemma 10.1.

Before we prove Lemma 10.2, we will conduct some non-rigorous

but intuitive discussion. The integral (10.6) for large t can be

approximated by the method of stationary phase. The stationary

point p0 of the phase ω(p)t − px is determined by the equation

v(p0)t = x; denoting by w(ξ) the solution of the equation v(p) = ξ,

we can write the stationary point in the form p0 = w(xt ). Applying

the method of stationary phase to f(x|t) for large t, we obtain

f(x|t) ≈ Ct−3/2f(w(x

t

))exp

(−iω

(w(x

t

))t+ iw

(x

t

)),

where

C = (2πi)−3/2| det γik|−1/2,

γik =∂2

∂pi∂pkω(p)|p=w(x

t).

Hence, f(x|t), for large t, decays as t−3/2. If xt 6∈ Uε, then

f(w(xt )) = 0 and the stationary point is not contained in the support

of f(p); in this case, we should expect the function f(x|t) to be

very small. Unfortunately, the simple considerations here cannot be

applied to estimate the function f(x|t) uniformly in x; we will provide

accurate but more cumbersome proofs for these inequalities.

Proof of Lemma 10.2. Let us first estimate f(x|t) when xt 6∈ Uε.

Page 257: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 210

210 Mathematical Foundations of Quantum Field Theory

We will use the formula∫exp(−iσ(p))χ(p)dp =

∫exp(−iσ(p))(Lnχ)(p)dp, (10.22)

where χ(p) is a smooth function with compact support, (Lχ)(p) =

−div(u(p)χ(p)); u(p) = ∇σ(p)|∇σ(p)|2 (see, for example, Fedoryuk (1971)).

In the case at hand, we have

σ(p) = ω(p)t− px, χ(p) = f(p),

u(p) =v(p)t− x|v(p)t− x|2

=1

v(p)− xt

|v(p)− xt |2

.

If xt 6∈ Uε, it is easy to check that

supp∈suppχ

|D(α)u(p)| ≤ C

|t|

(here, D(α) = ∂|α|

∂pα11 p

α22 p

α33

, |α| = α1 + α2 + α3, and C, here and in

the following, denotes a quantity independent of x and t, though it

may depend on other parameters, for example, here it depends on α

and ε). The inequality

supp∈suppχ

|(Lnχ)(p)| ≤ C|t|−n (10.23)

follows. The formula (10.22) and the inequality (10.23) lead to

the inequality |f(x|t)| ≤ C1+|t|n , where x

t 6∈ Uε (n is an arbitrary

number).3 To prove the inequality (10.9), we should consider the

cases |x| ≤ at and |x| > at, where a = 2 supp∈supp f |v(p)|, separately.

In the first case, the inequality (10.9) follows from (10.23); in the

second case, we can use the fact that

supp∈supp f|x|>at

|D(α)u(p)| ≤ C

|x|.

3In the proof, we have assumed that the function χ(p) is smooth. One can relaxthis condition assuming that this function is m times continuously differentiable.Then we have the inequality (10.23) for n = m.

Page 258: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 211

Axiomatic Scattering Theory 211

To prove inequality (10.7), let us take a smooth function µ(x)

equal to zero for |x| ≤ ν1 and equal to one for |x| ≥ ν2.

Let us split the integral representing the function f(x|t) in two

as follows:

I1 =

∫exp(−iω(p)t+ ipx)µ

(tρ(v(p)− x

t

))f(p)dp; (10.24)

I2 =

∫exp(−iω(p)t+ ipx)

(1− µ

(tρ(v(p)− x

t

)))f(p)dp,

(10.25)

where the number ρ is in the interval 512 < ρ < 1

2 . The integral I1 can

be estimated with the formula (10.22). Let us introduce the notation

‖φ‖n = supp∈Γ|α|≤n

|D(α)φ (p)|,

where Γ denotes the set of points p such that |v(p) − xt | ≥ ν1|t|−ρ

(outside of Γ the integrand in the integral I1 is zero). It is easy to

check that ‖u‖n ≤ C|t|−1+(n+1)ρ.

Furthermore, we have

‖Lχ‖n ≤ C‖uχ‖n+1 ≤ C∑

α+β=n+1

‖u‖α‖χ‖β,

which implies that

‖Lrχ‖n ≤ C∑

α1+···+αr+β=n+r

‖u‖α1 . . . ‖u‖αr‖χ‖β

≤ C∑β

|t|−r+(n+2r−β)ρ‖χ‖β.

To estimate the integral I1, we should introduce

χ(p) = µ(tρ(v(p)− x

t

))into formula (10.22). Then, ‖χ‖β ≤ C|t|βρ implies

|I1| ≤ C‖Lrχ‖0 ≤∑β

C|t|−r+(2r−β)ρ‖χ‖β ≤ C|t|−r+2rρ.

When r ≥ 1.5−2ρ+1 , we can show that |I1| ≤ C|t|−3/2.

Page 259: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 212

212 Mathematical Foundations of Quantum Field Theory

In integral I2, let us replace ω(p) with

ω1(p) = ω(p0) + v(p0)(p− p0) +∑i=1

∑j=1

1

2

∂2ω(p0)

∂pi∂pj

× (pi − p0i)(pj − p0j),

where p0 is any point with |v(p0)− xt | ≤ ν2|t|−ρ. We will show that

the error ∆ resulting from this change does not exceed C|t|−3/2. To

show this, let us use the formula w = v(p)− xt to make a substitution

in the integral

∆ =

∫(exp(−iω(p)t+ ipx)− exp(−iω1(p)t+ ipx))

×(

1− µ(t−ρ(v(p)− x

t

)))f(p)dp.

Since the function ω(p) is strongly convex and the function f(p) has

compact support, it follows that the Jacobian of this transformation

is bounded above and below by positive constants.

By the properties of the functions µ, we can assume that the

integral ∆ is taken over the domain Γ1, where |w| ≤ ν2|t|−ρ; the

volume of this domain equals C|t|−3ρ. Over the domain Γ1, we can

use the estimate

|(exp(−iω(p)t+ ipx)− exp(−iω1(p)t+ ipx))f(p)|

≤ C|t||p− p0|3 ≤ C1|t||w|3 ≤ C2|t|−3ρ+1,

and we then obtain that |δ| ≤ C|t|1−6ρ ≤ C|t|1−6· 512 = C|t|−3/2.

In the remaining part of the proof, we need to estimate the

expression

I2 −∆ =

∫exp(−iω1(p)t+ ipx)f(p)dp

−∫

exp(−iω1(p) + ipx)µ(t−ρ(v(p)− x

t

))dp.

The second term in this expression can be estimated in the same way

as the integral I1. The first term can be estimated if we rewrite it

Page 260: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 213

Axiomatic Scattering Theory 213

using the Fourier transform, in the form∫G(x− y, t)f(y)dy,

where

f(x) =

∫f(p) exp(ipx)

dp

(2π)3,

G(x, t) =

(2π

it

)3/2

|det γjk|−1/2 exp

i

2

∑j,k

λjkxjxk

,γjk =

∂2ω(p0)

∂pj∂pk,

λjk is the inverse matrix of the matrix γjk.

The inequality (10.8) follows from the inequality (10.7) in the case

x ∈ Uεt and from the inequality (10.9) in the case x 6∈ Uεt.

Remark 10.1. With some extra conditions on the behavior of the

function ω(p) at infinity, we can prove Lemma 10.2 for any smooth,

rapidly decaying function f(p).

Proof of Lemma 10.3. Let us begin by proving the simple but

important identity that connects the truncated vacuum expectation

value

〈AB(t)〉T = 〈AB(t)Φ,Φ〉 − 〈AΦ,Φ〉 〈BΦ,Φ〉

with the vacuum expectation value of the commutator [A,B(t)]. To

prove this, let us consider the smooth function h(ω), satisfying the

conditions h(ω) = 0 for ω ≤ 0 and h(ω) = 1 for ω ≥ δ [the number

δ > 0 is chosen such that all points of the spectrum of the operator

H, with the exception of 0, lie on the ray (δ,+∞)]. Let us note that

h(H) = 1− P0,

h(−H) = 0,(10.26)

where P0 is a projection operator on the vacuum vector Φ (these

equations can be checked by using the isomorphism between the

spaces H and L2(M), which sends the operator H to an operator

Page 261: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 214

214 Mathematical Foundations of Quantum Field Theory

of multiplication by a function). Let us assign to a function f ∈ Sthe function

fh(t) =

∫h(ω) exp(iω(τ − t))f(τ)

dωdτ

2π.

It is easy to check that the function fh(t) also belongs to the space S(the operator mapping f to the function fh is in fact a composition of

the Fourier transform over t, the operator of multiplication by h(ω),

and the inverse Fourier transform; all these operators transform Sinto itself). If the function f only depends on t and not on the other

variables, then the function fh is constructed by means of the Fourier

transform over t (other variables are considered as parameters). If

the function f(t, x1, . . . , xn) belongs to the space S(En+1), then the

function fh belongs to it as well.

Let us show that for any function f ∈ S, we have∫f(t) 〈AB(t)〉T dt =

∫fh(t) 〈[A,B(t)]〉 dt. (10.27)

Indeed, we have∫fh(t) 〈AB(t)〉 dt =

∫fh(t) 〈A exp(iHt)B〉 dt

= 〈Ah(H)f(H)B〉 = 〈A(1− P0)f(H)B〉

=

∫f(t) 〈A(1− P0) exp(iHt)B〉 dt =

∫f(t) 〈AB(t)〉T dt,

(10.28)∫fh(t) 〈B(t)A〉 dt =

∫fh(t) 〈B exp(−iHt)A〉 dt

= 〈Bh(−H)f(−H)A〉 = 0. (10.29)

Here, we have used the relation (10.26) and the equations∫fh(t) exp(iωt)dt = h(ω)f(ω),∫f(t) exp(iωt)dt = f(ω).

Combining (10.28) and (10.29), we obtain (10.27).

Page 262: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 215

Axiomatic Scattering Theory 215

Let us now prove the lemma for the case n = 2. We will suppose

that A1, A2, C ∈ A, α ∈ S, A2 =∫α(x, t)C(x, t)dxdt and introduce

the notation D =∫αh(x, t)C(x, t)dxdt. From the relation (10.27), it

is easy to see that

〈A1A2(x, t)〉T = 〈[A1, D(x, t)]〉 .

Since D ∈ A and all operators in A asymptotically commute, it

follows that

|wT2 (x1, t1,x2, t2)| = | 〈A1(x1, t1)A2(x2, t2)〉T |

= | 〈[A1, D(x2 − x1, t2 − t1)]〉 | ≤ C(1 + |t2 − t1|r)1 + |x2 − x1|n

.

Now that we have proved the lemma for the case n = 2, let us

give the proof in the case of arbitrary Wightman functions. Let us

consider the function

wn(x1, t1, . . . ,xn, tn) = 〈A1(x, t1) . . . An(xn, tn)〉

and the two functions

w(1)k (xi1 , ti1 , . . . ,xik , tik) = 〈Ai1(xi1 , ti1) . . . Aik(xik , tik)〉 ,

w(1)l (xj1 , tj1 , . . . ,xjl , tjl) = 〈Aj1(xj1 , tj1) . . . Ajl(xjl , tjl)〉 ,

obtained by splitting the operatorsA1, . . . , An, entering the definition

of the function wn, into two groups Ai1 , . . . , Aik and Aj1 , . . . , Ajk(here, k + l = n, the indices i1, . . . , ik and j1, . . . , jl are in ascending

order, and the operators Ai are assumed to be smooth).

The set of indices i1, . . . , ik, the indices of the operators in the

first group, will be denoted by K; the set of indices j1, . . . , jl of the

second group of operators will be denoted by L; clearly K ∪ L =

1, 2, . . . , n.It can be shown that when all the points xi, where i ∈ K, are

far from the points xj , where j ∈ L, the following approximation

holds:

wn(x1, t1, . . . ,xn, tn)

≈ w(1)k (xi1 , ti1 , . . . ,xik , tik)w

(2)l (xj1 , tj1 , . . . ,xjl , tjl).

Page 263: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 216

216 Mathematical Foundations of Quantum Field Theory

More precisely, for fixed t1, . . . , tn, the absolute value of the

function

wn(x1, t1, . . . ,xn, tn)− w(1)k (xi1 , ti1 , . . . ,xik , tik)w

(2)l

× (xj1 , tj1 , . . . ,xjl , tjl)

does not exceed Cd−p. Here, d = mini∈K,j∈L |xi−xj |; p is an arbitrary

number; C is a constant depending on p, but not on x1, . . . ,xn.

(This property of Wightman functions is often called asymptotic

factorization or the cluster property.)

To prove the asymptotic factorization property, we will first

consider the case when the set K consists of the points (1, 2, . . . , k)

and the set L consists of the points (k + 1, . . . , n). In this case, the

difference we aim to estimate can be written in the form

〈RU〉T = 〈RU〉 − 〈R〉 〈U〉 ,where

R = A1(x1, t1) . . . Ak(xk, tk),

U = Ak+1(xk+1, tk+1) . . . An(xn, tn).

Recall that the operators A1, . . . , An are smooth and can therefore

be written in the form

Ai =

∫fi(x, t)Bi(x, t)dxdt,

where fi ∈ S, Bi ∈ A. Noting that

U =

∫fk+1(ξ1, τ1) . . . fn(ξn−k, τn−k)Bk+1(xk+1 + ξ1, tk+1 + τ1)

. . . Bn(xn + ξn−k, tn + τn−k)dn−kξdn−kτ

=

∫α(t, ξ1 − xk+1, . . . , ξn−k − xn, σ1, . . . , σn−k−1)

× exp(iHt)Bk+1(ξ1, tk+1)Bk+2(ξ2, tk+2 + σ1)

. . . Bn(ξn−k, tn + σn−k−1) exp(−iHt)dtdn−kξdn−k−1σ,

where

α(t, ξ1, . . . , ξn−k, σ1, . . . , σn−k−1)

= fk+1(ξ1, t)fk+2(ξ2, t+ σ1) · · · fn(ξn−k, t+ σn−k−1),

Page 264: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 217

Axiomatic Scattering Theory 217

we can apply the relation (10.27). This leads to

〈R,U〉T =

∫αh(t, ξ1 − xn+1, . . . , ξn−k − xn, σ1, . . . , σn−k−1)

×〈[R,Bk+1(ξ1, t+ tk+1)Bk+2(ξ2, t+ tk+2 + σ1)

. . . Bn(ξn−k, t+ tn + σn−k−1)]〉 dtdn−kξdn−k−1σ. (10.30)

The commutator on the right-hand side of equation (10.30) can be

written in the form∑1≤i≤k

k+1≤j≤n

Cij [Ai(xi, ti), Bj(ξj−k, t+ tj + σj−k−1)]Dij ,

where Cij , Dij are operators whose norms are bounded by a constant

not depending on xi, ti, ξi, σi. Using this fact, asymptotic commu-

tativity, and formula (10.30), we obtain the asymptotic factorization

property for the caseK = 1, . . . , k, L = k+1, . . . , n. We can show

that the general case can be reduced to this special case. Suppose the

set K consists of the points (i1, . . . , ik) and the set L consists of the

points (j1, . . . , jl), where i1 < · · · < ik, j1 < · · · < jl. If the quantity

d = mini∈K,j∈L |xi − xj | is large, we can use the approximation

〈A1(x1, t1) . . . An(xn, tn)〉 ≈ 〈Ai1(xi1 , ti1) . . . Aik(xik , tik)

×Aj1(xj1 , tj1) . . . Ajl(xjl , tjl)〉 , (10.31)

whose error does not exceed Cd−m, where m is an arbitrary number;

C is a constant depending on m,n, t1, . . . , tn. We can justify this

approximation by using the relation

〈EAi(xi, ti)Aj(xj , tj)F 〉 ≈ 〈EAj(xj , tj)Ai(xi, ti)F 〉 ,

for i ∈ K, j ∈ L, that we have proved already (here, E, F are

compositions of the operators Aα(xα, tα), where α ∈ K ∪ L).

From the relation (10.31) and the asymptotic factorization prop-

erty, we can now obtain the asymptotic factorization property in the

general case by the same method as in the special case shown earlier.

Let us now show how to derive Lemma 10.3 from the asymptotic

factorization property of Wightman functions. We will suppose

Page 265: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 218

218 Mathematical Foundations of Quantum Field Theory

that the lemma holds for functions wTn−1 and then show it for

functions wTn . However, we will first prove the following geometric

lemma.

Let N be a set consisting of the points x1, . . . ,xn in Euclidean

space and let D be the diameter of N (i.e. D = maxxi,xj∈N |xi−xj |).Then, we can partition N into two subsets P and Q in such a way

so that ρ(P,Q) = minxi∈P,xj∈Q |xi − xj | ≥ D2n .

We will prove this fact by induction on n. Let D = |xα−xβ|. Let

us delete from N the point xγ which does not coincide with xα or

with xβ, obtaining as a result the set N ′, consisting of n − 1 points

and having diameter D. By the inductive hypothesis, we can split the

set N ′ into two sets P ′ and Q′ so that ρ(P ′, Q′) = minxi∈P,xj∈Q |xi−xj | ≥ D

2n−1 . By the triangle inequality, the point xγ cannot satisfy

the inequalities ρ(P ′,xγ) = minxi∈P ′ |xi − xγ | < D2n and ρ(Q′,xγ) =

minxi∈Q′ |xi − xγ | < D2n simultaneously; for concreteness, suppose

that ρ(P ′,xγ) ≥ D2n . Then, it is clear that the necessary partition

can be constructed by taking P to be P ′ and taking Q to be

Q′ ∪ xγ.Let us now suppose that Lemma 10.3 holds for functions wTk with

k < n. To prove the lemma for functions wTn , we need to estimate

wTn (x1, t1, . . . ,xn, tn) under the condition max1≤i≤j≤n |xi − xj | = D

(with the times t1, . . . , tn fixed). Using the geometric lemma we just

proved, let us partition the set 1, . . . , n into the subsets K and L

such that mini∈K,j∈L |xi − xj | ≥ D2n . To estimate the functions wTn ,

let us note that in the right-hand side of the relation (10.11) (the

recurrence relation defining the functions wTn ) all factors have the

form wαr(πr) with the set πr containing both elements of K and L.

By induction, we can obtain an estimate of the form CD−m, where

m is arbitrary and C depends on m. It therefore follows that every

term on the right-hand side of (10.11) contains at least one factor of

the form just described and therefore admits an estimate of the form

CD−m, since the factors wTαi(πi) do not exceed a constant, depending

only on ‖A1‖, . . . , ‖An‖ and the number n (this statement becomes

clear if we note that

| 〈A1(x1, t1) . . . An(xn, tn)〉 | ≤ ‖A1‖ . . . ‖An‖).

Page 266: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 219

Axiomatic Scattering Theory 219

Hence, with error not exceeding CD−m, we can write the

approximation

wn(x1, t1, . . . ,xn, tn) ≈ wTn (x1, t1, . . . ,xn, tn)

+n∑k=2

∑ρ

wTα1(π1) . . . wTαk(πk), (10.32)

where the sum is taken over partitions ρ of the set 1, . . . , n such

that every set πj is completely contained in either K or L.

It is easy to check that the sum in (10.32) equals the product

wk(xi1 , ti1 , . . . ,xik , tik)wl(xj1 , tj1 , . . . ,xjl , tjl), where (i1, . . . , ik) are

elements of K and (j1, . . . , jl) are elements of L.

Hence, we see that

wTn (x1, t1, . . . ,xn, tn)

≈ wn(x1, t1, . . . ,xn, tn)

−wk(xi1 , ti1 , . . . ,xik , tik)wl(xj1 , tj1 , . . . ,xjl , tjl).

To finish the proof of Lemma 10.3, we apply the asymptotic

factorization property of Wightman functions.

Remark 10.2. Because the operators Ai are smooth, the Wightman

functions wn(x1, t1, . . . ,xn, tn) and the truncated Wightman func-

tions wTn (x1, t1, . . . ,xn, tn) are smooth, and hence the derivatives of

these functions can be viewed as Wightman and truncated Wightman

functions constructed with different smooth operators. (This is clear

if we note that for a smooth operator A =∫f(τ,x)B(x, τ)dξdτ ,

where f ∈ S(E4), the following relation holds:

D(α)A(x, t) = D(α)

∫f(τ − t, ξ − x)B(ξ, τ)dξdτ

=

∫g(τ − t, ξ − x)B(ξ, τ)dξdτ = C(x, t),

where D(α) = ∂|α|

∂tα0∂xα11 ∂x

α22 ∂x

α33

is a differential operator; g(t,x) =

(−1)|α|D(α)f(t,x); C is a smooth operator defined by C =∫g(τ, ξ)B(ξ, τ)dξdτ ; operator derivatives are understood as norm

derivatives.) Using this fact and the statement of Lemma 10.3,

Page 267: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 220

220 Mathematical Foundations of Quantum Field Theory

we can say that for fixed t1, . . . , tn and fixed xn, the function

wTn (x1, t1, . . . ,xn, tn), constructed with smooth operators, belongs to

the space S(E3(n−1)).

The requirement that the operators A1, . . . , An be smooth in

Lemma 10.3 can be relaxed. However, if we completely remove this

requirement, then we have to weaken the statement of the lemma.

Namely, one can prove that without the smoothness condition on the

operators A1, . . . , An, the function

ν(x1, t1, . . . ,xn, tn)

=

∫γ(x1 − ξ1, t1 − τ1, . . . ,xn − ξn, tn − τn)

×wTn (ξ1, τ1, . . . , ξn, τn)dnξdnτ,

where γ ∈ S(E4n), for fixed t1, . . . , tn and fixed xn, belongs to the

space S(E3(n−1)).

Proof of Lemma 10.4. To prove this lemma, we need to estimate

the quantity

I = 〈A0(f0, t) . . . An(fn, t)〉T

=

∫f0(x0|t) . . . fn(xn|t) 〈A0(x0, t) . . . An(xn, t)〉T dx0 . . . dxn.

(10.33)

Let us change variables in integral (10.33) to the new variables

x0, ξ1, . . . , ξn, where ξj = xj − x0, and apply the estimate

| 〈A0(x0, t) . . . An(xn, t)〉T | ≤C

1 + ‖ξ‖m

=C

1 + (ξ21 + · · ·+ ξ2

n)m/2,

following from Lemma 10.3 (here, m is arbitrary and C depends on

m). Applying this estimate, we obtain

|I| ≤∫|f0(x0|t)|

(C

1 + |t|3/2

)n C

1 + ‖ξ‖mdx0d

≤ const(1 + |t|)−32n

∫|f0(x0|t)|dx0

≤ const(1 + |t|)−32

(n−1). (10.34)

Page 268: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 221

Axiomatic Scattering Theory 221

In the derivation above, we have used the inequalities |fj(x|t)| ≤C(1 + |t|3/2)−1, for j ≥ 1, and

∫|f0(x|t)|dx ≤ C(1 + |t|)3/2, which

follow from Lemma 10.2; the number m is chosen to be large enough

so that the integral∫

(1 + ‖ξ‖)−mdnξ converges.

The inequality (10.34) gives the necessary estimate of the quan-

tity I. To finish the proof of Lemma 10.4, we need an analogous

estimate in the case when in the expression 〈A1(f1, t) . . . An(fn, t)〉T

some of the operators Aj(fj , t) are replaced by the operators

Aj(fj , t) = ddtAj(fj , t).

Let us note that

Aj(fj , t) =d

dt

∫fj(x|t)Aj(x, t)dx

=

∫ (d

dtfj(x|t)

)Aj(x, t)dx +

∫fj(x|t)

d

dtAj(x, t)dx

=

∫gj(x|t)Aj(x, t)dx +

∫fj(x|t)Kj(x, t)dx,

where we have used the notation

K = i[H,A],

gj(x|t) =d

dtfj(x|t)

=

∫exp(−iω(p)t+ ipx)gj(p)

dp

(2π)3,

gj(p) = −iω(p)fj(p).

Hence, the operator Aj(fj , t) can be written in the form

Aj(fj , t) = Aj(gj , t) +Kj(fj , t), (10.35)

where gj is a smooth function with compact support; Kj is a smooth

operator. [The smoothness of Kj follows if we note that if the

operator Aj is represented in the form Aj =∫φj(τ, ξ)Bj(ξ, τ)dξdτ ,

the operator Kj is equal to −∫

( ddτ φj(τ, ξ))Bj(ξ, τ)dξdτ .]

By formula (10.35), the proof of the lemma in the case when

some of the operators Aj(fj , t) are replaced by the operators Aj(fj , t)

clearly reduces to the case already considered. The proof of the

lemma in the case when some of the operators Aj(fj , t) are replaced

Page 269: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 222

222 Mathematical Foundations of Quantum Field Theory

with the operators (Aj(fj , t))∗ or (Aj(fj , t))

∗ requires no new

ideas.

10.3 Asymptotic fields (in- and out-operators)

In Section 10.1, we defined the Møller matrices S± corresponding to

the asymptotically Abelian algebra A. With the Møller matrices, one

can define the in- and out-operators (asymptotic fields) a+in

out

(k) and

a inout

(k) by the relations

S−a+(k) = a+

in(k)S−,

S−a(k) = ain(k)S−,

S+a+(k) = a+

out(k)S+,

S+a+(k) = aout(k)S+.

These equations clearly define a+in(k), ain(k) as operator gener-

alized functions on the space Hin = S−Has and a+out(k), aout(k) as

operator generalized functions on the space Hout = S+Has.

In this section, we will show that in- and out-operators describe

the asymptotic behavior of the operators A(x, t), where A ∈ A, as

t→ ±∞ (this justifies the name asymptotic fields).

Before we begin precise formulations, let us introduce the follow-

ing definition.

Let us assign to every function f ∈ S(E3) the set U(f) ∈ E3

consisting of points of the form ∂ω(k)∂k , where k ∈ supp f (i.e. k

belongs to the support of f).

A family of functions f1, . . . , fn will be called non-overlapping if

the sets U(fj) do not intersect pairwise (i.e. the sets U(fi) and U(fj)

do not have common points if i 6= j).4

4Under the assumption of strong convexity on the functions ω(k), the familyf1, . . . , fn will be non-overlapping if and only if the supports of these functionssuppfj are pairwise non-intersecting. Therefore, in the case at hand, the definitionof non-overlapping family can be simplified. However, in the case when we havemultiple particles, the definition in the main text is necessary.

Page 270: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 223

Axiomatic Scattering Theory 223

The following remark elucidates the physical meaning of the new

notion.

If f1, . . . , fn is a non-overlapping family of functions, then the

dynamics on the space Has as t → ±∞ transform the vector

a+(f1) . . . a+(fn)θ into a set of far apart particles. More precisely,

we have

exp(−iHast)a+(f1) . . . a+(fn)θ = a+(f t1) . . . a+(f tn)θ

= (2π)32n

∫f1(x1|t) . . . fn(xn|t)a+(x1) . . . a+(xn)dx1 . . . dxnθ,

where

f tj (k) = fj(k) exp(−iω(k)t),

fj(x|t) =

∫f tj (k) exp(ikx)

dk

(2π)3

=

∫fj(k) exp(−iω(k)t+ ikx)

dk

(2π)3.

It follows from Lemma 10.2 that the function fj(x|t), for large t, is

small outside the set tUε(fj), where the symbol Uε(fj) denotes an

ε-neighborhood of the set U(fj). Under the assumed conditions, for

small enough ε and large enough t, the sets tUε(fj) are far apart.

Let us consider vectors of the form Ψ−(B1, . . . , Bn|f1, . . . , fn)

[correspondingly, of the form Ψ+(B1, . . . , Bn|f1, . . . , fn)], where

B1, . . . , Bn are good operators and f1, . . . , fn are functions with

compact support. We will call them non-overlapping in-vectors (out-

vectors) if f1, . . . , fn is a non-overlapping family of functions. The

set of linear combinations of non-overlapping in-vectors (out-vectors)

will be denoted by Din (correspondingly, Dout); Din and Dout are

linear subspaces that are dense in Hin and Hout, respectively.

Let us now prove a theorem that describes the asymptotic

behavior of operators A(x, t), where A ∈ A, in terms of in- and

out-operators. For concreteness, we will formulate the theorem in

the case t→ −∞.

Page 271: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 224

224 Mathematical Foundations of Quantum Field Theory

If A1, . . . , An are good operators, Ψ1 ∈ Din, then

limt→−∞

A1(f1, t, σ1) . . . An(fn, t, σn)Ψ1

= ain(φ1f1, σ1) . . . ain(φnfn, σn)Ψ1. (10.36)

If A1, . . . , An are smooth operators in the algebra A, satisfying the

condition 〈AjΦ,Φ〉 = 0, Ψ1 ∈ Din, Ψ2 ∈ Din, then

limt→−∞

〈A1(f1, t, σ1) . . . An(fn, t, σn)Ψ1,Ψ2〉

=⟨ain(φ1f1, σ1) . . . ain(φnfn, σn)Ψ1,Ψ2

⟩(10.37)

under the condition that the σj, corresponding to smooth operators

Aj that are not good, coincide.

In (10.36) and (10.37), the functions fj are smooth and have

compact support. The functions φj are given by the relation φj(k) =

〈AjΦ,Φ(k)〉 and A(f, t, σ) are defined by the formulas

A(f, t, 1) = A(f, t) =

∫f(x|t)A(x, t)dx,

A(f, t,−1) = A∗(f, t) =

∫f(x|t)A∗(x, t)dx.

(In other words, for arbitrary smooth operators A1, . . . , An, we have

weak convergence of the operators A1(f1, t, σ1) . . . An(fn, t, σn) to the

operators ain(φ1f1, σ1) . . . ain(φnfn, σn) in the limit as t → −∞ on

the setDin under the condition that σ1 = · · · = σn; for good operators

A1, . . . , An, one can prove that the convergence is strong on Din and

the condition σi = σj is not necessary.)

Let us begin the proof of the theorem.

Lemma 10.5. Let A1, . . . , An ∈ A, f1, . . . , fn be smooth functions

with compact support and σ1, . . . , σn = ±1. If for at least one pair of

indices i, j, where 1 ≤ i ≤ j ≤ n, the sets U(fi) and U(fj) do not

intersect and the σi = σj, then for any m

|〈A1(f1, t, σ1) . . . An(fn, t, σn)〉T | ≤ C(1 + |t|m)−1.

Page 272: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 225

Axiomatic Scattering Theory 225

This statement holds if we replace some of the operators

Ai(fi, t, σi) with their time derivatives ddtAi(fi, t, σi) in the expression

〈A1(f1, t, σ1) . . . An(fn, t, σn)〉T .

Proof. Let us write the expression

〈A1(f1, t, σ1) . . . An(fn, t, σn)〉T (10.38)

in the form∫f1(p1, σ1) exp(−iσ1ω(p1)t) . . . fn(pn, σn)

× exp(−iσnω(pn)t)wTn (p1, σ1, . . . ,pn, σn)dnp.

[Here, we have introduced the notation

wTn (p1, σ1, . . . ,pn, σn)

=

∫exp

(i∑

pjxj

)〈A1(x1, t, σ1) . . . An(xn, t, σn)〉T dnx,

A(x, t, 1) = A(x, t), A(x, t,−1) = A∗(x, t).]

It follows from Lemma 10.3 in Section 10.1 that the function

wTn has the form wTn (p1, σ1, . . . ,pn, σn) = νn(p2, . . . ,pn,

σ1, . . . , σn)δ(p1 + · · ·+ pn), where νn is a smooth function [one can

check that νn ∈ S(E3(n−1))].

Let us assume for concreteness that σ1 = σ2 and U(f1) does not

intersect U(f2). It then becomes convenient to write (10.38) in the

form ∫exp(−iΩ(p2, . . . ,pn)t)χ(p2, . . . ,pn)dp2 . . . dpn,

where χ ∈ S(E3(n−1));

Ω(p2, . . . ,pn) = σ1ω(−p2 − · · · − pn)

+σ2ω(p2) + σ3ω(p3) + · · ·+ σnω(pn).

Page 273: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 226

226 Mathematical Foundations of Quantum Field Theory

Let us note that

∂Ω

∂p2= −σ1

∂ω

∂p

∣∣∣∣p=−p2−···−pn

+ σ2∂ω

∂p

∣∣∣∣p=p2

= σ2

(∂ω

∂p

∣∣∣∣p=p2

− ∂ω

∂p

∣∣∣∣p=p1

),

from which it is clear that

∂Ω

∂p26= 0

on the support of the function χ.

We can therefore use formula (10.22), assuming that all variables,

except for p2, are fixed. In the case at hand, we have

u(p) =1

t

(∂Ω

∂p2

) ∣∣∣∣ ∂Ω

∂p2

∣∣∣∣−2

,

and therefore

|(Lrχ)(p)| ≤ C|t|−r. (10.39)

The inequality (10.39) gives us the inequality we seek.

The proof in the case when the operators Ai(fi, t, σi) are replaced

by their derivatives is similar to the proof of Lemma 10.4.

Lemma 10.6. If f1, . . . , fn are non-overlapping smooth functions

with compact support and B1, . . . , Bn are good operators, then for

all m

limt→−∞

tm‖Ψ−(B1, . . . , Bn|f1, . . . , fn)

−Ψ(B1, . . . , Bn|f1, . . . , fn|t)‖ = 0.

Proof. It is enough to show that for any m, we have∣∣∣∣⟨ d

dtΨ(B1, . . . , Bn|f1, . . . , fn|t),

d

dtΨ(B1, . . . , Bn|f1, . . . , fn|t)

⟩∣∣∣∣≤ C(1 + |t|m)−1.

Page 274: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 227

Axiomatic Scattering Theory 227

This inequality can be obtained in the same way as the inequality

with m = 3 in Section 10.1, by using the just-proved Lemma 10.5

instead of Lemma 10.4.

Lemma 10.7. The relations (10.36) and (10.37) hold in the case of

Ψ1 = Φ.

Proof. To prove this lemma, we first note that the norm of the

vector ξ(t), defined by the formula

ξ(t) = A1(f1, t, σ1) . . . An(fn, t, σn)Φ,

is bounded above by a constant not depending on t. Indeed,

〈ξ(t), ξ(t)〉 can be expanded in terms of truncated Wightman

functions, and by Lemma 10.4 from Section 10.1, all the resulting

truncated Wightman functions are bounded. To prove the lemma, it

is enough to consider the case when the vector Ψ2 has the form

Ψ2 = Ψ−(B1, . . . , Bm|g1, . . . , gm) = limt→−∞

Ψ2(t),

where

Ψ2(t) = Ψ(B1, . . . , Bm|g1, . . . , gm|t).

It follows from the boundedness of ‖ξ(t)‖ and ‖Ψ2(t)‖ that

limt→−∞

〈ξ(t),Ψ2〉 = limt→−∞

〈ξ(t),Ψ2(t)〉.

The inner product

〈ξ(t),Ψ2(t)〉

= 〈A1(f1, t, σ1) . . . An(fn, t, σn)Φ, B1(g1, t) . . . Bm(gm, t)Φ〉

can be expanded in terms of truncated Wightman functions. By

Lemma 10.4 of Section 10.1, as t→ −∞, the products that differ from

zero contain two-point truncated Wightman functions. It therefore

follows that

limt→+∞

〈ξ(t),Ψ2(t)〉

=⟨ain(φ1f1, σ1) . . . ain(φnfn, σn)Φ, a+

in(ψ1, g1) . . . a+in(ψmgm)Φ

⟩=⟨ain(φ1f1, σ1) . . . ain(φn, fn, σn)Φ,Ψ−(B1, . . . , Bm|g1, . . . , gm)

⟩(10.40)

Page 275: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 228

228 Mathematical Foundations of Quantum Field Theory

(here, the symbol ψi denotes a generalized function defined by the

formula BiΦ =∫ψi(k)Φ(k)dk).

In the derivation of (10.40), we should use the formula

limt→∞〈Ai(fi, t, σi)Aj(fj , t, σj)〉T

=⟨ain(φif i, σi)ain(φjf j , σj)

⟩,

which holds if at least one of the operators Ai, Aj is good or is adjoint

to a good operator or, without this assumption, if we have σi = σj (in

particular, this formula can be applied in the case Aj = Bj , fj = gj ;

then φj = ψj).

The relation (10.40) implies equation (10.37) for Ψ1 = Φ.

In the case when A1, . . . , An are good operators, we can, as in

Section 10.1, prove that the vector ξ(t) has a limit as t→ −∞.

It follows from the statements we have already proved that

the vector ξ − η, where ξ = limt→−∞ ξ(t), η = ain(φ1f1, σ1)

. . . ain(φnfn, σn)Φ, is orthogonal to the space Hin. On the other

hand, a simple calculation, based on expansion in terms of truncated

Wightman functions, shows that 〈ξ, ξ〉 = 〈η, η, 〉. Since η ∈ Hin, this

implies that ξ = η. Therefore, relation (10.36) holds for Ψ1 = Φ,

which finishes the proof of Lemma 10.6 in this case.

Let us now prove the relations (10.36) and (10.37) in the general

case.

Let Ψ1 = Ψ−(E1, . . . , Er|h1, . . . , hr), where E1, . . . , Er are good

operators, h1, . . . , hr are non-overlapping functions, and Ψ1(t) =

Ψ(E1, . . . , Er|h1, . . . , hr|t).It follows from (10.8) that ‖Ai(f, t, σ)‖ ≤ C(1 + |t|3/2) and

therefore from Lemma 10.3 it follows that for t→∞ we have

‖A1(f1, t, σ1) . . . An(fn, t, σn)(Ψ1 −Ψ1(t))‖ → 0.

Therefore, we have

limt→−∞

A1(f1, t, σ1) . . . An(fn, t, σn)Ψ1

= limt→−∞

A1(f1, t, σ1) . . . An(fn, t, σn)Ψ1(t)

= limt→−∞

A1(f1, t, σ1) . . . An(fn, t, σn)E1(h1, t, 1) . . . Er(hr, t, 1)Φ.

To finish the proof, it remains to apply Lemma 10.3.

Page 276: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 229

Axiomatic Scattering Theory 229

Remark 10.3. Relation (10.37) holds in the case when Ψ2 is an

arbitrary vector in Hin.

10.4 Dressing operators

In this section, we will prove that the scattering matrix construction

provided in Section 10.1 is equivalent to a different construction

which is more closely connected to the physical picture of scattering.

Let us begin by introducing the notion of a dressing operator.

An operator D, acting from space Has to the space H, will be

called in-dressing (out-dressing), if

S− = slimt→−∞

exp(iHt)D exp(−iHast), (10.41)

S+ = slimt→+∞

exp(iHt)D exp(−iHast). (10.42)

If both relations are satisfied, then the operator D will be called

a dressing operator.

It is useful to note that it follows from relation (10.4) that both

in-dressing and out-dressing operators satisfy the conditions

Dθ = Φ, (10.43)

Da+(k)θ = Φ(k). (10.44)

However, these conditions are not sufficient for an operator to be

a dressing operator. Note that, in particular, the S− operator is an

in-dressing operator and the S+ operator is an out-dressing operator.

Let us now give sufficient conditions for an operator D to be in-

dressing. Namely, let us prove the following theorem.

Theorem 10.1. Let us suppose that (a) the operator D has norm 1,

(b) for all non-overlapping families of functions with compact support

f1(k), . . . , fn(k) and for all smooth operators A1, . . . , An ∈ A, we

have, for t < 0,

〈Da+(ft1) . . . a+(f

tn)θ,A1(g1) . . . An(gn)Φ〉

≈∑π∈P

⟨Φ(f t1), Ai1(gi1)Φ

⟩. . .⟨Φ(f tn), Ain(gin)Φ

⟩(10.45)

Page 277: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 230

230 Mathematical Foundations of Quantum Field Theory

with error not exceeding cνn(g)|t|−N (here, f ti (k) = fi(k) exp

(−iω(k)t), P is the set of permutations π = (i1, . . . , in) of indices

(1, . . . , n), ν(g) = maxi(sup |gi(x)| +∫|gi(x)|dx); gi ∈ S(E3); N is

an arbitrary number; C is a constant depending on N but not on the

functions gi). Then the operator D is in-dressing.

The condition for an operator to be out-dressing can be obtained

from the theorem above by replacing t < 0 with t > 0.

To prove the above theorem, let us consider the vector x = S−y,

where x = Ψ−(B1, . . . , Bn|f1, . . . , fn); y = a+(φ1f1) . . . a+(φnfn)θ;

B1, . . . , Bn are good operators and f1, . . . , fn are non-overlapping

functions, and

x(t) = Ψ−(B1, . . . , Bn|f t1, . . . , f tn)

= S−a+(φ1f

t1) . . . a+(φnf

tn)θ.

It is clear that

exp(−iHast)y = a+(φ1ft1) . . . a+(φnf

tn)θ (10.46)

and therefore, by the relation (10.5), we have

x(t) = exp(−iHt)x. (10.47)

For large t, the vector x(t) negligibly differs from the vector ξ(t) =

Ψ(B1, . . . , Bn|f t1, . . . , f tn|0) = B1(f t1, 0) . . . Bn(f tn, 0)Φ (to prove this

statement, note that for t→ ±∞ we have

‖x(t)− ξ(t)‖ = ‖ exp(iHt)(x(t)− ξ(t))‖

= ‖Ψ−(B1, . . . , Bn|f1, . . . , fn)

−Ψ(B1, . . . , Bn|f1, . . . , fn|t)‖ → 0). (10.48)

On the other hand, by relation (10.46), we have that

limt→−∞

〈D exp(−iHast)y, ξ(t)〉

= limt→−∞

〈Da+(φ1ft1) . . . a+(φnf

tn)θ,B1(f t1, 0) . . . Bn(f tn, 0)Φ〉

= limt→−∞

∑π∈P

⟨Φ(φ1f

t1), Bi1(f ti1 , 0)Φ

⟩. . .⟨Φ(φnf

tn), Bin(f tin , 0)Φ

⟩= 〈y, y〉 . (10.49)

Page 278: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 231

Axiomatic Scattering Theory 231

Recalling that ‖D exp(−iHast)y‖ ≤ ‖y‖ and that limt→−∞ ‖ξ(t) −x(t)‖ = 0, we obtain from (10.49) the relation

limt→−∞

〈D exp(−iHast)y, x(t)〉 = 〈y, y〉 , (10.50)

from which it follows by (10.47) that

limt→−∞

〈exp(iHt)D exp(−iHast)y, S−y〉

= limt→−∞

〈exp(iHt)D exp(−iHast)y, x〉

= limt→−∞

〈D exp(−iHast)y, x(t)〉 = 〈y, y〉 = 〈S−y, S−y〉.

Here, we apply the following simple lemma.

If lim 〈βi, η〉 = 〈η, η〉 and ‖βt‖ ≤ ‖η‖, then limβt exists and equals

η (to prove this statement, we write 〈βt, η〉 in the form

〈βt, η〉 = ‖βt‖ · ‖η‖ cosφt,

where φt is the angle between βt and η, and note that it follows from

the inequalities ‖βt‖ ≤ ‖η‖, | cosφt| ≤ 1 that ‖βt‖ cosφt converges to

‖η‖ only in the case when lim ‖βt‖ = ‖η‖, lim cosφt = 1).

Applying the lemma to the vectors βt = exp(iHt)D exp(−iHast)y,

η = S−y, we see that limt→−∞ exp(iHt)D exp(−iHast)y exists and

equals S−y.

Since the vectors y = a+(φ1f1) . . . a+(φnfn)θ, which we have

shown to satisfy the relation

limt→−∞

exp(iHt)D exp(−iHast)y = S−y,

constitute a total set in the space Has, it is clear that the relation

holds for a dense set of vectors and hence for any vector y ∈ Has (see

Appendix A.5).

We have thus obtained sufficient conditions for an operator D to

be in-dressing and out-dressing. These conditions can be relaxed by

changing the requirement that ‖D‖ = 1 with

limt→−∞

(t→+∞)

‖Da+(ft1) . . . a+(f

tn)θ‖ = ‖a+(f

t1) . . . a+(f

tn)θ‖ (10.51)

(the proof remains the same).

Page 279: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 232

232 Mathematical Foundations of Quantum Field Theory

The above statements also give conditions guaranteeing that

the operator D is dressing (since the operator is dressing if it is

simultaneously in- and out-dressing). These conditions can be easily

formulated in a slightly different form. Namely, we have the following

theorem.

Theorem 10.2. Let us suppose the operator D has norm 1 and

for all functions f1(x), . . . , fn(x) ∈ S(E3), whose supports are far

from each other, and for all smooth operators A1, . . . , An ∈ A, the

following approximation holds:⟨Da+(f1) . . . a+(fn)θ,A1(g1) . . . An(gn)Φ

⟩≈∑π∈P〈Φ(f1), Ai1(gi1)Φ〉 . . . 〈Φ(fn), Ain(gin)Φ〉 (10.52)

with error not exceeding Cνn(f)νn(g)d−N (here, fi(x) =∫fi(k)

exp(ikx) dk(2π)3

; the functions gi ∈ S(E3); d is the minimal distance

between the supports of the functions fi(x); N is an arbitrary number;

C is a constant depending on N but not depending on the func-

tions f1, . . . , fn, g1, . . . , gn; the sum is taken over all permutations

π = (i1, . . . , in)). Then the operator D is a dressing operator.

Indeed, relation (10.52) implies relation (10.45) for t ≤ 0 as

well as for t ≥ 0. To show this, note that by Lemma 10.2 from

Section 10.1 the “essential supports” of the functions fi(x|t) =∫exp(ikx)f ti (k) dk

(2π)3are far from each other in the limit t → ±∞.

(By this lemma, the function fi(x|t) is small outside the set tU(i)ε ,

where U(i)ε is an ε-neighborhood of the set U (i) consisting of points of

the form ∂ω(k)∂k , where k runs over suppfi. Hence, we can say that the

set tU(i)ε is the “essential support” of the function fi(x|t), recognizing

that the values of the function outside of this set are very small.)

To provide more rigorous reasoning, let us introduce smooth

functions hi(x) ∈ S, equal to 1 on the set U(i)ε and 0 outside

the set U(i)2ε . Then, the functions tri = hi(tx)fi(x|t) will have

far away supports as t → ±∞ and the quantity 〈Da+(tr1) . . .

Page 280: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 233

Axiomatic Scattering Theory 233

a+(trn)θ,A1(g1) . . . An(gn)Φ〉 can be approximated with the rela-

tion (10.52). From the other side, we have

a+(ft1) . . . a+(f

tn)θ ≈ a+(tr1) . . . a+(trn)θ

by Lemma 10.2 of Section 10.1. Using these remarks, we obtain that

relation (10.52) implies relation (10.45).

Let us note that the condition ‖D‖ = 1 can be relaxed here as

well, by assuming that either the condition (10.51) or the condition

‖Da+(f1) . . . a+(fn)θ‖ ≈ ‖a+(f1) . . . a+(fn)θ‖ (10.53)

holds. (In (10.53), we assume that the functions f1(x), . . . , fn(x) ∈ Shave distant supports; the error should not exceed

Cνn(f)d−N ,

where N is an arbitrary number, d is the minimal distance between

the supports, and C depends on N).

The following statement is used in Section 11.5.

Theorem 10.3. Let S and S be scattering matrices constructed for

energy operators H and H, the momentum operator P, and a family

of operators A (this family must be asymptotically commutative

relative to the operators H,P, as well as relative to the operators

H, P). Let us suppose that T is a unitary operator satisfying the

following conditions:

(1) TΦ = Φ,

(2) TΦ(k) = Φ(k),

(3) if A ∈ A, then T−1AT ∈ A (here, Φ and Φ are ground states

and Φ(k) and Φ(k) are single-particle states corresponding to the

energy operators H and H).

If D is a dressing operator for the operator H, satisfying the

conditions of Theorem 10.2, then the operator D = TD is a dressing

operator for the energy operator H.

The proof of this theorem consists of checking that the conditions

of Theorem 10.2 apply to the operator D. This verification can be

Page 281: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 234

234 Mathematical Foundations of Quantum Field Theory

done with the following transformations:

〈Da+(f1) . . . a+(fn)θ,A1(g1) . . . An(gn)Φ〉

= 〈Da+(f1) . . . a+(fn)θ, T−1A1(g1) . . . An(gn)TT−1Φ〉

= 〈Da+(f1) . . . a+(fn)θ,AT1 (g1) . . . ATn (gn)Φ〉

≈∑π

⟨Φ(f1), ATi1(gi1)Φ

⟩. . .⟨Φ(fn), ATin(gin)Φ

⟩=∑π

〈TΦ(f1), Ai1(gi1)TΦ〉 . . . 〈TΦ(fn), Ain(gin)TΦ〉

=∑π

〈Φ(f1), Ai1(gi1)Φ〉 . . . 〈Φ(fn), Ain(gin)Φ〉

(here, we have used the notation T−1AiT = ATi ).

The statements we have proved in this section allow us to give a

new definition of Møller matrices and the scattering matrix. Namely,

if the operator D satisfies conditions (10.52) and (10.53), then we

can define Møller matrices using equations (10.41) and (10.42). The

scattering matrix, as usual, can be defined by the relation S = S∗+S−;

in the case when the scattering matrix is unitary, this definition is

equivalent to the formula

S = S−1+ S−

= slimt→+∞,t0→−∞

exp(iHast)D−1 exp(−iH(t− t0))D exp(−iHast0).

(10.54)

Let us now consider how the the formulas (10.41), (10.42) and

(10.54) are connected to the physical picture of particle scattering.

For this discussion, we will fix an operator D, satisfying the

conditions (10.52) and (10.53).

Let us first consider the collision of two particles. A single particle

with the wave function f will be written as the vector

Φ(f) =

∫f(k)Φ(k)dk = Da+(f)θ.

Let us ask the following question: which vector in the space H should

be used to describe a state with two particles having wave functions

f1 and f2?

Page 282: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 235

Axiomatic Scattering Theory 235

Before we answer this question, note that its formulation in

such a general form is physically meaningless: we should assume

that the particles are far from each other. More formally, if two

particles have wave functions f1(k), f2(k), we should assume that

the supports of the functions f1(x), f2(x) (the wave functions of these

particles in coordinate representation) are far from each other. With

this narrowed framing of the question, we can now give a definite

answer. Namely, we should represent such a state with the vector

Da+(f1)a+(f2)θ.

Indeed, condition (10.52), imposed on the operator D, corre-

sponds to the physical representation of the state with two distant

particles Φ(f1) and Φ(f2).

Analogously, the vector Da+(f1) . . . a+(fn)θ, in the case when the

supports of the functions f1(x), . . . , fn(x) are far from each other,

describes the state of n spatially separated particles.

It should now be easy to see that the collision of two particles can

be described by the vector

x(t) = exp(−iHt)S−a+(f1)a+(f2)θ,

where f1, f2 are non-overlapping functions. Indeed, from the rela-

tion (10.41), it follows that for t→ −∞, we have

x(t) ≈ D exp(−iHast)a+(f1)a+(f2)θ = Da+(f

t1)a+(f

t2)θ,

and hence as t → −∞, x(t) describes the state of two spatially

separated particles Φ(f t1) and Φ(f t2) (technically, it is wrong to say

that the supports of the functions f t1, f t2 are far from each other;

as discussed above, it is in fact their essential supports that can be

shown to be far from each other).

In particular, the vector x(0) = S−a+(f1)a+(f2)θ depicts the

state of a system of two particles at t = 0, whose wave functions at

t → −∞ are f t1, f t2 (if each particle moves freely, then they can be

described at t = 0 by the wave functions f1 and f2). Analogously, we

can interpret other vectors of the form S−ξ and S+η.

It should now be clear that the operator S defines the transition

from the initial state to the final state in a particle collision. Indeed,

Page 283: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 236

236 Mathematical Foundations of Quantum Field Theory

if η = Sξ, then the vector

x(t) = exp(−iHt)S−ξ = exp(−iHt)S+η

depicts the collision process where we begin (at t→ −∞) in the state

x(t) ≈ D exp(−iHast)ξ,

and end (at t→ +∞) in the state

x(t) ≈ D exp(−iHast)η.

Let us now investigate the relationship between the matrix

entries of the scattering matrix (scattering amplitudes) and the

experimentally observed quantities.

For simplicity, let us consider the collision of two particles. We

assume that a beam of particles with momentum q scatters on a

resting particle. Let us assume that the beam has unit flux (the

meaning of this, in both the classical and quantum situations, is

explained in the Chapter 5, Section 5.2). In the simplest case, two

particles remain at the end of the scattering process, however we will

consider the more general situation where new particles can be born

and we can thus have more than two particles at the end. We should

predict the probability (or, more precisely, the probability density)

that at the end we observe n particles with momenta (p1, . . . ,pn).

It will be more convenient to consider the probability σG that the

vector of momenta lies in the domain G in 3n-dimensional space E3n

(if the particles are identical, we should assume that the domain G is

invariant with respect to permutations). This number describes how

many collisions per second end up with n particles, whose momentum

vector p1, . . . ,pn belongs to the domain G. This number is called the

effective collision cross-section (and called the effective differential

cross-section in the case of an infinitesimally small domain G).

Let us now show that the collision cross-section σG can be

expressed using the scattering amplitudes by the following formula:

σG =1

n!

1

|v(q)|

∫G|sn,2(p1, . . . ,pn|q, 0)|2δ(p1 + · · ·+ pn − q)

× δ(ω(p1) + · · ·+ ω(pn)− ω(q)− ω(0))dp1 . . . dpn. (10.55)

Page 284: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 237

Axiomatic Scattering Theory 237

Here, sn,k are functions that are related to the matrix elements Sn,kof the scattering matrix by the relation

Sn,k(p1, . . . ,pn|q1, . . . ,qk)

=⟨Sa+(q1) . . . a+(qk)θ, a

+(p1)a+(pn)θ⟩

= sn,k(p1, . . . ,pn|q1, . . . ,qk)δ(p1 + · · ·+ pn − q1 − · · · − qk)

× δ(ω(p1) + · · ·+ ω(pn)− ω(q1)− · · · − ω(qk)) (10.56)

(sometimes the term scattering amplitudes refers to the functions sn,k

as well). The symbol v(k), as always, denotes the function ∂ω(k)∂k ; we

assume that v(0) = 0. Formula (10.55) is proved below under the

assumption that q 6= 0 and the function sn,2 is continuous at points

(p1, . . . ,pn|q, 0), where (p1, . . . ,pn) ∈ G.

To understand the collision cross-sections in the case at hand,

we will use the construction described in Section 5.2. In general, the

following discussion is largely similar to the discussion in Section 5.2.

Namely, let us assume that the process of collision is described by

the vector D exp(−iH0t)ξα for t→ −∞, where

ξα =

(∫exp(ikα)f(k)a+(k)dk

)(∫g(k)a+(k)dk

)θ, (10.57)

f(k) and g(k) are normalized wave functions differing from zero only

in a small neighborhood of the points q and 0, respectively; α is a

vector orthogonal to the vector v(q) (as explained in Section 5.2, this

vector is analogous to the impact parameter in classical mechanics.)

The collision process is described by the vector

xα(t) = exp(−iHt)S−ξα,

and at the end of the process (with t→ +∞) we obtain the state

D exp(−iH0t)(Sξα).

The probability that for initial state (10.57) we obtain the final

state Sξα of n particles with momenta (p1, . . . ,pn), belonging to

Page 285: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 238

238 Mathematical Foundations of Quantum Field Theory

the domain G, can be written in the form

wα(G) =1

n!

∫G

∣∣∣∣∫ dk1dk2Sn,2(p1, . . . ,pn|k1,k2)

× f(k1) exp(ik1α)g(k2)

∣∣∣∣2 dp1 . . . dpn (10.58)[we have used the relation

Sξα =∑m

1

m!

∫Sm,2(p1, . . . ,pm|k1,k2)f(k1)

× exp(ik1α)g(k2)a+(p1) . . . a+(pm)θdp1 . . . dpmdk1dk2

].

The collision cross-section σG is equal to5

σG =

∫α⊥v(q)

dαwα(G). (10.59)

Substituting into (10.59) the expression (10.58) and integrating over

α, we see that

σG =1

n!

∫Gdp1 . . . dpn

∫dk1dk

′1dk2dk

′2(sn,2(p1,

. . . ,pn|k1,k2)sn,2(p1, . . . ,pn|k′1,k′2)f(k1)f(k′1)

× g(k2)g(k′2)δ(kT1 − k′T1 ))

=1

n!

∫Gdp1 . . . dpn

∫dk1dk

′1dk2dk

′2(sn,2(p1, . . . ,pn|k1,k2)

× sn,2(p1, . . . ,pn|k′1,k′2)f(k1)f(k′1)g(k2)g(k′2)

× δ(kT1 − k′T1 )δ(p1 + · · ·+ pn − k1 − k2)δ(p1 + · · ·+ pn

−k′1 − k′2)δ(ω(p1) + · · ·+ ω(pn)− ω(k1)− ω(k2))

× δ(ω(p1) + · · ·+ ω(pn)− ω(k′1)− ω(k′2)) (10.60)

5Strictly speaking, this formula contains a limit: the probabilities wα(G) shouldbe defined with the functions fν(k), gν(k), whose supports contract to the pointsq, 0, respectively, as ν → ∞.

Page 286: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 239

Axiomatic Scattering Theory 239

[here the symbols kT1 , k′T1 denote the projections of the vectors k1,

k′1 on the plane orthogonal to the vector v(q)].

In further calculations, we will use the following relations:

δ(p1 + · · ·+ pn − k1 − k2)δ(p1 + · · ·+ pn − k′1 − k′2)

= δ(p1 + · · ·+ pn − k1 − k2)δ(k1 + k2 − k′1 − k′2),

δ(ω(p1) + · · ·+ ω(pn)− ω(k1)− ω(k2))δ(ω(p1) + · · ·

+ω(pn)− ω(k′1)− ω(k′2))

= δ(ω(p1) + · · ·+ ω(pn)

−ω(k1)− ω(k2))δ(ω(k1) + ω(k2)− ω(k′1)− ω(k′2)),

δ(kT1 − k′T1 )δ(k1 + k2 − k′1 − k′2)δ(ω(k1) + ω(k2)− ω(k′1)− ω(k′2))

=|v(q)|

(v(k1)− v(k2))v(q)δ(k1 − k′1)δ(k2 − k′2)

+α(k1,k2)δ(k′1 − β(k1,k2))δ(k′2 − γ(k1,k2)),

where β(k1,k2), γ(k1,k2) are non-trivial solutions of the following

system:

ω(k1) + ω(k2)− ω(β)− ω(γ) = 0,

k1 + k2 − β − γ = 0,

kT1 − βT = 0

(i.e. solutions other than β = k1,γ = k2).

Using these formulas, we can transform expression (10.60) into

the form

σG =1

n!

∫Gdp1 . . . dpn

∫dk1dk2

(|v(q)|

(v(k1)− v(k2)) · v(q)

× |sn,2(p1, . . . ,pn|k1,k2)|2|f(k1)|2|g(k2)|2

× δ(p1 + · · ·+ pn − k1 − k2)δ(ω(p1) + · · ·+ ω(pn)

− ω(k1)− ω(k2))

)

Page 287: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 240

240 Mathematical Foundations of Quantum Field Theory

+1

n!

∫Gdp1 . . . dpn

∫dk1dk2(sn,2(p1,

. . . ,pn|k1,k2)sn,2(p1, . . . ,pn|β(k1,k2),γ(k1,k2))

× f(k1)f(β(k1,k2)g(k2)g(γ(k1,k2))α(k1,k2)

× δ(p1 + · · ·+ pn − k1 − k2)δ(ω(p1)

+ · · ·+ ω(pn)− ω(k1)− ω(k2))).

To obtain the final expression for σG, we should use the fact that

for normalized functions f(k) with support in a small neighborhood

of the point q and for functions λ(k), continuous at the point q, we

have the relations ∫λ(k)|f(k)|2dk = λ(q),∫

λ(k)f(k)f(ρ(k))dk = 0 (10.61)

(in formulas (10.61) there is a limit in which the support of the

function f(k) contracts to the point q; ρ(k) is a function that is

continuous at the point q and satisfying the condition ρ(q) 6= q).

10.5 Generalizations

In this section, we describe some generalizations of the results of

previous sections of this chapter.

1. The condition of strong convexity imposed on the dispersion

law ω(p) can be substantially relaxed. Namely, it is enough to require

that for any open set G ⊂ E3 the function ω(p) is not linear (i.e. for

any neighborhood of a point p0 ∈ E3 there is another point p1 ∈ E3

such that ∂ω(p)∂p |p=p1 6=

∂ω(p)∂p |p=p0).

In this case, the definition of the Møller matrix should be modi-

fied: in the formula (10.1), we should consider only non-overlapping

families of functions f1, . . . , fn. The main modification required in the

proofs is to use Lemma 10.5 of Section 10.3 in place of Lemma 10.2

of Section 10.1.

These modifications in the definitions and proofs allow us to

construct a scattering theory in the one- and two-dimensional cases,

Page 288: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 241

Axiomatic Scattering Theory 241

i.e. in the case when the momentum operator P has only one or

two components instead of three. (In higher dimensions, there are no

essential modifications.)

2. Up to this point, we have only considered situations with a

single one-particle state. The generalization to an arbitrary number

of one-particle states is straightforward.

A good operator B, in this case, can be defined as a smooth

operator with (1) B∗Φ = 0; (2) we can find a particle Φi(k) and a

function φB(k) for which BΦ = Φi(φB) =∫φB(k)Φi(k)dk. [Recall

that in Section 10.1 we have agreed to fix a complete orthogonal

system of particles Φ1(k), . . . ,Φs(k).]

The modifications of the statements and their proofs consist

mostly of adding indices; perhaps the most important change comes

from the need to use the word “total” instead of “dense” in the

formulation of Lemma 10.1.

It is easy to check that the Møller matrices and the scattering

matrix do not depend on the choice of complete system of particles

(i.e. they depend only on the operators H,P, and the algebra A).

3. All the constructions in this chapter can be adapted, with

the appropriate modifications, to the case when we replace the

condition of asymptotic commutatitivty by the condition of asymp-

totic anticommutatitivity of the family of operators A (asymptotic

anticommutativity means that for any two operators A,B ∈ A and

any n, we can find C and r such that

‖ [A,B(x, t)]+‖ ≤ C1 + |t|r

1 + |x|n

).

If we have an asymptotically anticommutative family A, we

impose an additional condition that any vector A1 . . . A2k+1Φ, where

Ai ∈ A, is orthogonal to the vacuum Φ (this condition becomes

necessary to construct the scattering matrix). The space H can then

be expanded into a direct sum of two subspaces Hg and Hu in such a

way that the vector A1 . . . AnΦ belongs to the space Hg, if n is even,

and to the space Hu, if n is odd. We will call vectors in the spaces

Hg and Hu even and odd, respectively.

Let us use the symbol A to denote the smallest algebra of

operators that contains the family A. The algebra A, like the linear

Page 289: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 242

242 Mathematical Foundations of Quantum Field Theory

space above, can be represented as a direct sum Ag + Au, where

every operator A ∈ Ag asymptotically commutes with every operator

B ∈ A and every operator A ∈ Au asymptotically anticommutes

with every operator B ∈ Au (this can be easily checked if we

note that the operator A1 . . . Am and the operator B1 . . . Bn, where

Ai ∈ A, Bj ∈ A, asymptotically commute in the case when one of

the numbers m,n is even and asymptotically anticommute in the case

when both numbers m and n are odd). Operators from the family Agwill be called even and operators from the family Au will be called

odd. It is easy to check that an even operator preserves the parity

of vectors (i.e. transforms the spaces Hg and Hu into themselves),

while an odd operator transforms an even vector into an odd vector

and an odd vector into an even vector.

Let us suppose that the particle system Φ1(k), . . . ,Φs(k) is chosen

in such a way that each particle has a definite parity. In other words,

the particles can be split into even, whose vectors Φi(f) ∈ Hg, and

odd, whose vectors Φi(f) ∈ Hu. Even particles should be considered

as bosons, while odd particles as fermions. This implies that the

space Has can be constructed as the Fock space representation of the

relations

[ai(k), aj(k′)]∓ = [a+

i (k), a+j (k′)]∓ = 0,

[ai(k), a+j (k′)]∓ = δijδ(k− k′),

where the anticommutator is used in the case when both particles

Φi(k),Φj(k) are odd and the commutator is used in the other cases.

One can say that

Has = F s(L2(E3 ×N1))⊗ F a(L2(E3 ×N2)),

where N1 is the set of even particles and N2 is the set of odd particles.

In the case considered in this subsection, the construction of the

scattering matrices is performed in the same way as when the family

A is asymptotically commutative.

In the situation at hand, the elementary particles are fermions

while bosons arise only as composite particles. If the elementary

particles can be either fermions or bosons, then the family A must

Page 290: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 243

Axiomatic Scattering Theory 243

be a union of an asymptotically commutative family A1 and an

asymptotically anticommutative family A2, where each operator

from A1 must asymptotically commute with any operator from A2.

All the results of Chapter 10, as well as the results of the following

chapters, can be easily extended to the case when we have both

fermions and bosons as elementary particles.

4. We have previously assumed that the single-particle spectrum

does not intersect the multi-particle spectrum (in other words, that

the laws of conservation of energy and momentum prohibit the decay

of particles). This condition is not always satisfied. However, in the

theory of elementary particles, if a particle is stable, then its decay in

all known cases is prohibited either by the conservation of energy and

momentum or by some other conservation law. In this case, one can

prove all the results established above by the same methods. More

precisely, in axiomatic scattering theory, it is enough to assume that

the space H expands into a direct sum of the spaces Ni, that are

invariant with respect to the operators H,P, in such a way that for

every i, the spectrum of the operators H,P in the subspace H1 ∩Nidoes not intersect the spectrum of these operators in the subspace

M∩Ni (here, M is the multi-particle subspace). In this case, however,

it is necessary to tighten the cyclicity requirement by assuming that

vectors of the form AΦ, where A ∈ A, are dense in each space Ni.5. Let us describe a generalization of the concept of an asymp-

totically commutative algebra that allows us to study asymptotic

commutative algebras containing unbounded operators.

Let us suppose that the energy operator H and the momentum

operator P = (P1, P2, P3) act on the Hilbert spaceH. We will assume,

as always, that H, P1, P2, P3 are commuting self-adjoint operators,

whose spectrum satisfies the conditions outlined in Section 10.1.

Let us fix a linear subspace D that is dense in H, contains the

ground state Φ of the energy operatorH, and is invariant with respect

to operators of the form exp(iHt− iPx). (In this subsection, we will

assume that these conditions always apply to the operators H,P and

the set D.) A family A, consisting of operators defined on the set D

that transform the set into itself, will be called asymptotically Abelian

algebra, if

Page 291: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 244

244 Mathematical Foundations of Quantum Field Theory

C1. For every operator A ∈ A, the operators A+ and A(x, t) also

belong to A; for every pair of operators A,B ∈ A, their linear

combination λA+ µB and product AB also belong to A.

C2. For any operators A1, . . . , Ar ∈ A and arbitrary function f ∈S(E4r), the operator∫

f(x1, t1, . . . ,xr, tr)A1(x1, t1) . . . Ar(xr, tr)drxdrt (10.62)

belongs to the algebra A and the number⟨∫f(x1, t1, . . . ,xr, tr)A1(x1, t1) . . . Ar(xr, tr)d

rxdrtΦ,Φ

⟩(10.63)

continuously depends on the function f ∈ S (the integral

in (10.62) is understood in a weak sense).6

C3. For any operators A,B,A1, . . . , Ar ∈ A and any number n, we

can find numbers C and s such that

| 〈[A(x, t), B]A1(ξ1, τ1) . . . Aj(ξj , τj)Φ,

Aj+1(ξj+1, τj+1) . . . Ar(ξr, τr)Φ〉 |

≤ C(1 + |t|s)1 + |x|n

(1 +

r∑i=1

|ξi|+r∑i=1

|τi|

)k, (10.64)

where k is a number not depending on n.

We will assume that the vector Φ is a cyclic vector of the

asymptotic commutative algebra A. Then, using the earlier consid-

erations of this chapter, we can define Møller matrices and scattering

matrices, corresponding to the operators H,P, and the algebra A,

prove their existence, and transfer to this case the results of earlier

considerations. (The definitions, theorems, and proofs do not require

any significant modifications.)

6The second part of the condition C2 implies that the function 〈A1(x1, t1) . . .Ar(xr, tr)Φ,Φ〉 is a locally summable function of moderate growth.

Page 292: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 245

Axiomatic Scattering Theory 245

Let us add to these results the following statement which will be

used in Section 11.4.

Suppose thatB(k|t) is an operator function, which is a generalized

function in the variable k, that satisfies the following conditions:

(1) exp(iHτ − iPa)B(k|t) exp(−iHτ + iPa) = exp(−ika)B(k|t+ τ);

(2) B+(k|t)Φ = 0;

(3) vectors of the form∫f(k)B(k|t)dkΦ belong to the single-particle

subspace;

(4) the functions ρm,n defined by the equation

ρm,n(k1, t1, . . . ,km, tm|k′1, t′1, . . . ,k′n−1, t′n−1)

× δ(k1 + · · ·+ km − ε1k′1 − · · · − εnk′n)

=⟨A1(k1|t1) . . . Am(km|tm)B(k′1, ε1|t′1) . . . B(k′n, εn|0)

⟩T,

where Ai(k|t) = (2π)−3∫Ai(x, t) exp(ikx)dx; Ai ∈ A; B(k,

1|t) = B+(k|t); B(k,−1|t) = B(k|t) belongs to the space

S(E3(m+n−1)) over the variables k1, . . . ,km,k′1, . . . ,k

′n−1 for

fixed t1, . . . , tm, t′1, . . . , t

′n−1. [We will call such generalized func-

tions “good”; this name is motivated by the fact that for every

good operator B, we can construct a good operator gener-

alized function B(k|t) = (2π)−3∫B(x, t) exp(ikx)dx.] Then,

for any system of smooth functions with compact support

f1(k), . . . , fm(k), we have

limt→±∞

B(f t1|t) . . . B(f tn|t)Φ = S±a+(φf1) . . . a+(φfn)θ,

where

f ti (k) = fi(k) exp(−iω(k)t),

B(f ti |t) =

∫f ti (k)B(k|t)dk,

B(k|0)Φ = φ(k)Φ(k).

The proof of this statement follows the reasoning in Section 10.1.

Let us now define the notion of asymptotically commutative

families of operator generalized functions.

Page 293: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 246

246 Mathematical Foundations of Quantum Field Theory

A family B of operator generalized functions A(x, t) is called

asymptotically commutative if there exists an asymptotically com-

mutative algebra A that contains all operators of the form

A(f) =∫f(x, t)A(x, t)dxdt, where A(x, t) belongs to the family

B, f ∈ S(E4). (We will assume that the vector Φ is a cyclic vector of

the family of operators A(f). The operators A(f) should have a com-

mon domain of definition D and transform it into itself; the domain

of definition D′ of operators belonging to the algebra A may coincide

with the set D, but in general must at least contain it. When we say

that the algebra A contains the operator A(f), we mean that in this

algebra, we can find an operator coinciding with A(f) on the set D.)

The scattering matrix corresponding to the operators H,P and

the asymptotically commutative family of operator generalized func-

tions B can be defined as the scattering matrix constructed for H,P,

and the asymptotic algebra A, containing all operators A(f). (It is

easy to see that this scattering matrix does not depend on the choice

of algebra A.) We can define Møller matrices analogously.

The following theorem gives sufficient conditions for the asymp-

totic commutativity of a family of operator generalized functions.

Let us suppose that B is a family of operator generalized functions

satisfying the following conditions:

D1. For any operator generalized function A(x, t) ∈ B, the operators

A(f) =∫f(x, t)A(x, t)dxdt, where f ∈ S(E4), are defined on a

set D and transform it into itself.

D2. The functional 〈A(f)ψ1, ψ2〉, where A ∈ B, ψ1, ψ2 ∈ D, f ∈S(E4), continuously depends on the function f ∈ S(E4) (i.e. the

real-valued generalized function 〈A(x, t)ψ1, ψ2〉 is a generalized

function of moderate growth).

D3. If A ∈ B, then

exp(iHτ − iPa)A(x, t) exp(−iHτ + iPa) = A(x + a, t+ τ).

D4. For every function A(x, t) in family B, the family also contains

the adjoint operator generalized function A+(x, t).

D5. For any operator generalized functions A1(x, t), . . . , Ar(x, t),

A(x, t), B(x, t) in B, we can find a δ > 0 such that the real-valued

Page 294: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 247

Axiomatic Scattering Theory 247

generalized function

T (x, t, ξ1, τ1, . . . , ξr, τr)

= 〈[A(x + a, t+ α), B(a, α)]A1(ξ1, τ1)

. . . Aj(ξj , τj)Φ, Aj+1(ξj+1, τj+1) . . . Ar(ξr, τr)Φ〉

in the domain satisfying |t| < δ|x| can be represented in the

form

T (x, t, ξ1, τ1, . . . , ξr, τr)

=∑

|α|≤N(n)

D(α)

(1 +∑i

|ξi|2 +∑i

|τi|2)k

× (1 + |x|2)−n σα,n(x, t, ξ1, τ1, . . . , ξn, τn)

(here, n is an arbitrary natural number; σα,n is a bounded

continuous function; D(α) is a differential operator of order |α|with constant coefficients; k is a number not depending on n).

D6. The vector Φ is a cyclic vector of the family of operators A(f),

where A(x, t) ∈ B, f ∈ S(E4).

Then, the family B is asymptotically Abelian.

To prove this theorem, let us consider the family A consisting of

linear combinations of operators of the form∫f(x1, t1, . . . ,xn, tn)A1(x1, t1) . . . An(xn, tn)dnxdnt, (10.65)

where f ∈ S(E4n), Ai(x, t) ∈ B. As it is shown in Appendix A.7,

expressions of the form (10.65) can be seen as operators defined on

some linear subspace D′ and transform D′ into itself (it follows that

we can take D′ to be the set of all vectors of the form AΨ, where

A ∈ A; Ψ ∈ D; expression of the form (10.65) defines an operator on

the set D by the operator analog of the kernel theorem). It is easy to

check that the set D′ contains the set D and is invariant with respect

to the operators exp(iHt− iPa).

Page 295: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 248

248 Mathematical Foundations of Quantum Field Theory

It is easy to check that the family A, when viewed as a family

of operators defined on the set D′, is asymptotically commutative.

Conditions C1 and C2 can be easily checked with the help of the

statements proven in Appendix A.7. For example, if A ∈ A is an

operator of the form (10.65), then the operators

A+ =

∫f(x1, t1, . . . ,xn, tn)A+

n (x1, t1) . . . A+1 (xn, tn)dnxdnt,

A(x, t) =

∫f(x1 − x, t1 − t, . . . ,xn − x, tn − t)

×A1(x1, t1) . . . An(xn, tn)dnxdnt

are also contained in A. Inequality (10.64) follows from condition D5.

Let us now give a simple example of objects satisfying the

conditions of axiomatic scattering theory.

Let us consider the Fock space H = F (L2(E3)) with the free

Hamiltonian H0 =∫ε(k)a+(k)a(k)dk and the momentum operator

P =∫

ka+(k)a(k)dk. If ε(k) is a smooth, strongly convex function,

satisfying the condition ε(k1)+ε(k2) > ε(k1 +k2), then the spectrum

of these operators satisfies the conditions outlined in Section 10.1.

[The ground state of this system is the Fock vacuum θ and the single-

particle state is defined by the formula Φ(k) = a+(k)θ.] The space

Has can be identified with the spaceH; the Hamiltonian Has becomes

the Hamiltonian H0 in this identification.

The operator generalized functions

a+(x, t) = (2π)−3/2

∫exp(−ikx)a+(k, t)dk,

a(x, t) = (2π)−3/2

∫exp(ikx)a(k, t)dk,

where

a+(k, t) = exp(iH0t)a+(k) exp(−iH0t)

= exp(iω(k)t)a+(k),

a(k, t) = exp(iH0t)a(k) exp(−iH0t) exp(−iω(k)t)a(k),

constitute an asymptotically commutative family B0 of generalized

functions (as always, we assume that these operator generalized

Page 296: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 249

Axiomatic Scattering Theory 249

functions act on the linear manifold S∞ consisting of vectors of the

form ∑n

∫φn(k1, . . . ,kn)a+(k1) . . . a+(kn)θdnk,

where the functions φn belong to the space S and only a finite number

of these functions are non-zero). The family B0 can be shown to

be asymptotically commutative with the aid of the theorem proven

above. A different proof of the same fact can be obtained if we note

that operators of the form∫f(x, t)a+(x, t)dxdt and

∫f(x, t)a(x, t)dxdt,

where f ∈ S(E4) are contained in the asymptotically commutative

algebra A0 consisting of operators of the form∑m,n

∫fm,n(k1, . . . ,km|p1, . . . ,pn)a+(k1)

. . . a+(km)a(p1) . . . a(pn)dmkdnp, (10.66)

where fm,n ∈ S(E3(m+n)), the sum in (10.66) is finite, and operators

of the form (10.66) are viewed as operators, defined on the set S∞. It

is easy to check directly that A0 is an asymptotically commutative

algebra; this can also be shown by results we formulate below. It is

easy to verify that the Møller matrix S± and the scattering matrix

S, constructed for the operators H0,P and the family A0, are trivial

(i.e. they are all the identity operator).

It is sometimes convenient to assume that an asymptotically

commutative algebra is endowed with a topology (this will be useful

in Sections 10.6 and 11.5). We will now give some conditions that

when applied to a topological operator algebra guarantee that it will

be asymptotically commutative.

Let us assume that on the Hilbert space H we have, as before, an

energy operator H, a momentum operator P, and a linear subspace

D, satisfying the conditions outlined above, and let us consider a set

of operators A obeying the following conditions:

T1. The set A consists of operators that are defined on the set D

and transform D into itself.

Page 297: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 250

250 Mathematical Foundations of Quantum Field Theory

T2. For any A,B in A, the following operators are also in A: the

linear combination λA + µB (i.e. the set can be viewed as a

linear space), the product AB (this implies that A is an algebra

of operators), and the adjoints A+, B+.

T3. The linear space A is equipped with a locally convex topology;

the space A must be complete, and the product and conjugation

operations must be continuous in the topology A.7

T4. The vector AΦ ∈ H continuously depends on A ∈ A in the

topology of the algebra A [i.e. we can find a seminorm p(A) on

A such that

‖AΦ‖ ≤ p(A)].

T5. If A ∈ A, then the operator A(x, t) is in A and it continuously

depends on x, t in the topology on A; for any seminorm p(A) on

A and any compact set F ⊂ A, we can find another seminorm

q(A) and a number k, such that for A ∈ F , x ∈ E3,−∞ < t <∞we have

p(A(x, t)) ≤ (1 + |x|k + |t|k)q(A).

T6. For any seminorm p(A) on A, any compact set F ⊂ A, and any

natural number n, there exists a seminorm q(A) such that

p([A(x), B]) ≤ q(A)q(B)

1 + |x|n

for all A,B ∈ F , x ∈ E3.

(If the locally convex topology on A is specified by a system

of seminorms ‖A‖α, then conditions T5 and T6 should be

reformulated in the following way: for all seminorms ‖A‖α,

compact sets F ⊂ A, and natural numbers n, there must exist

a seminorm ‖A‖λ in the system and numbers k,C such that

‖A(x, t)‖α ≤ C(1 + |x|k + |t|k)‖A‖λ,

‖[A(x), B]‖α ≤C‖A‖λ‖B‖λ

1 + |x|n

for any A,B ∈ F .)

7Conditions T1–T3 mean that the set A is a complete locally convex topologicalalgebra with involution.

Page 298: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 251

Axiomatic Scattering Theory 251

T7. The vector Φ is a cyclic vector with respect to the family of

operators A.

We can now state the following proposition:

A family of operators A satisfying conditions T1–T6 is an asymp-

totically commutative algebra.

Indeed, condition C1 follows from conditions T1 and T2. To

check C2, note that for any A1, . . . , An ∈ A the operators

A1(x1, t1) . . . An(xn, tn) continuously depend on x1, t1, . . . ,xn, tn in

the topology on A and for any seminorm p(A) on A we can find a

number k such that

p(A1(x1, t1) . . . An(xn, tn)) ≤ 1 + |x|k + |t|k (10.67)

(this follows from conditions T3 and T5). These remarks and the

completeness of A imply that the integral (10.62) converges in the

topology A and defines an element of the algebra A. It follows from

condition T4 and inequality (10.67) that

〈A1(x1, t1) . . . An(xn, tn)Φ,Φ〉

is a continuous function of moderate growth in the variables

x1, t1, . . . ,xn, tn. This proves the second part of condition C2. Finally,

we can easily check C3 by using T3–T6.

The statement we have just proved allows us to say that conditions

T1–T7 are sufficient to construct a scattering theory starting with

the operators H,P and the algebra A.

Let us now return to the operators H0 =∫ε(k)a+(k)a(k)dk,P =∫

ka+(k)a(k)dk and the algebra A0, consisting of operators of the

form (10.66). We will show that the algebra can be endowed with a

topology in a way that satisfies conditions T1–T7. The algebra A0

can be represented as the union of linear spaces A(m,n) consisting of

operators of the form

A =

∫fm,n(k1, . . . ,km|p1, . . . ,pn)a+(k1) . . . a+(km)a(p1)

. . . a(pn)dmkdnp.

Elements of the space A(m+n) are in one-to-one correspondence with

functions in the space S(E3(m+n)); this allows us to transfer the

Page 299: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 252

252 Mathematical Foundations of Quantum Field Theory

topology of the space S(E3(m+n)) to the space A(m+n). Hence, we

can consider A(m+n) to be a complete locally convex space.

Let us consider a system of seminorms on algebra A0 that are

continuous functions on each subspace A(m,n) ⊂ A0. This system of

seminorms defines a topology on A0 which is called the inductive

limit topology ; the algebra A0 is complete in this topology (see, for

example, Robertson et al. (1980)).

It is easy to check that the algebra A0, with the topology

introduced above, satisfies conditions T1–T7. To verify conditions

T5 and T6, it helps to use the following lemmas.

Lemma 10.8. For any norm ‖f‖α,β on the space S(E3(m+n)) we

can find a norm ‖f‖γ,δ on S and numbers C, k such that for any

function f ∈ S(E3(m+n)) we have

‖UtVxf‖α,β ≤ C(1 + |x|k + |t|k)‖f‖γ,δ[here, Vx and Ut denote operators in S(E3(m+n)) transforming the

function f(k1, . . . ,km|p1, . . . ,pn) to the functions

exp

i m∑j=1

kj − tn∑j=1

pj

x

f(k1, . . . ,km|p1, . . . ,pn),

exp

−i m∑j=1

ε(kj)−n∑j=1

ε(pj)

t

f(k1, . . . ,km|p1, . . . ,pn),

respectively ].

Lemma 10.9. Foranynorm‖f‖α,β onthespaceS(E3(m+m′+n+n′−2r))

and any number N , we can find norms ‖f‖γ,δ and ‖f‖γ′,δ′ on the

spaces S(E3(m+n)) and S(E3(m′+n′)) and a number C so that

‖λr(Vxf, g)‖α,β ≤C

1 + |x|N‖f‖γ,δ‖g‖γ′,δ′ .

[Here, r > 0 and λr(f, g) denotes the function∫f(k1, . . . ,km|p1, . . . ,pn−r,q1, . . . ,qr)

×g(q1, . . . ,qr,k′1, . . . ,km′−r|p′1, . . . ,p′n)dq1 . . . dqr.]

Page 300: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 253

Axiomatic Scattering Theory 253

It follows from Lemma 10.8 that for any seminorm p(A) on the

space A(m+n) we can find a seminorm q(A) on A(m+n) and a number

k so that for any operator A ∈ Am,n we have

p(A(x, t)) ≤ (1 + |x|k + |t|k)q(A).

It follows from Lemma 10.9 that for any seminorm p(A) on

the space A0 and for any number N we can find seminorms q(A)

and q′(A) on the spaces A(m,n) and A(m′,n′) so that for any A ∈A(m,n), B ∈ A(m′,n′), we have

p([A(x), B]) ≤ q(A)q′(B)

1 + |x|N.

To finish checking conditions T5 and T6, we should note that every

compact subset of the algebra A0 is contained in a direct sum of a

finite number of subspaces A(m,n) (see Robertson et al. (1980)).

Therefore, the algebra A0 satisfies conditions T1–T7; as we have

remarked above, this implies that A0 is asymptotically commutative.

10.6 Adiabatic theorem in axiomatic scattering

theory

We will now show how in axiomatic scattering theory one can express

Møller matrices and scattering matrices in terms of their adiabatic

analogs.

First, we will define the adiabatic Møller matrix and the adiabatic

scattering matrix in a slightly more general situation than was

considered in Section 4.1.

Let the self-adjoint operator H(g) act on the Hilbert space H,

with the parameter g taking values in the interval [0, 1]. We will

fix a continuous function h(τ) of a real variable τ which rapidly

decays at infinity and equals 1 at τ = 0. The symbol Uα(t, t0)

will denote the evolution operator, constructed with the time-

dependent Hamiltonian H(h(ατ)), and the symbol Sα(t, t0) will

denote the operator exp(iH(0)t)Uα(t, t0) exp(−iH(0)t0). Otherwise,

we can define the operator Sα(t, t0) as the solution to

Page 301: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 254

254 Mathematical Foundations of Quantum Field Theory

the equation

idSα(t, t0)

dt= exp(iH(0)t)(H(h(αt))−H(0)) exp(−iH(0)t)Sα(t, t0)

with the initial condition Sα(t0, t0) = 1.

We will similarly use U(t, t0|g(τ)) to denote the evolution operator

corresponding to the time-dependent Hamiltonian H(g(τ)), where

g(τ) is a function taking values in [0, 1]; using this notation, we can

write Uα(t, t0) = U(t, t0|h(ατ)).

We will call the operators Sα(0,±∞) = slimt→±∞ Sα(0, t) the

adiabatic Møller matrices and the the operator Sα = Sα(∞,−∞) =

slim t→∞t0→−∞

Sα(t, t0) will be called the adiabatic S-matrix. The con-

struction of the adiabatic S-matrix and the adiabatic Møller matrices

provided in Section 4.1 for the pair of operators H,H0 corresponds

to H(g) = H0 + g(H −H0) and h(τ) = exp(−|τ |).In what follows, for the sake of simplifying the proofs, we will

assume that the function h(τ) satisfies a few somewhat stronger

conditions than we have assumed above. Namely, we will assume

that this function is smooth, even, has compact support, and satisfies

h(0) = 1 in a neighborhood of the point τ = 0. The radius of this

neighborhood will be denoted by the symbol δ and the radius of the

support of the function h(τ) will be denoted by ∆ (i.e. h(τ) = 1 for

|τ | < δ and h(τ) = 0 for |τ | > ∆).

Let us now consider the situation when the energy operator in

an axiomatic scattering theory depends on a parameter g. More

precisely, suppose that we have an energy operator H(g) that acts on

the Hilbert space H, where the parameter g is in the interval [0, 1], a

momentum operator P = (P1, P2, P3), and a family of operators A,

satisfying the following conditions:

(a) The self-adjoint operators H(g), P1, P2, P3 commute for every

g. In the space H, there exists a vector Φ (the ground state of the

energy operator H(g), not depending on the parameter g) and a

vector function Φ(k|g), generalized with respect to k and continuous

in the parameter g [single-particle state of the operator H(g)], for

which we have

Page 302: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 255

Axiomatic Scattering Theory 255

(1) H(g)Φ = PΦ = 0;

(2) H(g)Φ(k|g) = ω(k|g)Φ(k|g),

where ω(k|g) is a positive function, infinitely differentiable in k, g

and strongly convex in k;

(3) PΦ(k|g) = kΦ(k|g);

(4) 〈Φ(k|g),Φ(k′|g)〉 = δ(k− k′);

(5) for any k0 ∈ E3, 0 ≤ g0 ≤ 1, we can find an operator A ∈ A,

such that 〈Φ(k0|g0), AΦ〉 6= 0;

(6) for any k0 ∈ E3, 0 ≤ g0 ≤ 1, there exist ε > 0, δ > 0, so that for

any point (ω,k) belonging to the multi-particle spectrum of the

operators (H(g),P) and satisfying |k− k0| < δ, |g − g0| < δ, we

have the inequality ω > ω(k0|g0) + ε.

(b) Let us denote by D the set of vectors of the form AΦ where

A ∈ A. We assume that D is dense in H and invariant with respect

to the operators exp(iPx) and U(t, t0|g(τ)), where g(τ) is a smooth

function taking values in [0, 1]. For any two operators A,B ∈ A,

the linear combination λA + µB, the product AB, and the adjoint

operators A+, B+ all belong to A (in other words, the family A is

an operator algebra with an involution). We will assume that A is

equipped with a locally convex topology, in which A is complete and

the product and involution operations are continuous.

The topology on A should satisfy the following conditions:

(1) 〈AΦ,Φ〉 continuously depends on A ∈ A;

(2) if A ∈ A, g(τ) is a smooth function with values in [0, 1], x ∈ E2,

and −∞ < t0, t1 <∞, then the operator

U(t1, t0|g(τ))A(x)U(t0, t1|g(τ)) ∈ A;

for every seminorm p(A) on A and every compact F ⊂ A we can

find a seminorm q(A) on A and a number k such that for A ∈ F ,

we have

p(U(t1, t0|g(τ))A(x)U(t0, t1|g(τ))

≤ (1 + |t1 − t0|k + |x|k)q(A). (10.68)

Page 303: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 256

256 Mathematical Foundations of Quantum Field Theory

In particular, A(x, t|g) ∈ A; we will suppose that this operator

is a smooth function in g with respect to the topology A, the

operator ∂m

∂gmA(x, t|g) continuously depends on x, t, g and

p

(∂m

∂gmA(x, t|g)

)≤ (1 + |x|k + |t|k)q(A), (10.69)

where p(A) is any seminorm on A; m is an arbitrary whole

number; the seminorm q(A) on A and the number k depend

on the seminorm p, the number m, and the compact set F . [Here

and further, A(x) denotes the operator exp(−iPx)A exp(iPx)

and A(x, t|g) denotes exp(iH(g)t)A(x) exp(−iH(g)t)];

(3) for any seminorm p(A) on A, any number n, and any compact

F ⊂ A, we can find a seminorm q(A) on A such that for all

A,B ∈ A,x ∈ E3, A ∈ F we have

p([A(x), B]) ≤ q(A)q(B)

1 + |x|n. (10.70)

These conditions are sufficient to guarantee that starting from the

operators H(g),P, and the algebra A, we can construct the Møller

matrices S±(g) and the scattering matrix S(g) = S∗+(g)S−(g). (Not

very precisely, one can say that we require the operators H(g),P, and

the algebra A to satisfy the conditions T1–T7 of the Section 10.5,

which guarantee the existence of Møller matrices, but we additionally

require these conditions to be satisfied uniformly in g).

We will now show that we can derive the following relations from

the above conditions:

S−(1) = slimα→0

Sα(0,−∞) exp

(i

α

∫ρ(k)a+

in(k)ain(k)dk

),

(10.71)

S+(1) = slimα→0

Sα(0,+∞) exp

(i

α

∫ρ(k)a+

out(k)aout(k)dk

),

(10.72)

Page 304: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 257

Axiomatic Scattering Theory 257

where

ρ(k) =

∫ 0

−∞(ω(k|h(σ))− ω(k|0))dσ,

ain(k) = S−(0)a(k)S−1− (0),

aout(k) = S+(0)a(k)S−1+ (0)

are in- and out-operators, constructed for the energy operator H(0).

The proof of these relations (the adiabatic theorem) is the main result

of this section.

Of course, one can easily state the analog of relations (10.71) and

(10.72) for the operators S±(g), where 0 ≤ g ≤ 1; we will not stop

to do this.

The relations (10.71) and (10.72) are clearly equivalent to the

relations

S−(1) = slimα→0

slimt→−∞

Uα(0, t)S−(0)Wα(t), (10.73)

S+(1) = slimα→0

slimt→+∞

Uα(0, t)S+(0)Wα(t), (10.74)

where Wα(t) is an operator in the space Has, defined by the formula

Wα(t) = exp

(i

∫rα(k|t)a+(k)a(k)dk

),

rα(k|t) =

∫ 0

tω(k|h(ατ))dτ

(to verify this equivalence, it is enough to note that

exp(−iH(0)t)S−(0) = S−(0) exp

(−i∫ω(k|0)ta+(k)a(k)dk

);

1

αρ(k) = lim

t→−∞(rα(k|t)− tω(k|0))).

We will prove the adiabatic theorem in the form (10.73) and

(10.74). The proof will use the following series of lemmas.

Lemma 10.10. For any seminorm p(A) on A, any compact F ⊂ A,

and any number n, we can find a seminorm q(A) on A and a number

Page 305: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 258

258 Mathematical Foundations of Quantum Field Theory

k, such that

p([U(t1, t0|g(τ))A(x)U(t0, t1|g(τ)), B])

≤ 1 + |t1 − t0|k

1 + |x|nq(A)q(B)

(here, A,B ∈ F , g(τ) is a smooth function with values in [0, 1],

−∞ < t1, t0 <∞,x ∈ E3).

The proof of this lemma quickly follows from inequalities (10.68)

and (10.70) and the remark that the operator U(t1, t0|g(τ)) com-

mutes with exp(iPx).

Lemma 10.11. If φ(x, t) is a piecewise-continuous function, decay-

ing faster than a power function, then the operator∫φ(x, t)A(x, t|g)dxdt,

where A ∈ A, 0 ≤ g ≤ 1, belongs to the algebra A and is infinitely

differentiable with respect to g in the topology of A, and furthermore

the operator

dk

dgk

∫φ(x, t)A(x, t|g)dxdt

continuously depends on φ ∈ T , A ∈ A, g ∈ [0, 1] in the topology

on A.

[We will assume that the space T of piecewise-continuous func-

tions with decay faster than a power function is endowed with a

topology by means of the norm family

‖φ‖λ = supx∈E3,−∞<t<∞

(1 + |x|λ + |t|λ)φ(x, t).

]The proof of this lemma is based on the completeness of the algebra

A and condition b(2).

Lemma 10.12. Suppose that Ag, Bg ∈ A are two families of

operators, infinitely differentiable in g in the topology on A. Let us

Page 306: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 259

Axiomatic Scattering Theory 259

assume that B∗gΦ = 0 for all g (this condition is satisfied if Bg is a

good operator). Then the function

ρ(k|g) =

∫〈Ag(x)Φ, BgΦ〉 exp(ikx)dx

is infinitely differentiable with respect to k, g.

Indeed, let us consider the function

ρ(x|g) = 〈Ag(x)Φ, BgΦ〉 =⟨B∗gAg(x)Φ,Φ

⟩.

By Lemma 10.10, the function

∂m

∂gmρ(x|g) =

m∑k=0

Ckm

⟨∂kB∗g∂gk

∂m−kAg(x)

∂gm−kΦ,Φ

=

m∑k=0

Ckm

⟨[∂kB∗g∂gk

,∂m−kAg(x)

∂gm−k

]Φ,Φ

obeys the inequality ∣∣∣∣ ∂m∂gm ρ(x|g)

∣∣∣∣ ≤ C

1 + |x|n

(here, m and n are arbitrary natural numbers). It follows from this

inequality that the function ρ(k|g) =∫ρ(x|g) exp(ikx)dx is infinitely

differentiable.

Before we move on to the next lemma, let us note that the vector

generalized function Φ(k|g) defined by conditions a(2), a(3), a(4)

is not uniquely specified: a vector generalized function Φ′(k|g) =

exp(iλ(k|g))Φ(k|g), where λ(k|g) is a real measurable function, will

satisfy the same conditions.

Lemma 10.13. For any operator A ∈ A, we can find a real

measurable function λ(k|g) so that⟨AΦ,Φ′(k|g)

⟩= 〈AΦ, exp(iλ(k|g))Φ(k|g)〉

is infinitely differentiable in k and g.

Page 307: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 260

260 Mathematical Foundations of Quantum Field Theory

Let us suppose that B1g , B

2g are two families of good operators that

are infinitely differentiable in g in the topology of A; the functions

r1(k|g) and r2(k|g) will be defined by the relation

BigΦ =

∫ri(k|g)Φ(k|g)dk.

Let us consider the function

ν(x|g) =⟨B1g(x)Φ, B2

gΦ⟩

and note that

ν(x|g) =

⟨exp(iPx)

∫r1(k|g)Φ(k|g)dk,

×∫r2(k|g)Φ(k|g)dk

⟩=

∫exp(ikx)r1(k|g)r2(k|g)dk. (10.75)

It follows from relation (10.75) and Lemma 10.12 that the function

r1(k|g)r2(k|g) is infinitely differentiable in k and g.

Let us now show that for all points (k0, g0) we can find an

infinitely differentiable family of good operators Bg, such that

〈Bg0Φ,Φ(k0|g0)〉 6= 0.

We will use the fact that for every operator A ∈ A and every

smooth function with compact support σ(k, ω|g), which is equal to

0 for ω ≤ 0 and for any point (ω,k) belonging to the multi-particle

spectrum of the operators H(g),P, we can construct a family of good

operators Bσ,Ag by means of the formula

Bσ,Ag =

∫σ(x, t|g)A(x, t|g)dxdt,

where

σ(x, t|g) =1

(2π)4

∫exp(ikx− iωt)σ(k, ω|g)dkdω

(see Lemma 10.1 in Section 10.1). By Lemma 10.11, this family is

infinitely differentiable in g. It is easy to check that⟨Bσ,Ag Φ,Φ(k|g)

⟩= σ(k, ω(k|g)|g) 〈AΦ,Φ(k|g)〉 . (10.76)

Page 308: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 261

Axiomatic Scattering Theory 261

For every point k0, g0, we can find a function σ such that

σ(k0, ω(k0|g0)|g0) 6= 0. This observation, combined with condi-

tion a(5) shows that for every point k0, g0 we can find an operator

A ∈ A such that ⟨Bσ,Ag0 Φ,Φ(k0|g0)

⟩6= 0.

We can now use the following statement.

Let F = φγ(k) be a family of measurable complex functions

on the convex subset M of the Euclidean space En, satisfying (a)

for any two functions φγ ∈ F , φδ ∈ F , the function φγ(k)φδ(k)

is infinitely differentiable; (b) for any point k ∈ M , we can find a

function φγ ∈ F , not equalling zero at the point k. Then we can find

a measurable real function λ(k) with the property that the product

of any function φγ ∈ F with exp(iλ(k)) is a smooth function.

(The proof of this statement, based on several theorems from the

topology of fiber bundles, is given in Fateev and Shvarts (1973).)

Applying this statement, we can find a real function λ(k|g) such

that for any family of good operators Bg ∈ A, which are infinitely

differentiable in g in the topology of A, the function

exp(iλ(k|g))r(k|g) = exp(iλ(k|g)) 〈BgΦ,Φ(k|g)〉

is infinitely differentiable.

It follows from relation (10.76) that the same choice of the

function λ(k|g) implies the differentiability of all functions of the

form exp(iλ(k|g)) 〈AΦ,Φ(k|g)〉, where A ∈ A; this finishes the proof

of the lemma.

Remark 10.4. Using the same arguments, one can prove, without

requiring condition a(5), that for any operator A ∈ A, the function

|〈Φ(k|g), AΦ〉|2 is infinitely differentiable. This allows us to consider

the value of the function |〈Φ(k|g), AΦ〉| at individual points. There-

fore, condition a(5) has a precise meaning.

In what follows, we will assume that the single-particle state

Φ(k|g) is chosen in such a way that for any operator A ∈ A, the

function 〈AΦ,Φ(k|g)〉 is infinitely differentiable in k, g [Lemma 10.13

Page 309: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 262

262 Mathematical Foundations of Quantum Field Theory

implies that this can always be done by replacing Φ(k|g) with

Φ′(k|g) = exp(iλ(k|g))Φ(k|g) if necessary].

Lemma 10.14. For any smooth function with compact support

f(k|g), we can find a family of good operators Bfg , infinitely dif-

ferentiable in g in the topology of A, such that

BfgΦ =

∫f(k|g)Φ(k|g)dk.

To prove this lemma, it is convenient to use the relation

Bψφg =

∫φ(x|g)Bψ

g (x)dx (10.77)

[more precisely, if a family of good operators Bψ

g ∈ A is infinitely

differentiable in g and

Bψg Φ =

∫ψ(k|g)Φ(k|g)dk,

φ(k|g) is smooth function with compact support, and

φ(x|g) =1

(2π)3

∫φ(k|g) exp(ikx)dk,

then formula (10.77) specifies a family of good operators, infinitely

differentiable in g and satisfying the condition

Bψφg Φ =

∫ψ(k|g)φ(k|g)Φ(k|g)dk

].

Let us note that for every point (k0, g0) there exists a neighbor-

hood U , an operator A ∈ A, and a function σ, such that the function

〈Bσ,Ag ,Φ,Φ(k|g)〉 is non-zero in the neighborhood U (here, Bσ,A

g is a

family of good operators, constructed in the proof of Lemma 10.12).

If the support of the function f(k|g) is contained in the neighborhood

U , then the necessary family Bfg can be obtained with the help of

relation (10.77), using Bψg = Bσ,A

g , Bψφg = Bf

g in the relation. In other

words, we should choose function (10.76) as the function ψ(k|g) and

Page 310: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 263

Axiomatic Scattering Theory 263

the function

f(k|g)

〈Bσ,Ag Φ,Φ(k|g)〉

as the function φ(k|g) (clearly, the latter function is smooth and has

compact support if we define it to be zero at the values where both

the numerator and the denominator are zero).

An arbitrary function f can be represented as a finite sum of

functions fi, for which the family Bfig can be obtained by the

construction above. It is clear that the operators

Bfg =

∑Bfig

are infinitely differentiable in g and satisfy the conditions

BfgΦ =

∫f(k|g)Φ(k|g)dk,

(Bfg )∗Φ = 0.

However, these functions do not, a priori, satisfy the condition of

smoothness entering the definition of a good operator. To satisfy

this condition, we can replace the operators Bfg by the operators

Bfg =

∫µ(x, t|g)B(x, t|g)dxdt,

where

µ(x, t|g) =1

(2π)4

∫µ(k, ω|g) exp(ikx− iωt)dkdω,

µ(k, ω|g) is a smooth function with compact support, equal to 1 if

(k, g) ∈ supp f , ω = ω(k|g).

Lemma 10.15. The function ν(k|g) defined by the formula⟨dΦ(f |g)

dg,Φ(f ′|g)

⟩=

∫ν(k|g)f(k)f ′(k)dk, (10.78)

where f(k), f ′(k) are smooth functions with compact support, is

smooth in k and g.

Page 311: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 264

264 Mathematical Foundations of Quantum Field Theory

Let us first note that, by Lemma 10.14, the vector Φ(f |g) can

be represented in the form BfgΦ, where the operator Bf

g is infinitely

differentiable in g; hence, the derivative dΦ(f |g)dg exists. Furthermore,

the bilinear form

A(f, f ′) =

⟨dΦ(f |g)

dg,Φ(f ′|g)

⟩is translation-invariant [i.e. A(f, f ′) = A(fx, f

′x), where fx(k) =

exp(−ikx)f(k)] and, therefore, can be written in the form (10.78).

Let us consider the function

ζ(x|g) =

⟨dΦ(f |g)

dg,Φ(f ′x|g)

⟩=

⟨dΦ(f |g)

dg, exp(−iPx)Φ(f ′|g)

=

⟨dBf

g

dgΦ, Bf ′

g (x)Φ

⟩.

It is clear that

ζ(x|g) =

∫exp(ikx)ν(k|g)f(k)f ′(k)dk. (10.79)

It follows from relation (10.79) and Lemma 10.12 that the function

ν(k|g)f(k)f ′(k) is smooth. Since f(k), f ′(k) are arbitrary smooth

functions, the proof of the lemma follows.

It follows from Lemma 10.15 that the single-particle state Φ(k|g)

can be chosen in such a way that for every f, f ′ we have⟨dΦ(f |g)

dg,Φ(f ′|g)

⟩= 0. (10.80)

[If equality (10.80) is not satisfied, then we can satisfy it if we replace

Φ(k|g) with the generalized vector function

Φ(k|g) = exp(iτ(k|g))Φ(k|g),

where τ(k|g) is a solution of the equation

i∂τ(k|g)

∂g= −ν(k|g).

]

Page 312: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 265

Axiomatic Scattering Theory 265

Lemma 10.15 implies that τ(k|g) is a smooth function and hence the

function 〈AΦ, Φ(k|g)〉 is smooth for all A ∈ A.

In what follows, we will assume that the single-particle state

Φ(k|g) is chosen such that it satisfies condition (10.80).

Let us denote by t0 a moment in time, satisfying the relation

t0 ≤ −∆α (recall that the symbol ∆ denotes the support radius of the

fixed function h(τ), so that h(τ) = 0 for τ ≤ αt0).

Lemma 10.16. For any smooth function with compact support f(k),

we have

Uα(t, t0)Φ(f |0)

= exp(irα(P|t)− irα(P|t0))(Φ(f |h(αt)) + αξ1(f |αt)

+ · · ·+ αnξn(f |αt) + αn+1ηn+1(f, α, t, t0)),

where ξi(f |σ) is defined by the equations

idξi−1(f |σ)

dσ= (H(h(σ))− ω(P|h(σ)))ξi(f |σ), (10.81)⟨

dξi(f |σ)

dσ,Φ(f ′|h(σ))

⟩= 0, (10.82)

ξ0(f |σ) = Φ(f |h(σ)), (10.83)

ξ0(f |σ0) = 0 for j > 0 (10.84)

and ‖ηn+1(f, α, t, t0)‖ ≤ C (where C does not depend on α, t, t0, but

may depend on f and n).

The proof of this lemma is based on arguments similar to the

arguments in Section 4.3. The vector Ψ(t) = Uα(t, t0)Φ(f |0) satisfies

the equation

idΨ

dt= H(h(αt))Ψ(t).

By using the change of variables σ = αt, this equation can be written

in the form

iαdΨ

dσ= H(h(σ))Ψ(σ). (10.85)

Page 313: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 266

266 Mathematical Foundations of Quantum Field Theory

We will look for solutions of (10.85) of the form

Ψ(σ) = exp

(− iα

∫ σ

σ0

ω(P|h(σ′))dσ′)

(ξ0(f |σ) + αξ1(f |σ)

+ · · ·+ αnξn(f |σ) + αn+1ηn+1(f, α, σ)).

Let us note that equation (10.81) specifies the vector ξi(f |σ) up

to a constant belonging to the kernel of the operator H(h(σ)) −ω(P|h(σ)) (i.e. the vector is specified up to a vector from the

single-particle subspace). Equations (10.82)–(10.84) fix this constant

vector uniquely. Further on (see Lemma 10.17), the vector ξi(f |σ) is

infinitely differentiable in σ.

The function ηn+1(f, α|σ) satisfies the equation

iαdηn+1(f, α|σ)

dσ= (H(h(σ))− ω(P|h(σ)))ηn+1(f, α|σ)

− idξn(f |σ)

dσ(10.86)

with initial condition ηn+1(f, α|σ0) = 0. From this equation and

the remark that∥∥dξn(f |σ)

∥∥ is bounded above by a constant, not

depending on σ, it is easy to obtain the inequality

‖ηn+1(f, α|σ)‖ ≤ C

α

(the constant C depends on f). Noting the fact that

ηn(f, α|σ) = ξn(f |σ) + αηn+1(f, α|σ),

we see that

‖ηn(f, α|σ)‖ ≤ const.

This finishes the proof of the lemma.

Lemma 10.17. The vector ξi(f |σ), entering the formulation of

Lemma 10.16, can be written in the form

ξi(f |σ) = Df,σi Φ,

where Df,σi is a family of operators, infinitely differentiable in σ in

the topology on A, hence, this vector is infinitely differentiable in σ.

Page 314: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 267

Axiomatic Scattering Theory 267

We will prove this lemma by induction. Suppose the operatorDf,σi−1

is given. Let us consider the vector ζi(f |σ), defined by the relations

idξi−1(f |σ)

dσ= (H(h(σ))− ω(P|h(σ)))ζi(f |σ),⟨

ζi(f |σ),Φ(f ′|h(σ))⟩

= 0,

where f , f ′ are arbitrary smooth functions with compact support.

This vector can be written in the form

ζi(f |σ) = γ(H(h(σ)),P|h(σ))dξi−1(f |σ)

dσ. (10.87)

Here,

γ(ω,p|g) = i(ω − ω(p|g))−1,

if p ∈ supp f , and (ω,p) belongs to the spectrum of the operators

H(g),P, but ω 6= ω(p|g); if ω = ω(p|g), then γ(ω,p|g) should be

equal to zero. The function γ(ω,p|g) can be chosen to be smooth

and equal to zero for large |p|.Using relation (10.87), we can write

ζi(f |σ) =

∫γ(t,x|h(σ))Ef,σi−1(x, t|σ)Φdxdt, (10.88)

where γ(t,x|g) is the Fourier transform of γ(ω,p|g) in the variables

ω,p and Ef,σi−1 = (d/dσ)Df,σi−1. Equation (10.88) implies that ζi(f |σ) =

F f,σi Φ, where F f,σi =∫γ(t,x|h(σ))Ef,σi−1(x, t|σ)dxdt (this operator is

infinitely differentiable in σ, by Lemma 10.11). The vector ξi(f |g)

can be expressed in terms of the vector ζi(f |σ) in the following way:

ξi(f |σ) = ζi(f |σ) + Φ(λσi f |h(σ)), (10.89)

where λσi (k) is a function satisfying the equation∫∂λσi (k)

∂σf(k)f ′(k)dk = −

⟨dζi(f |σ)

dσ,Φ(f ′|h(σ))

⟩(10.90)

with the initial condition λσ0i (k) = 0 for σ0 < −∆. By introduc-

ing the vector generalized functions ξi(k|σ) and ζi(k|σ) with the

Page 315: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 268

268 Mathematical Foundations of Quantum Field Theory

relations ξi(f |σ) =∫f(k)ξi(k|σ)dk, ζi(f |σ) =

∫f(k)ζi(k|σ)dk, the

formulas (10.89) and (10.90) can be written in the form

ξi(k|σ) = ζi(k|σ) + λσi (k)Φ(k|h(σ)),

∂λσi (k)

∂σδ(k− k′) = −

⟨∂ζi(k|σ)

∂σ,Φ(k′|h(σ))

⟩.

The function λσi (k) is infinitely differentiable in σ; this can be shown

using the ideas in the proof of Lemma 10.15. We can therefore select

an operator Df,σi of the form

Df,σi = F f,σi +B

λσi f

h(σ).

It is easy to see that the operator Df,σi is infinitely differentiable

in σ and, therefore, satisfies the conditions of the lemma (the

differentiability of the operator F f,σi has already been shown, while

operator Bλσi f

h(σ) can be written in the form

Bλσi f

h(σ) = Bµσi f

h(σ) =

∫µσi (x)Bf

h(σ)(x)dx,

where µσi (k) = β(k)λσi (k) and the symbol β(k) denotes a smooth

function with compact support, equal to 1 when k ∈ suppf).

Remark 10.5. For all σ in the interval [−δ, δ], where the function

h equals 1, we can show the following equations by induction:

dξi(f |σ)

dσ= 0; ζi(f |σ) = 0; F f,σi = 0;

∂λσi (k)

∂σ= 0; Df,σ

i = Bλσi f1 .

Let us define the functions φt,α(x), φt,α(k) by the formulas

φt,α(x) = (2π)−3

∫φt,α(k) exp(ikx)dk,

φt,α(k) = φ(k) exp(irα(k|t)).

Lemma 10.18. Let φ(k) be a smooth function with compact support.

For any whole number n and any ε > 0 we can find a number Cn,ε,

Page 316: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 269

Axiomatic Scattering Theory 269

not depending on x, t, and α, so that

|φt,α(x)| ≤ Cn,ε1 + (ρ(x, V t

α(φ))− ε|t|)n

(here, V tα(φ) is a set of points which can be written in the form

∇rα(k|t), where k ∈ suppφ, and ρ(x, A) is the distance between the

point x and the set A). There exists a constant C, not depending on

x, t, α, so that

|φt,α(x)| ≤ C|t|−3/2,∫|φt,α(x)|dx ≤ C(1 + |t|3/2). (10.91)

The proof is analogous to the proof of Lemma 10.2 in Section 10.1.

Lemma 10.19. If f(k), φ(k) are smooth functions with compact

support, then the vector

Γ(t) = Uα(t, t0)S−(0)Wα(t0)a+(f φ)θ,

where t0 ≤ −∆α , can be written in the form

Γ(t) = QtsΦ +Rs(t),

where

Qts = Qts(f, φ, α) =

∫φt,α(x)N t,α

s (x)dx, (10.92)

N t,αs =

s∑i=0

Df,αti ,

‖Rs(t)‖ ≤ Cs(f, φ)αs+1(1 + |t|3/2). (10.93)

Indeed,

Wα(i0)a+(f φ)θ = a+(f φt0,α)θ.

Therefore,

Γ(t) = Uα(t, t0)Φ(fφt0,α|0)

= Uα(t, t0)φt0,α(P)Φ(f |0)

= φt0,α(P)Uα(t, t0)Φ(f |0).

Page 317: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 270

270 Mathematical Foundations of Quantum Field Theory

Hence, we can conclude from Lemmas 10.16–10.18 that

Γ(t) = φt0,α(P) exp(irα(P|t)− irα(P|t0))

×

(s∑i=0

αiξi(f |αt) + αs+1ηs+1(f, α, t, t0)

)

= φt,α(P)

(s∑i=0

Df,αti Φ

)+Rs(t)

= φt,α(P)N t,αs Φ +Rs(t)

=

∫φt,α(x)N t,α

s (x)Φdx +Rs(t),

which proves the necessary statement [to obtain an estimate for Rs,

we should use an estimate for ‖ηs+1‖, used in Lemma 10.16, and

inequality (10.91)].

Remark 10.6. The relation

Uα(t′, t)Uα(t, t0) = Uα(t′, t0)

implies that

Uα(t′, t)Γ(t) = Γ(t′). (10.94)

Using equation (10.94) and Lemma 10.18, we can conclude that

Uα(t′, t)QtsΦ = Qt′s Φ +Rs(t, t

′), (10.95)

where ‖Rs(t, t′)‖ ≤ Cαs+1 and the constant C depends on f and φ

(but does not depend t, t′, α). We will use relation (10.95) later.

Lemma 10.20. Let φ1(k), φ2(k) be non-overlapping smooth func-

tions with compact support. Then there exists a number c > 0 such

that the distance between V tα(φ1) and V t

α(φ2) is larger than c|t|.

It follows from the strong convexity of ω(k|g) that

d2ω(k|g) ≥ λdk2 (10.96)

Page 318: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 271

Axiomatic Scattering Theory 271

(here, λ is positive and does not depend on k, g, if k runs over a

bounded set; the second differential is taken over only the variable k).

Inequality (10.96) implies that d2rα(k|t) ≥ λ|t|dk2, from which we

can conclude that the second derivative of rα(k|t), in any direction,

cannot be less than λ|t|. Indeed, we can set

e =k1 − k2

|k1 − k2|, ν(ζ) = rα(k2 + ζe|t)

and conclude that ν ′′(ζ) ≥ λ|t|. Hence, it is clear that

|∇rα(k1|t)−∇rα(k2|t)|

≥ |v′(|k1 − k2|)− v′(0)| ≥ λ|t||k1 − k2|.

Therefore, if the distance between the sets suppφ1 and suppφ2 is

larger than ε, then the distance between the sets V tα(φ1) and V t

α(φ2)

is larger than λ|t|ε. This finishes the proof of the lemma.

Lemma 10.21. Let f1(k), f2(k), φ1(k), φ2(k) be smooth functions

with compact support and suppose that the supports of the functions

φ1, φ2 do not intersect. Then

p([Uα(t+ τ, t)Q1,ts Uα(t, t+ τ), Q2,t

s ]) ≤ C|τ |r

1 + |t|n

[here, p(A) is an arbitrary seminorm on the algebra A; n is an

arbitrary number; −∞ < τ , t <∞, |τ | ≤ |t|; C and r do not depend

on t, τ , and α; the symbol Qi,ts denotes the operator

Qi,ts = Qts(fi, φi, α),

defined by the relation (10.92)].

The proof of this lemma follows directly from Lemmas 10.10,

10.18, and 10.20.

Lemma 10.22. Let φ1(k), . . . , φn(k) be smooth functions with

compact and non-intersecting supports, let f1(k), . . . , fn(k) be smooth

functions with compact support, and let t ≤ −aα−ε, t0 ≤ −∆α , a > 0,

Page 319: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 272

272 Mathematical Foundations of Quantum Field Theory

ε > 0, s ≥ 32n+ 1. Then the vector

Ψn(t) = Uα(t, t0)S−(0)Wα(t0)b+(f1φ1) . . . b+(fnφn)θ

can be written in the form

Ψn(t) = Q1,ts . . . Qn,ts Φ + π(t), (10.97)

where

Qi,ts = Qts(fi, φi, α),

‖π(t)‖ ≤ C|t− t0|α2,

and the number C does not depend on t, t0, and α.

It is easy to see that the vector Ψn(t) does not depend on t0, if t0is less than −∆/α, hence we may assume that t0 is the largest integer

obeying t0 < −∆/α. We will assume that t is an integer and we will

prove the relation (10.97) by induction on t; the generalization to

real-valued t does not present any obstacles.

Let us suppose that

Ψn(t) ≈ Q1,ts . . . Qn,ts Φ. (10.98)

Using equation (10.95), we can conclude that

Ψn(t+ 1) = Uα(t+ 1, t)Ψn(t)

≈ Uα(t+ 1, t)Q1,ts . . . Qn,ts Φ

= Uα(t+ 1, t)Q1,ts . . . Qn−1,t

s Uα(t, t+ 1)Uα(t+ 1, t)Qn,ts Φ

≈ Uα(t+ 1, t)Q1,ts . . . Qn−2,t

s Uα(t, t+ 1)Uα(t+ 1, t)

×Qn−1,ts Uα(t, t+ 1)Qn,t+1

s Φ (10.99)

An additional error, arising in the transition from approxima-

tion (10.98) to approximation (10.99) by (10.95), can be written in

the form

Uα(t+ 1, t)Q1,ts . . . Qn−1,t

s Uα(t, t+ 1)Rs(t, t+ 1),

Page 320: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 273

Axiomatic Scattering Theory 273

where ‖Rs(t, t + 1)‖ ≤ Cαs+1(1 + |t|3/2) and any seminorm of

the operator Qi,ts on the A does not exceed C(1 + |t|3/2) [the

last statement follows from (10.91)]. Using these remarks and the

inequalities |t| ≤ |t0|, s ≥ 32n + 1, we can show that as we go

from (10.98) to (10.99) the error does not increase by more than

Cα2.

By Lemma 10.21, we can see that

Ψn(t+ 1) ≈ Uα(t+ 1, t)Q1,ts . . . Qn−2,t

s

×Uα(t, t+ 1)Qn,t+1s Uα(t+ 1, t)Qn−1,t

s Φ. (10.100)

Similarly, the additional error obtained from transferring to (10.100)

from (10.99) does not exceed Cα2 (moreover, this error does not

exceed Cαr, where r is arbitrary). To prove this, we should use, in

addition to Lemma 10.21, the inequality t < −aα−ε and the estimate

for Qi,ts shown above.

Using equation (10.95) again, we obtain

Ψn(t+ 1) ≈ Uα(t+ 1, t)Q1,ts

. . . Qn−2,ts Uα(t, t+ 1)Qn,t+1

s Qn−1,t+1s Φ.

Applying equation (10.95) several times and then applying

Lemma 10.21, we obtain

Ψn(t+ 1) ≈ Qn,t+1s . . . Q1,t+1

s Φ,

≈ Q1,t+1s . . . Qn,t+1

s Φ. (10.101)

The error arising from each transformation does not exceed Cα2.

Hence, we have derived (10.101) from (10.98) and have proven that

the error in (10.101) also does not exceed Cα2. The inductive step

is shown. Since the case t = t0 for equation (10.97) is obvious,

Lemma 10.22 is proven.

We can now finally easily show the main result of this section.

In the formulation of Lemma 10.22, let us take t = − δα , where δ

is such that h(τ) ≡ 1 for −δ ≤ τ ≤ δ. Then we can show that

φt,αi (x) = φi(x, t)

= (2π)−3

∫φi(k) exp(ikx− iω(k|1)t)dk, (10.102)

Page 321: Mathematical Foundations of Quantum Field Theory

March 27, 2020 15:20 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch10 page 274

274 Mathematical Foundations of Quantum Field Theory

Ψn(t) ≈n∏i=1

∫ φi(x, t)

s∑j=0

αiDfi,−δj (x)

dx

Φ,

Dfi,−δ0 = Bfi

1 ,

Dfi,−δj = B

λ−δj fi1 . (10.103)

The error in (10.102) does not exceed Cα2(− δα + ∆

α ) = const α and

inequality (10.103) follows from the remark to Lemma 10.17.

By the definition of the Møller matrix S− and the rela-

tions (10.102) and (10.103) we obtain

S−(1)b+(f1φ1) . . . b+(fnφn)θ

= limt→−∞

n∏i=1

(∫φi(x, t)B

fi1 (x, t|1)dx

= limt→−∞

exp(iH(1)t)n∏i=1

(∫φi(x, t)B

fi1 (x)dx

= limt→−∞

exp(−iH(1)t)

n∏i=1

(∫φi(x, t)D

fi,−δ0 (x)dx

= limα→0

exp

(−iH(1)

δ

α

) n∏i=1

(∫φi

(x,− δ

α

)

×

s∑j=0

αjDfi,−δj (x)

dx

Φ

= limα→0

exp

(−iH(1)

δ

α

)Ψn

(− δα

)= lim

α→0Uα(0, t0)S−(0)Wα(t0)b+(f1φ1) . . . b+(fnφn)θ. (10.104)

To derive equality (10.73) from relation (10.104), it is enough to

recall that if limAnx = Ax for all x in a dense set and the

norms of the operators An are uniformly bounded, then slimAn =A.

Relation (10.74) follows analogously.

Page 322: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 275

Chapter 11

Translation-Invariant Hamiltonians(Further Investigations)

11.1 Connections between the axiomatic theory

and the Hamiltonian formalism

Let H be a formal translation-invariant Hamiltonian of the

form (8.3). Let us suppose that there exists an operator realization

(H, H, P, a(k, ε, t),Φ) of the Hamiltonian H (recall that H is a

Hilbert space; H, P are the operators of energy and momentum;

a(k, ε, t) are operator generalized functions; Φ is a ground state).

Let us consider a family A consisting of two operator generalized

functions

a(x, 1, t) = (2π)−3/2

∫exp(ikx)a(k, 1, t)dk,

a(x,−1, t) = (2π)−3/2

∫exp(−ikx)a(k,−1, t)dk.

We will use the term scattering matrix of the formal Hamiltonian

H to refer to the scattering matrix, constructed with the operators

H, P, and the family A (here, we assume that the family A and the

operators H, P satisfy the conditions that guarantee the existence of

the scattering matrix; see Section 10.5).

The relation between this definition and the definition of the

scattering matrix provided in Chapter 9 will be addressed later.

Currently, we will note that in the case when the Hamiltonian H

satisfies the conditions in Section 9.2, then it follows from the results

of Section 10.3 that the scattering matrix constructed in Section 9.2

coincides with the above definition.

275

Page 323: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 276

276 Mathematical Foundations of Quantum Field Theory

Indeed, let λ(k) ∈ S(E3), f(k), and µ(τ) be smooth functions

with compact support and B =∫λ(k)µ(τ)a+(k, τ)dkdτ . It is clear

that ∫f(x|t)B(x, t)dx

=

∫f(x|t) exp(−ikx)λ(k)µ(τ)a+(k, t+ τ)dkdxdτ

=

∫f(k)λ(k)µ(τ)a+(k, t+ τ)dkdτ,

and, therefore, in the conditions of Section 9.2, we have

wlimt→−∞

∫f(x|t)B(x, t)dx =

∫f(k)λ(k)µ(τ)Λ

−1(k)a+

in(k, τ)dkdτ

=

∫f(k)λ(k)µ(ω(k))Λ

−1(k)a+

in(k)dk,

(11.1)

where a+in(k, τ), ain(k, τ) are in-operators, defined in Section 9.2;

µ(ω(k)) =∫µ(τ) exp(iω(k)τ)dτ .

Relations (9.18) and (9.21) imply that⟨BΦ, a+

in(p)Φ⟩

=

∫λ(k)µ(τ)

⟨a+(k, τ)Φ, a+

in(p)Φ⟩dkdτ

=

∫λ(k)µ(ω(k))

⟨a+(k)Φ, a+

in(p)Φ⟩dk

= λ(p)µ(ω(p))ρ(1,p) = λ(p)µ(ω(p))Λ−1

(p).

(11.2)

Comparing equations (11.1) and (11.2) with relation (10.37),

we see that the in-operators, defined in Section 9.2, coincide with

the in-operators defined in Section 10.3; an analogous proof can be

carried out for out-operators and for scattering matrices.

We should note, however, that the present definition of scattering

matrix can also be used in the presence of bound states. (We define

bound state as a single-particle state Φ(k) satisfying the condition

Page 324: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 277

Translation-Invariant Hamiltonians (Further Investigations) 277

〈Φ(k), a+(k′, t)Φ〉 = 0. Other single-particle states are called ele-

mentary. The construction of Section 9.2 works only for elementary

particles.)

The questions of the existence of an operator realization of the

formal Hamiltonian H and whether the operators H, P, and the

operator A satisfy the requirements in Section 10.5 turn out to be

non-trivial.

In this section, we will show that these questions can be answered

in the framework of perturbation theory; more precisely, we will

construct the relevant objects as formal power series in the coupling

constant.1

Let us consider the Hermitian and translation-invariant formal

expression

H = H0 + V

=

∫ε(k)a+(k)a(k)dk +

∑r≥1

gr∑m,n

∫v(r)m,n(k1, . . . ,km|p1, . . . ,pn)

× δ(k1 + · · ·+ km − p1 − · · · − pn)a+(k1) · · · a+(km)a(p1)

· · · a(pn)dmkdnp. (11.3)

a+(k), a(k) are, as usual, symbols satisfying CCR. We will assume

that ε(k) is a smooth function whose derivatives are of moderate

growth, the functions v(r)m,n belong to the space S and for every r, only

a finite number of the functions v(r)m,n are non-zero. The number g will

be called the coupling constant. If the sum in (11.3) is infinite, then

we will consider it simply as a formal power series without assuming

1In other words, we will consider formal expressions∑

nAngn, where An are

numbers, vectors, or operators; if An are operators, we will assume that theyare all defined on the same domain. The sums and products of two (numberor operator) formal series

∑nAng

n and∑

nBngn are understood as the series∑

n(An + Bn)gn and∑

n(∑

k+l=nAkBl)gn. The rest of this section will be

considered in the framework of perturbation theory (without always stating so).In order to give precise mathematical meaning to the results in this section andtheir proofs, all the relations should be understood in terms of formal power series.

Page 325: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 278

278 Mathematical Foundations of Quantum Field Theory

anything about its convergence. The set of all formal expressions

having these properties will be denoted by M.

An expression H ∈ M can be seen as a formal translation-

invariant Hamiltonian (the most interesting are the Hamiltonians

in the set M that can be written in the form H = H0 + gW , where

W does not already contain the coupling constant g; however, it will

be convenient to consider the general case).

An operator realization of a Hamiltonian H ∈ M will be defined

to be a Hilbert space H with the action of the energy operator H =∑Hrg

r and the momentum operator P, the operator generalized

functions a(k, ε, t) =∑grar(k, ε, t) and the vector Φ, satisfying the

operator realization conditions in every order with respect to g (the

series∑Hrg

r and∑grar(k, ε, t) are to be understood as formal

power series in g without convergence assumptions). For the operator

H =∑grHr, we can construct the operators exp(iHt) =

∑grUr(t)

which constitute a group of unitary operators.2 The operators Ur(t)

and the operators∫f(k)ar(k, ε, t)dk must be defined on the same

domain D and must transform it to itself; the set D must contain

the vector Φ.

In Section 11.3, we will show that for every Hamiltonian H ∈M,

we can construct an operator realization in the above sense (for

Hamiltonians that generate vacuum polarization, we will assume

that the function ε(k), entering in the definition of the Hamiltonian

H0, satisfies the condition ε(k1 + k2) < ε(k1) + ε(k2), however this

requirement is not necessary).

Any translation-invariant Hamiltonian H = H0 + W generates

a family of Hamiltonians H0 + gW , depending on the parameter g.

If the Hamiltonian H0 +W satisfies the conditions which guarantee

that the Heisenberg equations (8.5) make sense (these conditions

are described in Section 8.1), then H0 + gW ∈ M; therefore, for

2The operators∑grUr(t) constitute a group of unitary operators if∑

k+l=r

Uk(t)Ul(t′) = Ur(t+ t′),

Uk(0) = 0 for k > 0,

U0(t) = 1.

Page 326: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 279

Translation-Invariant Hamiltonians (Further Investigations) 279

these Hamiltonians, an operator realization can be constructed in

the framework of perturbation theory.

In Section 11.3, we will show that the operators H =∑grHr, P

and the family of operator generalized functions consisting of two

functions a(x, ε, t) =∑grar(x, ε, t) =

∑gr∫

exp(iεkx)ar(k, ε, t)dk,

ε = ±1, satisfy the conditions of Section 10.5 in every order.

The proof uses the additional assumption that the Hamiltonian in

consideration belongs to the subset M of M, whose Hamiltonians

obey the condition that their functions ε(k) satisfy ε(k1 + k2) <

ε(k1) + ε(k2) and are not linear on any open subset.

It follows that the Møller matrices and the scattering matrices

exist in the framework of perturbation theory (more precisely, the

Møller matrices S± and the scattering matrix S can be written as

formal power series S± =∑grS

(r)± , S =

∑grS(r)). To derive this, it

is necessary to analyze the construction of Møller matrices, given in

Chapter 10, and check that the statements proven in this chapter

hold for formal power series in g as well. To see this, note that

we must modify the Møller matrix construction by assuming that

the functions f1(k), . . . , fn(k), entering the formula (10.1), form a

non-overlapping family of functions3 (for details, see Section 11.5).

11.2 Heisenberg equations and canonical

transformations

Let us consider a formal translation-invariant Hamiltonian H,

belonging to the class M described in Section 11.1. As noted

in Section 8.1, if v(r)m,0 6≡ 0 (the Hamiltonian generates vacuum

polarization), then the expression H does not define a self-adjoint

operator on Fock space. Furthermore, the operator exp(iHt) cannot

be constructed in this case even in the framework of perturbation

theory (in terms of a formal power series in g).

This can be easily shown by calculating the operator S(t, t0) =

exp(iH0t) exp(−iH(t − t0)) exp(−iH0t0) by means of the diagram

techniques developed in Section 6.4. For example, in the expression

3The same modification of the Møller matrix definition is used for a differentpurpose in Section 10.5, part 5.

Page 327: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 280

280 Mathematical Foundations of Quantum Field Theory

Fig. 11.1 Divergent diagram.

corresponding to the diagram depicted in Fig. 11.1, we run into the

meaningless expression

δ(k1 + · · ·+ km)δ(k1 − km+1) · · · δ(km − k2m)δ(km+1 + · · ·+ k2m)

= δ2(k1 + · · ·+ km)δ(k1 − km+1) · · · δ(km − k2m).

A similar situation arises for all diagrams without outer vertices (for

vacuum loops): in the expression corresponding to the diagram, the

integrand contains a product of δ-functions whose arguments are

linearly dependent, rendering the expression meaningless.

Therefore, we cannot write a solution to the Schrodinger equation

idψdt = Hψ even in perturbation theory.

However, using the expression (11.3), we can formally write

the Heisenberg equations which remain meaningful in perturbation

theory.

Indeed, the Heisenberg equations,

∂a(k, t)

∂t= i[H, a(k, t)], (11.4)

∂a+(k, t)

∂t= i[H, a+(k, t)], (11.5)

can be written in the form

i∂a(k, t)

∂t− ε(k)a(k, t) = I(k, t), (11.6)

−i∂a+(k, t)

∂t− ε(k)a+(k, t) = I+(k, t), (11.7)

Page 328: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 281

Translation-Invariant Hamiltonians (Further Investigations) 281

where

I(k, t) =∑r≥1

gr∑m,n

m

∫v(r)m,n(k,k1, . . . ,km−1|p1, . . . ,pn) (11.8)

× δ

k +

m−1∑i=1

ki −n∑j=1

pj

a+(k1, t) (11.9)

· · · a+(km−1, t)a(p1, t) · · · a(pn, t)dm−1kdnp. (11.10)

(These equations can be obtained by calculating the commutator

on the right-hand side of the equations (11.4) and (11.5) by using

CCR, distributivity, and the relation [AB,C] = [A,C]B +A[B,C].)

By using the initial conditions a+(k, 0) = a+(k), a(k, 0) = a(k),

equations (11.6) and (11.7) can be written in the form

a(k, t) = exp(−iε(k)t)a(k)

+

∫ t

0exp(−iε(k)(t− τ))I(k, τ)dτ, (11.11)

a+(k, t) = exp(iε(k)t)a+(k)

+

∫ t

0exp(iε(k)(t− τ))I+(k, τ)dτ. (11.12)

We then look for solutions to equations (11.11) and (11.12) in the

form

a(k, t) = exp(−iε(k)t)a(k)

+∑r≥1

gr∑m,n

∫f r,tm,n(k1, . . . ,km|p1, . . . ,pn)

× δ

k +

m∑i=1

ki −n∑j=1

pj

a+(k1) · · · a+(km)a(p1)

· · · a(pn)dmkdnp, (11.13)

Page 329: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 282

282 Mathematical Foundations of Quantum Field Theory

a+(k, t) = exp(iε(k)t)a+(k)

+∑r≥1

gr∑m,n

∫f r,tm,n(k1, . . . ,km|p1, . . . ,pn)

× δ

k +m∑i=1

ki −n∑j=1

pj

a+(p1) · · · a+(pn)a(k1)

· · · a(kn)dmkdnp. (11.14)

Substituting these expressions for a(k, t), a+(k, t) into (11.11) and

(11.12) and equating terms having the same order in g, we obtain a

recurrence relation for f r,tm,n, from which it is clear that the functions

f r,sm,n belong to the space S and for every order of g (i.e. for every r),

only a finite number of functions f r,tm,n are non-zero.

(Here, we view a(k, t) as formal power series consisting of the

symbols a(k), a+(k), obeying CCR; all expressions arising in the

solution can be converted to normal form using CCR. The series

in g are viewed formally.)

It is easy to check that ‖f r,tm,n‖α,β does not grow faster than a

power of t, i.e.

‖f r,tm,n‖α,β ≤ C(1 + |t|k) (11.15)

(here, ‖ · ‖α,β is a norm on the space S; C and k depend on r, α, β).

Inequality (11.15) follows from the following statement.

Theorem 11.1. Let σ(k1, . . . , kn) be a smooth real function whose

derivatives do not grow faster than a power. Then for any norm

‖ · ‖α,β on the space S, we can find a norm ‖ · ‖γ,δ on the space Sand numbers C, r such that for any function φ ∈ S

‖ exp(iσ(k1, . . . , kn)t)φ(k1, . . . , kn)‖α,β≤ C(1 + |t|r)‖φ(k1, . . . , kn)‖γ,δ. (11.16)

Indeed, using inequality (11.16), we see that the norm of the

function

gt(k1, . . . , kn) =

∫ t

0exp(iσ(k1, . . . , kn)(t− τ))hτ (k1, . . . , kn)dτ

Page 330: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 283

Translation-Invariant Hamiltonians (Further Investigations) 283

satisfies the inequality

‖gt‖α,β ≤ C|t|(1 + |t|r) sup0≤τ≤t

‖hτ‖γ,δ,

which implies inequality (11.15) by induction. It is easy to prove that

for fixed t, the expression a+(k, t), a(k, t) satisfies CCR:

[a(k, t), a(k′, t)] = [a+(k, t), a+(k′, t)] = 0,

[a(k, t), a+(k′, t)] = δ(k− k′) (11.17)

(it is important to note that the calculations of the commutators

in (11.17) do not require limits: all the involved series depend

on a finite number of convergent integrals). In order to check

relation (11.17), we can differentiate them in t, for example,

∂t[a(k, t), a+(k′, t)] = i[[H, a(k, t)], a+(k′, t)]

+ i[a(k, t), [H, a+(k′, t)]]

= i[H, [a(k, t), a+(k′, t)]],

which, with the initial condition [a(k, 0), a+(k′, 0)] = [a(k), a+(k′)] =

δ(k− k′), allows us to conclude that [a(k, t), a+(k′, t)] = δ(k− k′).

We will say that the formulas

b(k) = α(k)a(k) +∑r≥1

gr∑m,n

∫h(r)m,n(k1, . . . ,km|p1, . . . ,pn)

× δ

k +m∑i=1

ki −n∑j=1

pj

a+(k1)

· · · a+(km)a(p1) · · · a(pn)dmkdnp, (11.18)

b+(k) = α(k)a+(k) (11.19)

+∑r≥1

gr∑m,n

∫h

(r)m,n(k1, . . . ,km|p1, . . . ,pn)

× δ(k +

m∑i=1

ki −n∑j=1

pj)a+(p1) · · · a+(pn)

× a(k1) · · · a(km)dmkdnp, (11.20)

Page 331: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 284

284 Mathematical Foundations of Quantum Field Theory

where α(k) is a smooth function whose derivatives do not grow faster

than a power of |k|, h(r)m,n ∈ S, and for every r, all but a finite number

of the functions h(r)m,n are zero, specify a canonical transformation if

the following relations are satisfied:

[b(k), b(k′)] = [b+(k), b+(k′)] = 0,

[b(k), b+(k′)] = δ(k− k′).

Roughly speaking, the canonical transformations turn the sym-

bols a+(k), a(k), satisfying CCR, into new symbols b+(k), b(k), also

satisfying CCR.

It is easy to prove that the set of canonical transformations

constitutes a group. This means that the composition of two canonical

transformations λ and µ produces a new canonical transformation λµ

and that the inverse of a canonical transformation is also a canonical

transformation.

It was shown above that the transformation from a+(k), a(k) to

the Heisenberg operators a+(k, t), a(k, t) is a canonical transforma-

tion; it therefore follows that any expression, H ∈ M defines a one-

parameter family of canonical transformations; we will denote these

transformations by σtH .

It is useful to note that the canonical transformation σtH can be

obtained as a limit of canonical transformations in finite volume.

Namely, using the HamiltonianHΩ, obtained as a finite-volume cutoff

of H, we can introduce the operators ak(t), a+k (t), where k ∈ TΩ, by

the formulas

ak(t) = exp(iHΩt)ak exp(−iHΩt),

a+k (t) = exp(iHΩt)a

+k exp(−iHΩt).

The operators a+k (t), ak(t), for fixed t, clearly satisfy CCR; the trans-

formation from the operators a+k , ak to the operators a+

k (t), ak(t)

specifies a canonical transformation in the volume Ω. Let us write

Page 332: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 285

Translation-Invariant Hamiltonians (Further Investigations) 285

the operators a+k (t), ak(t) in normal form:

ak(t) = exp(−iε(k)t)ak

+∑r≥1

gr∑

ki∈TΩpj∈TΩ

Ωf r,tm,n(k1, . . . ,km|p1, . . . ,pn)

× δk1+···+kmp1+···+pn

(2π

L

) 32

(m+n−2)

a+k1· · · a+

kmap1 · · · apn ,

a+k (t) = exp(iε(k)t)a+

k

+∑r≥1

gr∑

ki∈TΩpj∈TΩ

Ωfr,tm,n(k1, . . . ,km|p1, . . . ,pn)

× δk1+···+kmp1+···+pn

(2π

L

) 32

(m+n−2)

a+p1· · · a+

pnak1 · · · akm .

For the functions Ωf r,tm,n, just as for the functions f r,tm,n, we can derive

recurrent formulas from the Heisenberg equations. It is easy to check

that the recurrent formulas admit the limit Ω → ∞, and therefore,

it follows that

limΩ→∞

Ωf r,tm,n = f r,tm,n.

This equality is what we meant by the statement that the canonical

transformation σtH can be obtained as a limit of canonical transfor-

mations in finite volume.

Let the Hamiltonian H ∈ M be given by the formula (11.3) and

the canonical transformation λ by the relations (11.18) and (11.20).

Let us further replace in H the symbols a+(k), a(k) by the

expressions b+(k), b(k), defined by the formulas (11.18) and (11.20),

and bring the resulting formal expression to normal form using CCR,

while discarding infinite constants.

It is easy to check that we will obtain an expression H ∈ M; we

will say that the Hamiltonian H is obtained from the Hamiltonian

H by the canonical transformation λ, and write H = λ(H).

Page 333: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 286

286 Mathematical Foundations of Quantum Field Theory

Let us note that if the canonical transformation λ can be written

in the form λ = σtA, where A ∈M, then

H = λ(H) = limΩ→∞

λΩ(H),

where

λΩ(H) = exp(itAΩ)HΩ exp(−itAΩ)− cΩ

(the limit should be understood in the sense of the limits of the

coefficients of the functions in normal form of the operator λΩ(H) to

the coefficients of the functions in the expression λ(H)). The proof

of this follows from the fact that, as stated above, the canonical

transformation σtA can be obtained as a limit of canonical trans-

formations in finite volume; the constant cΩ arises from discarding

infinite constants after the canonical transformation; we can assume

that

cΩ = 〈exp(itAΩ)HΩ exp(−itAΩ)θ, θ〉 .

If we formally calculate the commutator [A,B] of two expressions

A,B ∈ M and multiply it by i, then we obtain a new expression

belonging to M. It is easy to check that the set M, equipped with

the operation i[A,B] and the obvious structure of vector space, is a

Lie algebra. The group of canonical transformations G can be seen

as an infinite-dimensional Lie group. It is possible to show that Mis a Lie algebra of the group G. The canonical transformations σtHclearly constitute a one-parameter subgroup of G, corresponding to

the element H of the Lie algebra M; the map taking the element

H ∈ M to the element σ1H is the exponential map from M to G.

Each element of the Lie group defines an automorphism of the

corresponding Lie algebra; in our case of interest, the canonical

transformation λ corresponds to the automorphism of the Lie algebra

M that maps the formal expression H to the formal expression λ(H).

11.3 Construction of an operator realization

Let us consider a Hamiltonian H =∑Hrg

r ∈ M that does not

generate vacuum polarization. The formal expressions Hr therefore

Page 334: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 287

Translation-Invariant Hamiltonians (Further Investigations) 287

define Hermitian operators Hr on Fock space with domain S∞ (see

Section 3.2, where it is shown that in the case of vacuum polarization,

this statement ceases to be true). In the framework of perturbation

theory, we can construct the operator exp(iHt) =∑grUr(t),

where the operators Ur(t) are defined on S∞ and transform S∞into itself. This statement can be proved by means of diagram

techniques described in Section 6.4; roughly speaking, the proof

depends on the remark that in the case at hand, all the diagrams

represent mathematically sensible expressions. (Before constructing

a perturbation theory series for exp(iHt), it is necessary to express

the Hamiltonian H in the form H ′0 + V ′, where

H ′0 = H0 +∑r≥1

gr∫v

(r)1,1(k)a+(k)a(k)dk,

and use the diagram techniques for the operator S(0, t) =

exp(−iH ′0t) exp(iHt).) Another proof is laid out in Section 11.5.

The operator generalized functions a(k, ε, t) =∑grar(k, ε, t) are

defined by the formula

a(k, ε, t) = exp(iHt)a(k, ε) exp(−iHt),

where a(k, 1) = a+(k), a(k,−1) = a(k) are the operators of

annihilation and creation in Fock space. Since the operators Ur(t)

and∫f(k)a(k, ε)dk transform the set S∞ into itself, the operators∫f(k)ar(k, ε, t)dk =

r∑s=0

Us(t)

(∫f(k)a(k, ε)dk

)Ur−s(−t)

inherit the property for any function f ∈ S.

It is clear that the energy operator H =∑Hrg

r, the momentum

operator P =∫ka+(k)a(k)dk, the operator generalized functions

a(k, ε, t) =∑grar(k, ε, t), and the vector Φ = θ constitute an

operator realization of the Hamiltonian H ∈M.

To construct an operator realization for a Hamiltonian generating

vacuum polarization, we will use the following statement.

Theorem 11.2. Let λ be a canonical transformation and let H =

λ(H) ∈ M be the Hamiltonian obtained from the Hamiltonian H

Page 335: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 288

288 Mathematical Foundations of Quantum Field Theory

by the canonical transformation λ. Then the operator realization of

the Hamiltonian H can be obtained from the operator realization

(H, H, P, a(k, ε, t),Φ) of the Hamiltonian H, by keeping the space

H, the operators of energy and momentum H, P, and the vector Φ

the same and replacing the operator generalized functions a(k, ε, t)

by the operator generalized functions

a(k,−1, t) = a(k, t) = α(k)a(k, t)

+∑r≥1

gr∑m,n

∫h(r)m,n(k1, . . . ,km|p1, . . . ,pn)

× δ

m∑i=1

ki −n∑j=1

pj − k

a+(k1, t)

· · · a+(km, t)a(p1, t) · · · a(pn, t)dmkdnp, (11.21)

a(k, 1, t) = a+(k, t) = α(k)a(k, t)

+∑r≥1

gr∑m,n

∫h

(r)m,n(k1, . . . ,km|p1, . . . ,pn)

× δ

m∑i=1

ki −n∑j=1

pj − k

a+(p1, t)

· · · a+(pn, t)a(k1, t) · · · a(kn, t)dmkdnp (11.22)

[we assume that the canonical transformation λ is given by the

formulas (11.18) and (11.20)].

To prove the formulated statement, let us note that the

only dependence on the concrete form of the translation-invariant

Hamiltonian in the definition of an operator realization is in the

condition that the operator generalized function a(k, ε, t) satisfies

the Heisenberg equations.

In order to check that this condition is satisfied, it is necessary to

differentiate equation (11.21) with respect to t, using the fact that

∂a(k, ε, t)

∂t= i[H, a(k, ε, t)]

Page 336: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 289

Translation-Invariant Hamiltonians (Further Investigations) 289

by the definition of the operator realization of the Hamiltonian H.

Expressing a in terms of a by the canonical transformation λ−1, we

obtain an equation for a(k, ε, t); simple but cumbersome calculations

allow us to bring the resulting equations to the form

∂a(k, ε, t)

∂t= i[H, a(k, ε, t)].

We will further consider the Hamiltonian H = H0 + V ∈ M,

assuming that the function ε(k), entering the definition of the

Hamiltonian H0, satisfies the inequality ε(k1 + k2) < ε(k1) + ε(k2).

With the help of the statement proven above, we can easily convert

the construction of the operator realization of this Hamiltonian to

the already-solved problem of construction of the operator realization

of a Hamiltonian that does not generate vacuum polarization. We

should use the following statement proven (in different terms) in

Section 9.4.

Theorem 11.3. Let the Hamiltonian H = H0 + V ∈ M be given

and suppose the function ε(k) satisfies the inequality ε(k1 + k2) <

ε(k1) + ε(k2). Then there exists a canonical transformation λ such

that the Hamiltonian H = λ(H) ∈M can be represented in the form

H = H0 + V , (11.23)

where

V =∑r

gr∑m,n

∫v(r)m,n(k1, . . . ,km|p1, . . . ,pn)

× δ

m∑i=1

ki −n∑j=1

pj

a+(k1) · · · a+(km)a(p1)

· · · a(pn)dmkdnp,

V does not contain summands of type (m, 0) or summands of type

(m′, 1) with m′ 6= 1(

i.e. v(r)m,0 = 0, v

(r)m′,1 = 0 for m′ 6= 1

).

The transformation λ was previously called a Faddeev transfor-

mation; let us recall that it was constructed in the form σ1A, where

A ∈M.

Page 337: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 290

290 Mathematical Foundations of Quantum Field Theory

One can say that the Faddeev canonical transformation transforms

the Hamiltonian H into the Hamiltonian H, for which the bare

vacuum coincides with the dressed vacuum and the bare single-

particle state coincides with the dressed single-particle state.

Let us now show that any two operator realizations of a

Hamiltonian H ∈ M are unitarily equivalent. With the help of

Faddeev transformation, we can reduce the problem to the case

when the Hamiltonian does not generate vacuum polarization [recall

that above we also assumed the condition ε(k1 + k2) < ε(k1) +

ε(k2)]. Indeed, from two unitarily equivalent operator realizations

of the Hamiltonian H by means of the Faddeev transformation,

we can obtain two unitarily equivalent operator realizations of the

Hamiltonian not generating vacuum polarization.

In order to prove the uniqueness of the operator realization in

the case when the Hamiltonian H ∈ M does not generate vacuum

polarization, it is enough to check that for any operator realiza-

tion (H′, H ′,P′, a′(k, ε, t),Φ′) of the Hamiltonian H, the relation

a′(k,−1, t)Φ′ = 0 is satisfied (then the unitary operator U estab-

lishing the unitary equivalence of this operator realization and the

operator realization constructed above can be defined as the operator

satisfying the condition Ua(k, ε, 0) = a′(k, ε, 0)U,UΦ = Φ′).4 Let

us now check the relation a′(k,−1, t)Φ′ = (∑grar(k,−1, t))Φ′ = 0.

Consider the equations for a′r(k,−1, t) following from the Heisenberg

equations. It is easy to check that they have the form

i∂a′r(k,−1, t)

∂t− ε(k)a′r(k,−1, t) = Fr, (11.24)

where Fr is an expression composed of a′s(k, 1, t) and a′s(k,−1, t)

with s < r. In order to prove the necessary relation, we should show

that a′r(k,−1, t)Φ′ = 0; we will perform the proof by induction in r.

4The existence of the operator U satisfying these conditions follows from theresults of Section 6.1 (more precisely, from their analogs concerning the unitaryequivalence of the Fock representations of CCR, written in the form of a formalexpression in powers of g). Using the Heisenberg equations, we can show thatUa(k, ε, t) = a′(k, ε, t)U . From the condition (b) in the definition of operatorrealization, we obtain that UH = H ′U,UP = P′U .

Page 338: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 291

Translation-Invariant Hamiltonians (Further Investigations) 291

Suppose that a′s(k,−1, t)Φ′ = 0 for s < r. By the absence of vacuum

polarization, every summand in the expression Fr will contain at

least one operator of the form as(k,−1, t) and therefore FrΦ′ = 0.

Using equation (11.24) on the vector Φ, we can see that

i∂

∂ta′r(k,−1, t)Φ′ = ε(k)a′r(k,−1, t)Φ′,

from which it follows that

a′r(k,−1, t)Φ′ = exp(−iε(k)t)a′r(k,−1, 0)Φ′.

Let us now calculate the expectation value of the energy in the state∫f(k)a′(k,−1, t)dkΦ′:⟨H ′∫f(k)a′(k,−1, t)dkΦ′,

∫f(k)a′(k,−1, t)dkΦ′

⟩=

⟨1

i

∫f(k)

∂ta′(k,−1, t)dkΦ′,

∫f(k)a′(k,−1, t)dkΦ′

⟩=∑s≥r

1

i

∫f(k)f(k′)

⟨∂

∂ta′s(k,−1, t)Φ′, a′s′(k

′,−1, t)Φ′⟩dkdk′

= C2rg2r + C2r+1g

2r+1 + · · · ,

where

C2r =1

i

∫f(k)f(k′)

⟨∂

∂ta′r(k, 1, t)Φ

′, a′r(k′,−1, t)Φ′

⟩dkdk′

= −∫ε(k)f(k)f(k′) exp(−i(ε(k)− ε(k′))t)

×⟨a′r(k,−1, 0)Φ′, a′r(k

′,−1, 0)Φ′⟩dkdk′.

Since ⟨a′r(k,−1, 0)Φ′, a′r(k

′,−1, 0)Φ′⟩

= σ(k)δ(k− k′),

where σ(k) ≥ 0, we see that

C2r = −∫ε(k)|f(k)|2σ(k)dk ≤ 0.

Page 339: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 292

292 Mathematical Foundations of Quantum Field Theory

If a′r(k,−1, t)Φ′ 6≡ 0, then σ(k) 6≡ 0, and therefore, we can find a

function f(k) such that C2n < 0. This contradicts the non-negativity

assumption of the operator H ′.

Let us now check the applicability conditions of axiomatic field

theory in each power of g. Let us first consider the spectrum of the

operators H, P in the operator realization (H, H, P, a(k, ε, t),Φ) of

a Hamiltonian of the form (11.23). For this Hamiltonian, the (bare)

Fock vacuum θ and the bare single-particle state a+(k)θ satisfy the

conditions

Hθ = Pθ = 0,

Ha+(k)θ =(ε(k) +

∑grv

(r)11 (k)

)a+(k)θ,

Pa+(k)θ = ka+(k)θ.

Therefore, θ can be viewed as the ground state of the operator H

(dressed vacuum) and a+(k)θ can be viewed as a single-particle state

of operators H, P (dressed single-particle state). The joint spectrum

of the operators H, P consists of the point 0, corresponding to the

ground state, the set of points (ω(k),k) = (ε(k) +∑grvr1,1(k),k)

(single-particle spectrum), and the multi-particle spectrum, whose

points have the form (ω+∑

r≥1 grωr,k+

∑r≥1 g

rkr), where the point

(ω,k) belongs to the multi-particle spectrum of the operators H0, P

(i.e. can be presented in the form (ε(k1)+ · · ·+ε(km),k1 + · · ·+km)).

If the Hamiltonian H belongs to the class M [i.e. the function ε(k)

is not linear on any open set and satisfies the condition ε(k1 +k2) <

ε(k1) + ε(k2)], then the spectrum of the operators H, P satisfies the

conditions formulated in Section 10.5.

In order to finish the exposition of the case at hand, it remains

to be checked that the family B of operator generalized functions,

consisting of two functions

a(x, ε, t) = (2π)−3/2

∫exp(iεkx)a(k, ε, t)dk, ε = ±1,

is asymptotically commutative, in the sense of Section 10.5, in every

order of perturbation theory with respect to g. In order to show that

B is asymptotically commutative, let us note that the operators of

Page 340: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 293

Translation-Invariant Hamiltonians (Further Investigations) 293

the form∫f(x)a(x, ε, t)dx that we have considered belong to the

algebra A, consisting of operators of the form∑gr∑m,n

∫f (r)m,n(k1, . . . ,km|p1, . . . ,pn)a+(k1)

· · · a+(km)a(p1) · · · a(pn)dmkdnp,

where the functions f(r)m,n ∈ S and for every r, only finitely many

functions f(r)m,n are non-zero. It is easy to check (using the results

of Section 11.2, specifically inequality (11.15)) that the algebra Ais asymptotically commutative in every order in g. (This statement

also follows from the results of Section 11.5.)

Therefore, the family B is asymptotically commutative.

Let us return to considering an arbitrary Hamiltonian H =

H0 + V ∈ M. Using the operator realization of this Hamiltonian

constructed by means of Faddeev transformation, we can obtain

an operator realization of the Hamiltonian H of the form (11.23).

In these constructions, the energy and momentum operators in the

operator realization of the Hamiltonian H coincide with the energy

and momentum operators in the operator realization of the Hamilto-

nian H; this remark implies that they satisfy the necessary spectrum

conditions. Operator generalized functions a(k, ε, t), entering in

the operator realization of the Hamiltonian H, are related to the

operator generalized functions a(k, ε, t) in the operator realization

of the Hamiltonian H through the canonical transformation (see

the relation (11.21)). It therefore follows that the family B of two

operator generalized functions a(x, ε, t), ε = ±1, with which we

may construct the scattering matrix for the Hamiltonian H, is an

asymptotically commutative algebra in every order of g, and there-

fore the operators∫f(x)a(x, ε, t)dx belong to the asymptotically

commutative family A.

These statements imply that the scattering matrix of the Hamil-

tonian H ∈ M can be constructed via axiomatic scattering theory in

terms of formal series in g.

It is now possible to apply axiomatic scattering theory to study

(in the framework of perturbation theory) scattering matrices in the

Hamiltonian formalism.

Page 341: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 294

294 Mathematical Foundations of Quantum Field Theory

First, let us note that the discussions imply that the scattering

matrix of the Hamiltonian H coincides with the scattering matrix of

the Hamiltonian H, obtained from H by means of Faddeev canonical

transformation.5 Indeed, these two scattering matrices are built on

the asymptotically commutative families B and B, which are both

contained in the same asymptotically commutative algebra A; by

the results of Chapter 10, the Møller and scattering matrices based

on these families are identical.

The statement we have just proved implies the following equiva-

lence theorem.

If the Hamiltonians H1 ∈ M and H2 ∈ M are related by a

canonical transformation, then their Møller and scattering matrices

coincide.

Indeed, let α be a canonical transformation transforming the

Hamiltonian H1 to the Hamiltonian H2; let β be the Faddeev

canonical transformation transforming the Hamiltonian H2 to the

Hamiltonian H of the form (11.23). Then the canonical transforma-

tion αβ transforms the Hamiltonian H1 to the Hamiltonian H.

The scattering matrices of the Hamiltonians H1 and H2 above

coincide with the scattering matrices of the Hamiltonian H and,

therefore, coincide with each other. The same reasoning applies to

Møller matrices.

11.4 Dressing operators for translation-invariant

Hamiltonians

In this section, we will consider translation-invariant Hamiltonians

from class M.

Let us first examine the case where the Hamiltonian H ∈ Mdoes not generate vacuum polarization. The dressing operator for

5It is important to note that in this statement in place of the Faddeevtransformation, we can select any canonical transformation transforming theHamiltonian H into a Hamiltonian of the form (11.23), not necessarily theconcrete transformation constructed in Section 9.4.

Page 342: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 295

Translation-Invariant Hamiltonians (Further Investigations) 295

this Hamiltonian is defined as an operator satisfying the condition

S± = slimt→±∞

exp(iHt)D exp(−iHast). (11.25)

(The Hamiltonian in consideration defines the operator H on Fock

space H = F (L2(E3))), the asymptotic space Has = F (L2(E3)) is

naturally identified with the space H, therefore, the Møller matrices

S± act on the space H. The symbol Has denotes the asymptotic

Hamiltonian which we also view as an operator on H:

Has =

∫ω(k)a+(k)a(k)dk,

where ω(k) is the energy of the single-particle state of the Hamilto-

nian H. All operators in the formula (11.25), just as for the rest of

this section, are viewed as formal series in g. The limit in (11.25) is

understood as a strong limit in each perturbation order on the set of

linear combinations of vectors of the form a+(f1) · · · a+(fn)θ, where

f1(k), . . . , fn(k) are non-overlapping smooth functions with compact

support.)

The following theorem describes a broad class of dressing

operators for Hamiltonians H ∈ M that do not generate vacuum

polarization.

Theorem 11.4. Let D be an operator on Fock space H = F (L2(E3)),

satisfying the following conditions:

(1) Dθ = θ; (11.26)

(2) the vector generalized function Da+(k)θ is δ-normalized and is

an eigenvector for the operator H;

(3) the operator D can be represented as a product of operators of

the form expA or N(expA), where A = A1 + iA2, A1 ∈ M,

A2 ∈M.

Then the operator D is a dressing operator for the Hamiltonian H.

(The symbol N(expA) denotes the normal exponent; see Section 6.3.)

This theorem can be proved with the help of description of

dressing operators in axiomatic scattering theory (see Section 10.4).

In the following section, we present a proof of this theorem in the

Page 343: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 296

296 Mathematical Foundations of Quantum Field Theory

case when the operator D is represented in the form of a product of

operators of the form expA, where A ∈ M. In the section at hand,

however, we will provide a different proof of the formulated theorem.

To begin, let us consider the case where D = N(expA), where

A =∑m

∫Am,1(k1, . . . ,km)δ(k1 + · · ·+ km − p)

× a+(k1) · · · a+(km)a(p)dk1 · · · dkmdp.

Let us first note that in this case, we have the relation

[D, a+(k)] = B(k)D, (11.27)

where B(k) is a generalized operator function defined by the formula

B(k) =∑m

∫Am,1(k1, . . . ,km)δ(k− k1 − · · · − km)

× a+(k1) · · · a+(km)dmk

(the relation (11.27) follows immediately from the equality

[N(An), a+(k)] = B(k) · nN(An−1),

which we obtained using the formula

[R1 · · ·Rm, R] =∑

1≤i≤mR1 · · ·Ri−1[Ri, R]Ri+1 · · ·Rm).

Let us rewrite relation (11.27) in the form

Da+(k) = E(k)D, (11.28)

where E(k) = a+(k) +B(k). Using formulas (11.28) and (11.26), we

see that

Da+(k1) · · · a+(kn)θ = E(k1) · · ·E(kn)θ,

in particular

Da+(k)θ = E(k)θ

= a+(k) +∑m

∫Am,1(k1, . . . ,km)

× δ(k− k1 − · · · − km)a+(k1) · · · a+(km)dmk. (11.29)

Page 344: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 297

Translation-Invariant Hamiltonians (Further Investigations) 297

It follows from relation (11.29) that

exp(iHt)D exp(−iH0t)a+(f1) · · · a+(fn)θ

= exp(iHt)Da+(f t1) · · · a+(f tn)θ

= exp(iHt)

(∫f t1(k1)E(k1)dk1

)· · ·(∫

f tn(kn)E(kn)dkn

=

(∫f t1(k1)E(k1, t)dk1

)· · ·(∫

f tn(kn)E(kn, t)dkn

(here, f t(k) = exp(−itω(k))f(k)).

To check that D is a dressing operator, it is sufficient to establish

that E is a good operator generalized function in the sense of

Section 10.5; this can be easily done following the analysis of

Heisenberg operators a(k, ε, t) provided in Section 11.2.

Let us remark that knowing the single-particle state Φ(k) =

Da+(k)θ and using the formula (11.29), we can find the function

Am,1 and the rest of the dressing operator of the considered type.

From this remark, it is easy to derive that the condition that the

operator A belongs to the class M + iM, for the case at hand, is

unnecessary (it is necessary to use the following expression of the

single-particle state in terms of perturbation theory

Φ(k) = Φ(k|g)

= exp(iσ(k|g))

(a+(k) +

∑m

∫φm(k1, . . . ,km)

× δ(k− k1 − · · · − km)a+(k1) · · · a+(km)dmk

)θ,

where exp(iσ(k|g)) is an arbitrary phase factor; φm(k1, . . . ,km) =∑m φ

(r)m (k1, . . . ,km)gr; the functions φ

(r)m belong to the space S, and

for every r only finitely many of these functions are non-zero).

The proven statement allows us to conclude that the definition of

scattering matrix given in Section 9.1 (formula (9.10)) is equivalent

to the other definitions.

Let us now return to the slightly more general case where D =

N(expA), where A ∈M+iM (i.e. A = A1+iA2, A1 ∈M, A2 ∈M).

Page 345: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 298

298 Mathematical Foundations of Quantum Field Theory

We will write A in the form A =∑Am,n, where Am,n =

∑r A

(r)m,ngr,

A(r)m,n =

∫A(r)m,n(k1, . . . ,km|p1, . . . ,pn)

× δ(k1 + · · ·+ km − p1 − · · · − pn)a+(k1)

· · · a+(km)a(p1) · · · a(pn)dmkdnp,

and define the operator A′ in the same way as the operator A except

only keeping terms of type (m, 1) (i.e. we set A′ =∑

mAm,1).

The operator D′ = N(expA′) satisfies the conditions

D′θ = θ,

D′a+(k)θ = Da+(k)θ

(terms containing no less than two annihilation operators go to zero

when acting on a+(k)θ).

As was shown above, the operator D′ = N(expA′) is a dressing

operator, i.e.

slimt→±∞

exp(iHt)D′ exp(−iHast) = S±. (11.30)

In order to derive from formula (11.30), the necessary relation

slimt→±∞

exp(iHt)D exp(−iHast) = S±,

it is enough to show that

slimt→±∞

exp(iHt)(D −D′) exp(−iHast) = 0. (11.31)

The proof of equation (11.31) in each order of perturbation theory

immediately follows from the relation

limt→±∞

N(A(r1)m1n1

· · ·A(rs)msns

)a+(f t1)· · · a+

(f ts)θ = 0,

where at least one of the operators A(ri)mini satisfies the condition

ni ≥ 2 (i.e. contains at least two annihilation operators), and the

functions f1(k), . . . , fs(k) constitute a non-overlapping family of

smooth functions with compact support. The above completes the

proof of the theorem in the case when the operator D takes the

Page 346: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 299

Translation-Invariant Hamiltonians (Further Investigations) 299

form N(expA). The proof of the theorem in the general case can

be reduced to the cases already considered using the following two

lemmas.

Lemma 11.1. Any operator of the form exp(A1 + iA2), where

A1, A2 ∈ M, can be written in the form N(exp(A′1 + iA′2)), where

A′1, A′2 ∈M.

Lemma 11.2. Any operator of the form exp(A1 + iA2)× exp(A′1 +

iA′2), where A1, A2, A′1, A

′2 ∈ M, can be written in the form

N(exp(A′′1 + iA′′2)), where A′′1, A′′2 ∈M.

We will not provide proofs of these lemmas.

Let us now consider the case of an arbitrary Hamiltonian H ∈M.

The symbol HΩ, as usual, denotes the operator H with finite volume

cutoff (see Section 8.1); the operator HΩ acts on the Fock space

FΩ = F (L2(Ω)). In perturbation theory, we can calculate the

operator exp(iHΩt) (in other words, HΩ can be viewed as a self-

adjoint operator in the framework of perturbation theory).

The family of operators DΩ will be called a family of dressing

operators for the Hamiltonian H if the scattering matrix S of the

Hamiltonian H can be written in the form

S = limt→∞t0→−∞

limΩ→∞

exp(i(cΩ +HasΩ)t)(DΩ)−1

× exp(−iHΩ(t− t0))DΩ exp(−i(cΩ +HasΩ)t0). (11.32)

(The operators DΩ act on the space FΩ, the symbols HasΩ denote the

asymptotic Hamiltonian Has with volume cutoff, and cΩ is a constant

depending on the volume. The limit in (11.32) can be understood in

the sense of the convergence of matrix elements (see Section 9.3) or

via the relation

S = slimt→∞t0→−∞

slimΩ→∞

iΩ exp(i(cΩ +HasΩ)t)(DΩ)−1

× exp(iHΩ(t− t0))DΩ exp(−i(cΩ +HasΩ)t0)i∗Ω,

where iΩ is a natural embedding of the Fock space FΩ = F (L2(Ω))

into the Fock space F (L2(E3)).)

Page 347: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 300

300 Mathematical Foundations of Quantum Field Theory

Let us sketch the proof of the following theorem.

Theorem 11.5. Let the family of operators DΩ satisfy the following

conditions:

(1) The operator DΩ transforms the Fock vector θ and the bare

single-particle state a+k θ into normalized eigenvectors of the

operator HΩ.

(2) The operator DΩ can be represented as the product of opera-

tors of the form N(expAΩ) and expAΩ, where AΩ are oper-

ators obtained from a volume cutoff of operators of the form

A = A1 + iA2 and A1, A2 are formal expressions from the class

M, not depending on Ω.

Then DΩ is a family of dressing operators. (More precisely, it

can be shown that formula (11.32) holds, where cΩ is defined by the

formula HΩDΩθ = cΩD

Ωθ.)

Let us begin with the remark that with the help of analogs

of Lemmas 11.1 and 11.2 for operators on the space FΩ, we can

reduce the proof to the case when the operator DΩ has the form

N(expAΩ), where A = A1 + iA2, A1 ∈ M, A2 ∈ M. Let us

first suppose that the Hamiltonian H does not generate vacuum

polarization. For such a Hamiltonian, the operators exp(iHΩt), DΩ =

N(expAΩ), exp(iHasΩt) have limits as Ω → ∞; they converge,

respectively, to the operators exp(iHt), N(expA) = D, exp(iHast).

This can be derived from the relations Hθ = 0, Hasθ = 0, Aθ = 0

(the last of these relations follows from the equation DΩθ = θ). It

is easy to check that on the right-hand side of equation (11.32), we

can transform the limit of products into a product of limits. Noting

that the constant cΩ in the case at hand is equal to 0, we obtain that

equation (11.32) is equivalent to the relation

S = slimt→∞t0→−∞

exp(iHast)D−1 exp(−iH(t− t0))

×D exp(−iHast0),

which follows from the already-proven relations (11.25) and the

formula S = S∗+S−. Thus, the necessary statement is proven for

Page 348: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 301

Translation-Invariant Hamiltonians (Further Investigations) 301

Hamiltonians not generating vacuum polarization. The case of an

arbitrary Hamiltonian H ∈ M reduces to the considered case by

Faddeev transformation. Indeed, by the results of Section 9.4, there

exists a formal expression W ∈M, such that as Ω→∞, we have

exp(iWΩ)HΩ exp(−iWΩ) ≈ H ′Ω + cΩ,

and, therefore,

exp(iWΩ)eiHΩt exp(−iWΩ) ≈ ei(H′Ω+cΩ)t (11.33)

where H ′Ω is obtained from a Hamiltonian H ′ that does not generate

vacuum polarization by means of volume cutoff; cΩ is the energy

of the ground state of the operator HΩ. Using formula (11.33), we

conclude that

limΩ→∞

exp(i(cΩ +HasΩ)t)(DΩ)−1 exp(−iHΩ(t− t0))

×DΩ exp(−i(cΩ +HasΩ)t0)

= limΩ→∞

exp(iH ′asΩt′)(D′Ω)−1 exp(−iH ′Ω(t− t0))

×D′Ω exp(−iH ′asΩt0), (11.34)

where

D′Ω = exp(iWΩ)DΩ

(to derive equation (11.34), one needs to use the coincidence of

the asymptotic Hamiltonians Has and H ′as, corresponding to the

Hamiltonians H and H ′). Applying the statement of the theorem

to a Hamiltonian H ′ not generating vacuum polarization, we see

that the operators D′Ω constitute a family of dressing operators

for the Hamiltonian H ′. This implies that the right-hand side of

equation (11.34), in the limit as t → ∞, t0 → −∞, equals

the scattering matrix of the Hamiltonian H ′. Since the scattering

matrices of the Hamiltonians H and H ′ coincide, equation (11.34)

proves the theorem in the general case.

11.5 Perturbation theory via the axiomatic approach

Let us suppose that on the Hilbert space H, we have defined

an energy operator H0, a momentum operator P, a linear dense

Page 349: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 302

302 Mathematical Foundations of Quantum Field Theory

subspace D of H, and a complete locally convex topological algebra

with involution A, consisting of operators acting on D. Let us

further assume that these objects satisfy conditions T1–T7 from

Section 10.5.

Suppose that W is an operator from the algebra A, satisfying the

condition WΦ = 0. Let us now prove that the family of operators

H(g) = H0 + gV = H0 + g

∫W (x)dx

together with momentum operator P and algebra A, satisfy conditions

T1–T7 from Section 10.5 in each power of the parameter g (the

integral is understood in the strong sense). Imposing some conditions

on the joint spectrum of the operators H0,P, we will prove that

starting with the operators H(g),P, and the algebra A, one can

construct the Møller matrices S±(g) as well as the scattering matrix

S(g), at least as formal series in powers of g.6

Let us first note that the operator V ∈∫W (x)dx is defined on

the set AΦ (i.e. on the set of vectors of the form AΦ, where A ∈ A)

and transforms this set into itself. Indeed,(∫W (x)dx

)AΦ =

∫[W (x), A]Φdx.

By condition T6, it follows that for any seminorm p(A) on A and

any n, we have

p([W (x), A]) ≤ C

1 + |x|n. (11.35)

By the completeness of the algebra A and equation (11.35), the

integral∫

[W (x), A]dx converges in the topology of A and defines

an element of the algebra A. This completes the proof.

Later in the chapter, we will prove that when calculating the

operator exp(−iH(g)t) = exp(−i(H0+gV )t) in terms of perturbation

6This statement, just as the rest of the statements in this section, applies tothe more general class of Hamiltonians of the form H0 +

∑r≥1 g

r∫Wr(x)dx, and

Wr ∈ A,WrΦ = 0, and the series in g is understood as a formal series.

Page 350: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 303

Translation-Invariant Hamiltonians (Further Investigations) 303

theory in each order of g, we obtain an operator, defined on the set

AΦ, that transforms this set into itself (this implies that the operator

H(g) can be seen as a self-adjoint operator, at least in perturbation

theory). Indeed, from the results of Section 4.1, it follows that

exp(−iH(g)t) =∑(g

i

)rUr(t),

where

Ur(t) =1

r!exp(−iH0t)

∫ t

0dτ1

∫ τ1

0dτ2 · · ·

∫ τr−1

0dτrV (τ1) · · ·V (τr)

=1

r!exp(−iH0t)

∫ t

0dτ1

∫ τ1

0dτ2 · · ·

∫ τr−1

0dτr

∫drxW (x1, τ1)

· · · W (xr, τr),

V (τ) = exp(iH0τ)V exp(−iH0τ) =

∫W (x, τ)dx,

W (x, τ) = exp(i(H0τ −Px))W exp(−i(H0τ −Px)).

Let us define the operator Rr(x1, t1, . . . ,xr, tr|A), where A ∈ A,

by induction and the relations

R0(A) = A,

Rr(x1, t1, . . . ,xr, tr|A)

= [W (x1, t1), Rr−1(x2, t2, . . . ,xr, tr|A)].

It is easy to check that the operator Rr continuously depends on

x1, t1, . . . ,xr, tr in the topology of the algebra A and

p(Rr(x1, t1, . . . ,xr, tr|A)) ≤ C(1 + |t1|s + · · ·+ |tr|s)1 + |x1|n + · · ·+ |xr|n

(11.36)

(here, p(A) is any seminorm on A, n is an arbitrary number and the

numbers s and C depend on the seminorm p and the number n).

Inequality (11.36) can be easily derived by induction and by means

of conditions T5 and T6.

Page 351: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 304

304 Mathematical Foundations of Quantum Field Theory

Let us now note that

W (x1, t1) · · ·W (xr, tr)AΦ = Rr(x1, t1, . . . ,xr, tr|A)Φ

(this can also be easily checked by induction). Thus,

Ur(t)AΦ = Br(t)Φ,

where

Br(t) =1

r!

∫ t

0dτ1

∫ τ1

0dτ2 · · ·

∫ τr−1

0dτr

∫drx exp(−iH0t)

×Rr(x1, τ1, . . . ,xr, τr|A) exp(iH0t).

From the inequality (11.36) and the completeness of algebra A,

it follows that for any A ∈ A, the integral defining the operator

Br(t) converges in the topology of A and, therefore, Br(t) ∈ A.

Thus, the operator Ur(t) is defined on the set AΦ and transforms

the set into itself. It therefore follows that the operator A(x, t|g) =

exp(iH(g)t − iPx)A exp(−iH(g)t + iPx), in each order of the

perturbation series, is defined on the set AΦ and then transforms

the set into itself:

A(x, t|g) =∑

grAr(x, t),

where the operator

Ar(x, t) =r∑

α=0

Uα(−t)A(x)Ur−α(t)

(1

i

)racts on the set AΦ).

Let us now show that the operator Ar(x, t) ∈ A is continuous

in x, t in the topology on A and that for any seminorm p(A) on A,

we have

p(Ar(x, t)) ≤ (1 + |x|s + |t|s)q(A), (11.37)

where the number s and the seminorm q(A) on A depend on the

seminorm p and the number r; A runs over a compact set F ⊂ A(in other words, we will prove that condition T5 from Section 10.5

holds for operators H(g),P, and the algebra A in each order of the

Page 352: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 305

Translation-Invariant Hamiltonians (Further Investigations) 305

perturbation theory in g). To do this, let us note that from the

Heisenberg equation,

1

i

∂A(x, t|g)

∂t= [H0 + gV,A(x, t|g)],

follows the relation

1

i

∂Ar(x, t)

∂t= [H0, Ar(x, t)] + [V,Ar−1(x, t)]. (11.38)

With the relation (11.38) and the initial condition Ar(x, 0) = 0 for

r ≥ 1, we can express Ar(x, t) through Ar−1(x, t):

Ar(x, t) = exp(iH0t)Ar(x, t) exp(−iH0t),

Ar(x, t) = i

∫ t

0exp(−iHτ)[V,Ar−1(x, τ)] exp(iHτ)dτ

= i

∫ t

0dτ

∫dξ exp(−iHτ)[W (ξ), Ar−1(x, τ)] exp(iHτ).

(11.39)

Using this expression, the necessary statement follows by induction.

We obtain

p(exp(−iHτ)[W (ξ), Ar−1(x, τ)] exp(iHτ)) ≤ p1(A)(1 + |x|2 + |t|s)1 + |ξ|n

,

(11.40)

where A ∈ F , p is an arbitrary seminorm, n is an arbitrary number,

and the seminorm p1 and the number s depend on p and n.

Inequality (11.40) implies that the integral in (11.39) converges

uniformly in x, t in the topology on A if x, t belongs to a bounded

set. From this remark, it follows that Ar(x, t) ∈ A and that it con-

tinuously depends on x, t; the inequality (11.37) follows immediately

from the inequality (11.40).

It is now easy to prove that the operators H(g),P, and the algebra

A, whose elements take the form D′ = AΦ, satisfy conditions T1–T7

in Section 10.5 in each order of the perturbation series in g. Indeed,

we have already shown that the set D′ is invariant in each order of

perturbation expansion of the operators exp(iH(g)t). Its invariance

Page 353: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 306

306 Mathematical Foundations of Quantum Field Theory

with respect to the operators exp(iPx) follows from the relation

exp(−iPx)AΦ = A(x)Φ. Conditions T1–T4, T6, and T7 do not

depend on the choice of the energy operator H(g) and, therefore,

are satisfied in the case at hand. We have already checked condition

T5 by perturbation theory.

Let us now show that the joint spectrum of the operators H(g),P

satisfies the conditions on the spectrum of energy and momentum

operators specified in Section 9.5 (at least in perturbation theory).

Let us suppose that the necessary conditions7 are satisfied for

g = 0 and let us further assume that for any point k ∈ E3, we can find

an operator A ∈ A, such that | 〈Φ0(k), AΦ〉 | 6= 0 (the symbol Φ0(k)

denotes a single-particle state of the operator H0). The arguments

applied in the slightly more complicated situation in the proof of

Lemmas 10.12–10.14 in Section 10.6, allow us to show that the single-

particle state Φ(k) can be chosen such that the functions 〈Φ(k), AΦ〉for all A ∈ A will be infinitely differentiable. If the single-particle

state Φ(k) is chosen this way, then for any smooth function with

compact support, we can find a good operator Bf such that BfΦ =∫f(k)Φ(k)dk.

It follows from WΦ = 0 that the ground state Φ of the operator

H0 is a stationary state of the operator H(g) for any g; this implies

that when we calculate the ground state of the operator H(g) in

perturbation theory, we obtain the vector Φ. We can find single-

particle states of the operator H(g) in perturbation theory in the

form

Φ(k|g) =∞∑r=0

grΦr(k),

using the relations

(H0 + gV )(∑

grΦr(k))

=(∑

grωr(k))(∑

grΦr(k)), (11.41)

7In order to prove the existence of Møller matrices and scattering matricesconstructed with the operators H(g),P, and the algebra A, it is enough to assumethat the joint spectrum of the operators H0,P satisfies conditions 4 and 5 fromSection 10.5.

Page 354: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 307

Translation-Invariant Hamiltonians (Further Investigations) 307

P(∑

grΦr(k))

= k(∑

grΦr(k)), (11.42)⟨

∂g

∑grΦr(k),

∑grΦr(k

′)

⟩= 0. (11.43)

By the relations (11.41) and (11.42), we obtain that∫ωr(k)f(k)f ′(k)dk

=⟨V Φr−1(f),Φ0(f ′)

⟩−⟨Φr−1(ω1f),Φ0(f ′)

⟩− · · ·

−⟨Φ1(ωr−1f),Φ0(f ′)

⟩, (11.44)

Φr(f) = γ(H0,P)(−V Φr−1(f) + Φr−1(ω1f) + · · ·+ Φ0(ωrf))

+ Φ0(λrf), (11.45)

where f(k) is a smooth function with compact support; Φr(f) =∫f(k)Φr(k); γ(ω,k) is a smooth function with compact support

equal to (ω−ω0(k))−1, if k ∈ supp f and (ω,k) belongs to the multi-

particle spectrum of the operators H,P and also if k = 0, ω = 0. It

should be equal to zero if ω = ω0(k). Using relations (11.43)–(11.45),

we can find functions λr(k), ωr(k), and vectors Φr(k); simultaneously

by induction on r, we can prove that ωr(k) and λr(k) are smooth

functions and the vector Φr(f) can be written in the form

Φr(f) = B(r)f Φ, (11.46)

where B(r)f ∈ A. (The representation (11.46) follows from equa-

tion (11.45) with the help of the reasoning employed in the proof

of Lemma 10.17 in Section 10.6; the infinite differentiability of ωr(k)

and λr(k) can be shown by a slight modification of the proof of

Lemma 10.15 in Section 10.6.)

Let us now show that the Møller matrices S±(g), corresponding

to the operators H(g),P, and the algebra A, can be constructed at

least in the framework of perturbation theory. More precisely, we will

construct the Møller matrix as a formal series

S±(g) =∑

grS(r)± ,

Page 355: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 308

308 Mathematical Foundations of Quantum Field Theory

where S(r)± are operators defined on the set D ⊂ Has, consisting of

linear combinations of vectors of the form a+(f1) · · · a+(fn)θ (here,

f1(k), . . . , fn(k) is a non-overlapping family of smooth functions with

compact support. (Recall that this condition on a family of functions

means that for any k ∈ supp fi,k′ ∈ supp fj , i 6= j, the gradients of

the functions ω0(k) at the points k and k′ do not coincide.)

To construct Møller matrices, we repeat the construction in

Section 10.1, considering all objects, entering into the formula (10.1),

as formal series in powers of g. The proof of the correctness of

this definition is nearly identical to the proof in Chapter 10. It is

important to note that the estimate (10.7) cannot be proven in every

order in g, in other words, if f(k) is a smooth function with compact

support with

f(x|t) = (2π)−3

∫exp

−it∑r≥0

grωr(k) + ikx

f(k)dk

=∑r≥0

grfr(x|t),

then we cannot prove the inequality

|fr(x|t) ≤ C|t|−3/2].

Therefore, we also cannot prove that the Lemma 10.4 of Section 10.1

holds in each order of perturbation theory. However, in the proof

of the correctness of the definition of the Møller matrix relying

on formula (10.1), we can substitute this lemma with Lemma 10.5

from Section 10.3, which holds in each order in g (the family of

functions f1, . . . , fn in formula (10.1) should be assumed to be a

non-overlapping family of smooth functions with compact support).

The scattering matrix S is defined, as always, by the relation

S = S∗+S−. (From what we have proved, it does not follow that S

has the form S =∑grS(r), where S(r) are operators with a common

domain, however, the matrix elements⟨Sa+(k1) · · · a+(km)θ, a+(p1) · · · a+(pn)θ

⟩=⟨S−a

+(k1) · · · a+(km)θ, S+a+(p1) · · · a+(pn)θ

⟩(11.47)

Page 356: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 309

Translation-Invariant Hamiltonians (Further Investigations) 309

for ki 6=kj ,pi 6=pj can be viewed as formal series in g. More

precisely, these matrix elements can be written in the form∑

r≥0 gr

S(r)(k1, . . . ,km |p1, . . . ,pn), where S(r) are generalized functions,

defined on the class of test functions φ(k1, . . . ,km |p1, . . . ,pn) ∈S, having support belonging to the set where ki 6= kj ,pi 6=pj . Hence, by scattering matrix, we will mean the set of matrix

elements (11.47).)

We can now use the results of Chapter 10 to investigate the Møller

matrices S±(g) and the scattering matrix S(g), corresponding to the

operators H(g),P, and the algebra A. Let us note that, in particular,

with the help of the statements proven above, we can easily prove

the assumptions in the adiabatic theorem from Section 10.6 in each

order of g. Unfortunately, this is not enough to prove, in each order of

perturbation theory, the theorem that gives an expression of Møller

matrix S±(g) in terms of the adiabatic Møller matrix Sα(0,±∞|g).

Inspection of the proof of the adiabatic theorem shows that the proof

depends on relations of the form∣∣∣∣exp

(i

αν(k|g)

)∣∣∣∣ ≤ const,

where ν(k|g) =∑

r≥0 grνr(k) is a real function; these relations

clearly do not hold in each order of g. In particular, to prove

Lemma 10.25, from Section 10.6, it is essential that

p(Qt(fi, φi, α)) = p(Qti) ≤ C(1 + |t|n),

where C is a constant not depending on α; this inequality with n = 32

follows from the inequality∫| ˜φt,αi (x)|dx ≤ C(1 + |t|3/2),

which does not hold in the framework of perturbation theory.

However, it is easy to verify that all the proofs in Section 10.6 hold in

the framework of perturbation theory if we allow the use of partial

summation of perturbation series. This allows us to say that the

adiabatic theorem can be proven in the framework of perturbation

theory with partial summation.

Page 357: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 310

310 Mathematical Foundations of Quantum Field Theory

Let us now come back to Hamiltonians that do not generate

vacuum polarization using the results we have proved.

Let us take as our Hilbert space H the space F (L2(E3)) and let

the energy operator H0 and the momentum operator P be defined

by the formulas

H0 =

∫ε(k)a+(k)dk,

P =

∫ka+(k)a(k)dk.

The function ε(k) is presumed to be a smooth function whose

derivatives grow not faster than a power. We also assume that the

function ε(k) satisfies the condition ε(k1 + k2) < ε(k1) + ε(k2) and

does not coincide with a linear function on any open set.) Let us

suppose that the algebra A is the algebra of operators of the form

∞∑m,n

∫fm,n(k1, . . . ,km|p1, . . . ,pn)

× a+(k1) · · · a+(km)a(p1) · · · a(pn)dmkdnp, (11.48)

where fm,n ∈ S and the summation in (11.48) is presumed to be finite

(operators of the form (11.48) can be viewed as operators defined on

the set S∞). In Section 10.5, it was shown that the operators H0,P,

and the algebra A satisfy conditions T1–T7. This allows us to apply

to them all the statements proven in this chapter.

Let us now note that the operator

H0 +∑r≥1

∫Wr(x)dx, (11.49)

where Wr ∈ A,Wrθ = 0, can be written in the form

H0 +∑r≥1

gr∑m,n

∫w(r)m,n(k1, . . . ,km|p1, . . . ,pn)

× δ(k1 + · · ·+ km − p1 − · · · − pn)

× a+(k1) · · · a+(km)a(p1) · · · a(pn)dmkdnp,

Page 358: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:3 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch11 page 311

Translation-Invariant Hamiltonians (Further Investigations) 311

i.e. it is a translation-invariant Hamiltonian not generating vacuum

polarization. (It is easy to check that the Hamiltonian (11.49) belongs

to the class M described in Section 11.1.)

As was shown in this chapter, we can associate Møller matrices

and scattering matrix written as formal series in powers of g to the

Hamiltonian (11.49) (we have already discussed this in Section 11.3,

where we have provided a different proof of this statement).

We can now apply the results of Chapter 10 to gain insight

about the scattering matrix of a translation-invariant Hamiltonian

not generating vacuum polarization.

For example, the following statement follows immediately from

Theorem 10.3, Section 10.4 if we note that for any free Hamiltonian

H0, the identity operator is a dressing operator and satisfies the

conditions of Theorem 10.2 in Section 10.4.

Let H ∈ M, B1, . . . , Bn ∈ M be Hamiltonians not generating

vacuum polarization. Suppose that the operator D = exp(iB1) · · ·exp(iBn) transforms the bare single-particle state a+(k)θ into the

single-particle state of the Hamiltonian H. Then the operator D is a

dressing operator for the Hamiltonian H.

Furthermore, let H = H0 + V ∈ M be a Hamiltonian not

generating vacuum polarization and let the function ε(k) be strictly

convex. Then from the results of Section 10.6, we can conclude that

the definition of the Møller matrix in terms of the adiabatic Møller

matrix given in Section 9.1 is equivalent to the other definitions.

It therefore follows that in the case at hand, Definition 9.2 of the

scattering matrix in Section 9.3 is equivalent to the other definitions

(both of these statements, as noted above, hold in the framework of

perturbation theory with partial summation).

Page 359: Mathematical Foundations of Quantum Field Theory

b2530 International Strategic Relations and China’s National Security: World at the Crossroads

b2530_FM.indd 6 01-Sep-16 11:03:06 AM

This page intentionally left blankThis page intentionally left blankThis page intentionally left blankThis page intentionally left blank

Page 360: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch12 page 313

Chapter 12

Axiomatic Lorentz-InvariantQuantum Field Theory

12.1 Axioms describing Lorentz-invariant scattering

matrices

Let H be a Hilbert space of states. To discuss the Lorentz invariance

of a theory, we should first of all assume that we have a unitary rep-

resentation of an inhomogeneous Lorentz group P (Poincare group)

on the space H. We will denote the unitary operator corresponding

to the element g ∈ P by U(g). The group P contains the group of

space–time translations T as a subgroup; we will define the energy

operator H and the momentum operator P by the relation

exp(−i(Ha0 −Pa)) = U(1, a), (12.1)

where the symbol (1, a) denotes the translation x′ = x+a (in general,

(Λ, a) denotes the element of P defined by the transformation

x′ = Λx+ a).

However, if we only have a representation of the Poincare group,

still the corresponding scattering matrix cannot be defined. We will

assume, as in the axiomatic approach to scattering theory described

in Chapter 10, that we have an asymptotically commutative algebra

A on the space H; we impose conditions guaranteeing the Lorentz

invariance of the scattering matrix.

To begin, we will assume that the presentation U(g) and the

algebra A satisfy the following axioms:

A1. Spectrality : The energy operator H has a single ground state

Φ; the vector Φ is Lorentz-invariant (i.e. U(g)Φ = Φ for any

transformation g ∈ P).

313

Page 361: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch12 page 314

314 Mathematical Foundations of Quantum Field Theory

The vector Φ is called the physical vacuum.

A2. Cyclicity : The vector Φ is a cyclic vector of the algebra A.

A3. Lorentz invariance of the algebra A: If the operator A ∈ A, then

for any transformation g ∈ P , the operator U(g)AU−1(g) also

belongs to the algebra A.

In what follows, axiom A3 can be replaced with the following

weaker condition.

A3′. For any transformation g ∈ P , the algebra A asymptotically

commutes with the algebra U(g)AU−1(g) (i.e. with the alge-

bra consisting of operators of the form U(g)AU−1(g), where

A ∈ A).

Let us analyze the consequences of axiom A1 on the joint

spectrum of the operators H,P.

Lorentz invariance of the vector Φ implies, in particular, that

exp(−iHt)Φ = Φ and exp(iPa)Φ = Φ; hence,

HΦ = PΦ = 0. (12.2)

It is easy to check that the converse holds as well (i.e. from

equation (12.2) and the uniqueness of the ground state, the Lorentz

invariance of Φ follows).

Since the ground state Φ of the energy operator H has zero energy,

by equation (12.2), the operator H is non-negative. This implies that

the joint spectrum of the operators H,P is contained in the half-

space consisting of points (ω,p) where ω ≥ 0. However, the joint

spectrum of the operators H,P must be invariant with respect to

homogenous Lorentz transformations (see Appendix A.9). Therefore,

the spectrum of the operators H,P lies in the cone W , consisting

of points (ω,p), that satisfy the inequality ω2 ≥ p2 (any point

with ω2 < p2 can be transformed to a point (ω′,p′) with ω′ < 0

by a homogenous Lorentz transformation). The physical vacuum Φ

corresponds to the point 0 = (0, 0, 0, 0) ∈W .

Let us consider an irreducible invariant subspace K of the spaceH[i.e. a subspace invariant with respect to the operators U(g) and not

containing non-trivial subspaces that are also invariant with respect

to the operators U(g)]. For example, the subspace A, consisting of

Page 362: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch12 page 315

Axiomatic Lorentz-Invariant Quantum Field Theory 315

vectors of the form λΦ (the vacuum subspace), is irreducible and

invariant. All other irreducible invariant subspaces cannot contain

the vector Φ (otherwise, they would contain the subspace A and not

be irreducible). We will therefore assume that the subspace K does

not contain the vector Φ.

A representation of the Poincare group P by the operators U(g),

viewed as operators on the subspace K, is irreducible. The energy

operator H on the space K cannot vanish, otherwise every vector

x ∈ K is a ground state of the operator H, which contradicts the

uniqueness of the ground state (axiom A1). On the other hand, the

same axiom implies that the operator H on the subspace K (and on

the whole space) is non-negative.

As follows from known facts about irreducible representations of

the group P (see Appendix A.9), the joint spectrum of the operators

H and P on the space K constitutes the set Uµ, consisting of points

(ω,p) ∈ E4 with ω2 − p2 = µ2, ω ≥ 0. Here, µ is an arbitrary non-

negative number, the set Uµ with µ > 0 forms the top-half of a

hyperboloid. Furthermore, one can prove that there exists a unitary

map (an isomorphism) Ψ between the space L2(E3 × N), where N

is a finite set, and the space K, having the property

U(g)Ψ(f) = Ψ(V (g)f), (12.3)

where V (g) is an irreducible presentation of type (µ, n), described in

Appendix A.9. It follows from the relation (12.3) that

HΨ(f) = Ψ(hf),

PΨ(f) = Ψ(pf),(12.4)

where h and p are operators in the space L2(E3×N) defined by the

formulas

(pf)(k, α) = kf(k, α),

(hf)(k, α) =√

k2 + µ2f(k, α).(12.5)

The mapping Ψ can be viewed as n generalized vector functions

Ψα(k) related to Ψ by the formula

Ψ(f) =∑α∈N

∫f(k, α)Ψα(k)dk

Page 363: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch12 page 316

316 Mathematical Foundations of Quantum Field Theory

(here, n is the number of elements in the set N and α runs over

the set N). From the unitarity of the mapping Ψ, it follows that the

generalized vector functions Ψα(k) satisfy the relations⟨Ψα(k),Ψα′(k′)

⟩= δα

′α δ(k− k′)

(orthogonality and normalization by a δ-function); from equa-

tions (12.4) and (12.5), it follows that

PΨα(k) = kΨα(k),

HΨα(k) =√

k2 + µ2Ψα(k).

This implies that the functions Ψα(k) can be viewed as single-particle

states in the sense of the definitions previously used in this book. In

ordinary terminology, the collection of functions Ψ1(k), . . . ,Ψn(k)

(i.e. the mapping Ψ) is identified with a particle having mass µ and

spin s = n−12 . A particle having spin 0 is called scalar.

The smallest subspace, containing all irreducible invariant sub-

spaces differing from A, will be called the single-particle subspace

and will be denoted by B.

It is easy to check that the space B can be represented as a direct

sum of mutually orthogonal irreducible invariant subspaces Ki; using

the isomorphisms Ψi, described above, between the spaces L2(E3 ×Ni) and Ki, we can construct an isomorphism Ψ =

∑i Ψi between

the space L2(E3 ×N) =∑

i L2(E3 ×Ni) and the space B (N is the

union of the sets Ni).

It is clear from these remarks that the single-particle subspace B

defined above coincides with the single-particle subspace in the sense

of the definition in Section 10.1.

The multi-particle subspace M will be defined as the orthogonal

completion of the direct sum A+B (the space A+B can be written

as the sum of all irreducible invariant subspaces).

The space Has will be defined as the Fock space F (B).

Using the isomorphism between the space B and the space

L2(E3 × N), constructed above, we can introduce the operator

generalized functions a+(k, s), a(k, s) on Has, where k ∈ L3, s ∈ N ,

Page 364: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch12 page 317

Axiomatic Lorentz-Invariant Quantum Field Theory 317

satisfying the relations

[a(k, s), a(k′, s′)] = [a+(k, s), a+(k′, s′)] = 0,

[a(k, s), a+(k′, s′)] = δss′δ(k− k′).

The space B is invariant with respect to the operators U(g). The

representation of P by the operators U(g) on the space B clearly

induces a representation on Has; we will denote the operators of

this representation by Uas(g). The asymptotic Hamiltonian Has and

the asymptotic momentum operator Pas are naturally defined by the

relation

exp(−i(Hasa0 −Pasa)) = Uas(1, a),

where (1, a) is a translation.

It is easy to see that the asymptotic Hamiltonian and the

asymptotic momentum operator can be expressed in terms of the

operators a+(k, s), a(k, s) by the formulas

Has =∑i

∑s∈Ni

∫ √k2 + µ2

i a+(k, s)a(k, s)dk,

Pas =∑s∈N

∫ka+(k, s)a(k, s)dk.

We will now formulate an additional axiom which will allow us

to construct a scattering theory in this case (the strong spectral

condition).

A4. The joint spectrum of the operators H,P on the subspace C of

the space H, containing all the irreducible invariant subspaces

of the operators U(g), does not intersect the joint spectrum of

these operators on the orthogonal complement of the space C.

The spectrum of the energy operator H contains a gap (i.e. there

exists an ε > 0, such that the single point of the spectrum of the

operator H on the ray (−∞, ε) is the point 0, corresponding to

the physical vacuum).

It is clear that the space C equals the direct sum of the

subspace A and the single-particle subspace B, therefore, the joint

Page 365: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch12 page 318

318 Mathematical Foundations of Quantum Field Theory

spectrum of the operators H,P on the space C is identical to the

union of the vacuum spectrum (i.e. the spectrum of the operators

H,P on the space A) and the single-particle spectrum (i.e. the

spectrum of the operators H,P on the space B). It follows from the

statements above that the vacuum spectrum consists of the single

point 0 and the single-particle spectrum can be written as the union

of several sets Uµi , specified by the relation ω2 − k2 = µ2i , ω ≥ 0.

The orthogonal complement to the space C = A+B, referred to

as the multi-particle space and denoted by M , the joint spectrum of

H,P on M will be called the multi-particle spectrum.

By using these definitions, we can reformulate axiom A4.

A4′. The vacuum spectrum, single-particle spectrum, and the multi-

particle spectrum are pairwise non-intersecting.

Thus, the spectrum of the operators H,P on the space H is

structured as follows. It contains the point 0, corresponding to the

vector Φ, and several sets Uµi with µi > 0, constituting the single-

particle spectrum, and the multi-particle spectrum, which does not

contain the point 0 and does not intersect any of sets Uµi .

Let us now formulate the following statement. If the axioms 1–4

are satisfied, then the Møller matrices S± and the scattering matrix

S can be defined as in Section 10.1; using the ideas of Section 10.1,

one can prove the existence of the Møller matrices and the scattering

matrix. The Møller matrices and scattering matrix defined in this

manner are Lorentz-invariant (i.e. they satisfy the equations

U(g)S± = S±Uas(g), (12.6)

Uas(g)S = SUas(g) (12.7)

for any transformation g ∈ P).

Indeed, the definitions and discussions in Section 10.1 can be

applied to the case at hand. (The conditions in Section 10.1 are not

necessarily filled, since the definition of the single-particle space in

Section 10.1 is different from the definition in this section; similarly,

Page 366: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch12 page 319

Axiomatic Lorentz-Invariant Quantum Field Theory 319

the definitions of the space Has differ. However, in light of the fact

that the single-particle space B defined in this section is contained in

the single-particle space in the sense of Section 10.1, the discussions

in Chapter 10 do not need to be modified in this case.) The Lorentz

invariance of the Møller matrices and the scattering matrix can be

proved by the discussions in Section 6, Chapter 6 in Jost (1965); we

will sketch another proof in Chapter 13.

Let us note that the results of this section hold if the asymp-

totically commutative family of operators is replaced by an asymp-

totically commutative family of operator generalized functions (in

the sense of Section 10.5). The domain of definition of these

operator generalized functions should be invariant with respect to

the operators U(g), where g ∈ P. Conditions A2 and A3 require

corresponding modifications (we ought to require that the vector Φ

is cyclic vector for this family and that for every generalized operator

function A(x, t), the corresponding generalized operator functions

Ag(x, t) = UgA(x, t)U−1g , where g is an arbitrary element of the

group P, should also belong to the family).

12.2 Axiomatics of local quantum field theory

To formulate the axioms of Lorentz-invariant theory, we added

extra conditions to the axiomatic scattering theory constructed in

Chapter 10. These extra conditions guarantee the Lorentz invariance

of the scattering matrix.

Here, we will provide a different set of axiomatics introduced by

Haag and Araki.

Let us fix a representation Ug of the Poincare group P on the

Hilbert space H, and, as in the previous section, let us require the

spectral condition (axiom A1).

Let us assume that for every bounded set O ⊂ E4, there is a

collection R(O) of bounded operators on the space H. We assume

that this collection is closed under addition and multiplication of

operators, multiplication by a scalar, involution A → A∗, and weak

limits (such a collection of operators is called a W ∗-algebra).

Page 367: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch12 page 320

320 Mathematical Foundations of Quantum Field Theory

Let us assume that the algebra R(O) has the following properties:

R1. If O1 ⊂ O2, then R(O1) ⊂ R(O2).

R2. Covariance with respect to the Poincare group:

UgR(O)U−1g = R(gO).

R3. Local commutativity: If any point of the set O1 is separated by

a space-like interval from any point of the set O2, then R(O1)

and R(O2) commute with each other.

R4. Cyclicity: The physical vacuum Φ is cyclic with respect to the

union R of the algebras R(O).

It is easy to see that the union R of the algebras R(O) is an

asymptotically commutative algebra. Indeed, let A ∈ R(O1) ⊂ R,

B ∈ R(O2) ⊂ R. Then, by condition R2, we have

A(x, t) = exp(iHt− iPx)A exp(−iHt+ iPx) ∈ R(O1 − x),

where x = (x, t). Since the sets O1 and O2 are bounded, there exists

an a such that for |x|2 − |t|2 ≥ a2, any point of the set O1 − x is

separated from any point of the set O2 by a space-like interval. Using

condition R3, we obtain that for |x|2 − |t|2 ≥ a2,

[A(x, t), B] = 0.

On the other hand, ‖A(x, t)‖ = ‖A‖, and therefore,

‖[A(x, t), B]‖ ≤ 2‖A‖ · ‖B‖.

Combining these relations, we see that

‖[A(x, t), B]‖ ≤ C 1 + |t|n

1 + |x|n

for any n (to show this, it is enough to note that

1 + |t|n

1 + |x|n> c > 0 for |x|2 − |t|2 ≥ a2

).

Furthermore, the algebra R is clearly invariant with respect to

Lorentz transformations.

Page 368: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch12 page 321

Axiomatic Lorentz-Invariant Quantum Field Theory 321

Therefore, the algebra R and the representation Ug satisfy all

conditions of Section 12.1, except for axiom A4.

If we require that axiom A4 holds, then the considerations

provided in Section 12.1 allow us to prove that there exists a Lorentz-

invariant scattering matrix.

Another important axiomatics is due to Wightman. Again, we

require that the Poincare group is represented in Hilbert space Hand the spectral condition A1 is satisfied.

We assume that an operator generalized function φ(x) is defined

on the spaceH (here, x ∈ E4). More precisely, we assume that for any

function f ∈ S(E4), there is an operator φ(f); all the operators φ(f)

are defined on the same linear set D and transform D into itself;

the operator φ(f) must depend linearly on f and for any vectors

ξ, η ∈ D, the functional 〈φ(f)ξ, η〉 must continuously depend on f in

the topology of the space S (i.e. it must define a generalized function

from the space S ′(E4)).

As before, we will write φ(f) =∫φ(x)f(x)dx; the operator

generalized function φ(x) will be referred to as a quantum field.

Let us assume that the quantum field φ(x) is Hermitian (i.e. the

operators φ(f) and φ(f) are Hermitian conjugate).

We say that the representation U(g) of the Poincare group P and

the operator generalized function φ(x) (the quantum field) satisfy

the Wightman axioms,1 if the following conditions are satisfied:

W1. Transformation law of quantum fields: The operators Ug keep

the set D fixed: UgD = D. The quantum field φ(x) transforms

according to the law

U(g)φ(x)U−1(g) = φ(gx). (12.8)

More precisely, equation (12.8) implies that

U(g)φ(f)U−1(g) = φ(fg),

where

fg(x) = f(g−1x).

1Here, we have formulated the Wightman axioms in the case of a scalar field;for the general case, see, for example, Bogolyubov et al. (1969).

Page 369: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch12 page 322

322 Mathematical Foundations of Quantum Field Theory

W2. Locality (local commutativity): If the points x, y ∈ E4 are sepa-

rated by a space-like interval, then φ(x) commutes with φ(y).

More precisely, we require that the operators φ(f) and φ(g)

commute in the case when the support of the functions f and g

are separated by a space-like interval.

W3. Cyclicity : The vector Φ (the ground state of the energy

operator) is a cyclic vector of the family of operators φ(f).

The Wightman functions wm(x1, . . . , xm) will be defined by the

relation

wm(x1, . . . , xm) = 〈φ(x1) · · ·φ(xm)Φ,Φ〉 .

By the kernel theorem (see Appendix A.7), the function wn can be

viewed as a generalized function from S′(E4m).

Using the considerations of Section 8.2, one can prove the

reconstruction theorem, which states that knowing the Wightman

functions wm, we can construct a Hilbert space H, a Lorentz group

representation, and a quantum field φ(x). It is easy to work out

the conditions on the functions wm(x1, . . . , xm) which guarantee

that these functions are Wightman functions for a quantum field,

satisfying the Wightman axioms (these conditions are analogous to

conditions 1–6 from Section 8.2).2

Let us now show that by adding axiom A4 from Section 12.1 to

the Wightman axioms, we can provide a definition of Møller matrices

(hence of the scattering matrix) and prove their existence and Lorentz

invariance. To do this, it is sufficient to consider a family of operator

generalized functions φ(gx), where g runs over a Poincare group P,

and note that this family is asymptotically commutative in the sense

of Section 10.5 and it satisfies conditions A1–A4 from Section 12.1.

(Axiom A3 follows immediately from axioms W1, axiom A2 from

axiom W3, condition D5 from Section 10.5 for δ < 1 follows from

local commutativity.)

2For more details, see Streater and Wightman (2016), Jost (1965), andBogolyubov et al. (1969).

Page 370: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch12 page 323

Axiomatic Lorentz-Invariant Quantum Field Theory 323

Thus, we can see that all the statements which can be obtained

from the axioms in Section 12.1 can be derived from Wightman’s

axioms. However, there are theorems which can be proved in

Wightman’s axiomatics that have not been proven (and, apparently,

cannot be proven) in the axiomatic formulation in Section 12.1.

(This is not surprising; Wightman axiomatized local quantum field

theory; the axioms of Section 12.1 are based on weaker assumption

of asymptotic commutativity).

The most important statement of this type are dispersion rela-

tions obtained first in Bogolyubov et al. (1956) and derived from

Wightman axioms in Hepp and Epstein (1971).

Let us also mention the CPT theorem, stating that the Wightman

axioms imply CPT invariance (i.e. the existence of additional

symmetries).

Let us formulate Borchers’ theorem:

If two quantum fields φ1(x), φ2(y) on a Hilbert space satisfy the

Wightman axioms and are mutually local (i.e. [φ1(x), φ2(y)] = 0 for

any points x, y separated by a space-like interval), then they define

the same scattering matrix.

The CPT theorem can be used in the proof of this theorem.

However, this theorem has another proof: it follows from the results of

Section 10.1 (from the theorem that two asymptotically commutative

algebras define the same scattering matrix).

Finally, let us analyze the connection between the Wightman

axioms and the Haag–Araki axioms. Let us assume that the operators

φ(f) =∫f(x)φ(x)dx, entering in the Wightman axioms, are not

only Hermitian but also essentially self-adjoint for any real smooth

function f with compact support. For every such function f , one can

construct a family Tf of bounded operators, arising as functions of

the operator φ(f) (i.e. having the form α(φ(f)), where α is a bounded

function). Let us use the symbol R(O) for the smallest W ∗-algebra,

containing all families Tf , where f is a smooth real function with

support in the set O. It is easy to check that the algebras R(O)

satisfy the Haag–Araki axioms.

Page 371: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch12 page 324

324 Mathematical Foundations of Quantum Field Theory

12.3 The problem of constructing a non-trivial

example

It is natural to try to construct a Lorentz-invariant quantum theory

by means of the quantization of Lorentz-invariant classical theory.

Let us recall that in Section 8.3, we considered the question of

quantizing a classical theory corresponding to the Hamiltonian

H(π, φ) =1

2

∫π2(x)dx + V (φ).

It is easy to see that such a classical system is Lorentz invariant if

H(π, φ) =1

2

∫π2(x)dx−1

2

∫φ(x)∆φ(x)dx+

∫U(φ(x))dx. (12.9)

Indeed, the Hamiltonian equations for the system, described by the

functional Hamiltonian (12.9), have the form

∂φ(x, t)

∂t= π(x, t),

∂π(x, t)

∂t= ∆φ(x, t)− F (φ(x, t)),

where

F (φ) =∂U(φ)

∂φ.

Thus, the function φ(x) = φ(x, t) satisfies the equation

φ(x) + F (φ(x)) = 0,

which is clearly Lorentz invariant (the equation does not change form

when replacing φ(x) by φ(Λx), where Λ is a Lorentz transformation).

We can also prove the Lorentz invariance of classical systems,

described by the Hamiltonian (12.9), by considering the action

functional of such a system. This functional can be presented in the

form

S =

∫ ((∂φ

∂t

)2

− (∇φ)2

)dx−

∫U(φ(x))dx,

which is invariant under Lorentz transformations.

Page 372: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch12 page 325

Axiomatic Lorentz-Invariant Quantum Field Theory 325

Let us now consider the simplest system with the Hamiltonian of

the form (12.9), a system with

U(φ(x)) =1

2m2φ2(x). (12.10)

After quantizing this system, we obtain a quantum Hamiltonian

H =1

2

∫π2(x)dx− 1

2

∫φ(x)∆φ(x)dx +

1

2m2

∫φ2(x)dx, (12.11)

where π(x), φ(x) are symbols satisfying the relations

π+(x) = π(x), φ+(x) = φ(x),

[π(x), π(y)] = [φ(x), φ(y)] = 0,

[π(x), φ(y)] =1

iδ(x− y).

The Hamiltonian (12.11) can be written in the form (8.26), where

ν(x) = −∆δ(x)+m2δ(x). An operator realization of Hamiltonians of

the form (8.26) (free Hamiltonians) is described in Section 8.3. Using

the results of Section 8.3, we can prove that the operator realization

of the Hamiltonian (12.11) can be constructed on the Fock spaceH =

F (L2(E3)); as the energy operators and the momentum operators,

we can select the operators

H =

∫ω(k)a+(k)a(k)dk,

P =

∫ka+(k)a(k)dk,

where ω(k) =√

k2 +m2; the operators (the operator generalized

functions) φ(x, t) are defined by the formula

φ(x, t) = (2π)−3/2

∫(a+(k) exp(iω(k)t− ikx)

+ a(k) exp(−iω(k)t+ ikx))dk√2ω(k)

. (12.12)

The ground state Φ coincides with the Fock vacuum θ.

Page 373: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch12 page 326

326 Mathematical Foundations of Quantum Field Theory

The operators φ(x) = φ(x, t) satisfy the Lorentz-invariant

equation

φ(x) +m2φ(x) = 0.

It is therefore natural to expect that the quantum system is Lorentz

invariant. Actually, the representation of the translation group on

the space F (L2(E3)), given rise by the operators H and P, can be

extended to a unitary representation of Poincare group in such a way

that

U(g)φ(x)U−1(g) = φ(gx)

(here, g is a Lorentz transformation and U(g) is the corresponding

unitary operator). The necessary Lorentz group representation can

be obtained by considering on the space L2(E3) the representation

of the type (m, 1), described in Appendix A.9. (We should note that

the representation on the space L2(E3) specifies a representation on

the space F (L2(E3)); we have already used this fact in Section 12.1).

It is easy to check that the operator generalized function φ(x, t)

defined by the formula (12.12) and the representation of the Poincare

group satisfy all the Wightman axioms.

The given quantum system is called a free scalar field with

mass m. Corresponding particles are non-interacting bosons with

spin 0. The scattering matrix is trivial as it should be for non-

interacting particles.

To obtain a non-trivial Lorentz-invariant scattering matrix, we

can try to quantize a classical theory with the Hamiltonian of the

form (12.9), where the function U(φ) is not quadratic. Unfortunately,

this approach leads to problems related to the fact that in every step

one encounters singular expressions which cannot be easily given a

precise meaning. For example, the expressions of the form φn(x),

where n ≥ 2, and similarly more general expressions ν(φ(x)), where

ν(φ) is a nonlinear function, are not well defined; therefore, the

equation

φ(x) + F (φ(x)) = 0,

for the field operators φ(x) does not have precise meaning.

Page 374: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch12 page 327

Axiomatic Lorentz-Invariant Quantum Field Theory 327

Let us suppose that the Hamiltonian (12.9) can be presented

in the form H = H0 + V , where H0 is a Hamiltonian of the

form (12.10), V =∫ν(φ(x))dx. After quantization, the resulting

formal Hamiltonian can be written in the form

H = H0 + V =1

2

∫π2(x)dx− 1

2

∫φ(x)∆φ(x)dx

+1

2m2

∫φ2(x)dx +

∫ν(φ(x))dx. (12.13)

If we try to express Hamiltonian in terms of the symbols a+(k), a(k),

as in Section 8.3, and bring the resulting expression to normal form

by means of CCR, then except for the standard infinite constants,

which we have previously thrown away, we will obtain other infinite

summands. Therefore, one often considers the formal Hamiltonian

H = H0 +N

∫ν(φ(x))dx, (12.14)

where N is the normal product symbol (i.e. we express∫ν(φ(x))dx

in terms of a+(x), a(k) by the formulas (8.27), and then reorder the

symbols in normal order, assuming they commute). Replacing the

formal Hamiltonian (12.13) by the formal Hamiltonian (12.14), we

can get rid of some infinities, but other infinities remain. In particu-

lar, calculating the Green functions of the Hamiltonian (12.14) in the

framework of perturbation theory, we encounter diverging integrals.3

Unfortunately, even now, a consistent method of overcoming these

challenges in the case of arbitrary function ν(φ) or even in the case

when the function ν(φ) is a polynomial does not exist. However, if

the function ν(φ) is a polynomial of order ≤ 4 (i.e. ν(φ) = aφ3+bφ4),

one can overcome these problems in the framework of perturbation

theory. There exists a method that allows us to construct a family of

Lorentz-invariant scattering matrices that correspond to the family

of Hamiltonians of the form H0 + N∫ν(φ(x))dx. This method was

introduced by Feynman; it is called covariant renormalization.

3We should note that in the one-dimensional case (i.e. in the case when in theHamiltonian (12.14) integration is performed over a single variable) calculations ofGreen functions of the Hamiltonian (12.14) do not give rise to diverging integrals.

Page 375: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch12 page 328

328 Mathematical Foundations of Quantum Field Theory

Let us consider for the sake of concreteness the case when ν(φ) =

bφ4. We consider the family of Hamiltonians of the form

H = H0(w) + bVΛ

=1

2

∫π2(x)dx +

1

2

∫w(x− y)φ(x)φ(y)dxdy

+ b

∫fΛ(x1 − x4,x2 − x4,x3 − x4)

+ φ(x1)φ(x2)φ(x3)φ(x4)d4x.

Here,w(x) is an arbitrary function, fΛ(x1,x2,x3) = Λ3f(Λx1,Λx2,Λx3),

where the function f ∈ S and satisfies the condition∫f(x1,x2,x3)

d3x = 1. It is easy to see that limΛ→∞ fΛ(x1,x2,x3) = δ(x1)

δ(x2)δ(x3) and therefore in the limit as Λ → ∞, the interaction

VΛ becomes V =∫φ4(x)dx; the replacement of V by VΛ is called

ultraviolet cutoff. (Defining VΛ, we can use the normal product

in place of the regular product; the class of Hamiltonians under

consideration does not change in this case.) Let us select the

function w(x) from the condition that energy of the particles, defined

by the Hamiltonian H0(w) + bVΛ, equals m for k = 0; we will

denote the resulting Hamiltonian H(m, b,Λ) (in the framework of

perturbation theory, the function w(x) is easy to find). The transition

to the Hamiltonian H(m, b,Λ) is called mass renormalization (the

number m corresponds to the mass of the particles defined by the

Hamiltonian). Let us denote the scattering matrix corresponding

to the Hamiltonian H(m, b,Λ) by S(m, b,Λ). In order to obtain a

Lorentz-invariant theory, we must take the limit Λ → ∞ (remove

the ultraviolet cutoff). However, taking the limit Λ→∞, we obtain

divergent expressions in the perturbation series.

In order to obtain a non-trivial Lorentz-invariant scattering

matrix in the framework of perturbation theory, we should assume

that the quantity b (bare charge) changes in the process of

removing the ultraviolet cutoff. In the perturbation theory, one can

prove the following statement: there exists a function b(m,Λ, g) such

that the scattering matrix S(m, b(m,Λ, g),Λ) has a finite, non-trivial,

Lorentz-invariant limit S(m, g) as Λ → ∞. (In order to find the

Page 376: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch12 page 329

Axiomatic Lorentz-Invariant Quantum Field Theory 329

function b(m,Λ, g), we suppose that in the process of removing the

cutoff we leave a physical quantity g fixed and express the bare charge

in terms of the cutoff parameter and this quantity. We can take as

g any matrix element of the scattering matrix; for some particular

choice, it is called physical charge or dressed charge.) The result

(which we have described in a non-standard form) is the theory of

renormalization. The Green functions Gn(x1, . . . , xn|m,Λ, g) of the

Hamiltonian H(m, b(m,Λ, g),Λ) do not have a finite limit as Λ→∞,

however, we can find a function Z(m,Λ, g) such that the functions

Z−n/2(m,Λ, g)Gn(x1, . . . , xn|m,Λ, g) have a finite Lorentz-invariant

limit as Λ→∞.

The statements above suggest the following way of constructing

objects satisfying the Wightman axioms and giving rise to non-trivial

scattering matrices. First, we must find the functions b(m,Λ, g) and

Z(m,Λ, g), such that a non-trivial finite limit

limΛ→∞

Z−n/2(m,Λ, g)wn(x1, . . . , xn|m,Λ, g) = wn(x1, . . . , xn|m, g)

exists. Here, wn(x1, . . . , xn|m, g) is a Wightman function of the

Hamiltonian H(m, b(m,Λ, g),Λ). One can hope that the functions

w(x1, . . . , xn|m, g) defined this way satisfy the conditions necessary

for constructing objects obeying the Wightman axioms and that

the corresponding scattering matrix is non-trivial. Unfortunately, at

present, this hypothesis is unproven. Moreover, in four-dimensional

space–time, an example of objects satisfying the Wightman functions

and giving rise to a non-trivial scattering matrix is not known.

Clearly, if we can construct a non-trivial scattering matrix in

Wightman’s axioms, then we will obtain a non-trivial example of

a scattering matrix in the axioms of Section 12.1. However, the other

direction does not hold — a non-trivial example in the axioms of

Section 12.1 can be easier to construct.

The situation in two-dimensional and three-dimensional space–

time is much better. A new branch of mathematical physics —

constructive field theory — was born from the attempts to find non-

trivial examples and in two and three dimensions, it was successful.

Two-dimensional conformal field theories were thoroughly analyzed

Page 377: Mathematical Foundations of Quantum Field Theory

March 27, 2020 16:33 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch12 page 330

330 Mathematical Foundations of Quantum Field Theory

in numerous papers starting with the seminal paper by Belavin et al.

(1984). Another way to construct two-dimensional theories comes

from integrable models.

In all dimensions, the considerations based on supersymmetry and

superstring theory led to much deeper understanding of quantum

field theories.

We are moving forward!

Page 378: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 331

Chapter 13

Methods of Quantum Field Theoryin Statistical Physics

13.1 Quantum statistical mechanics

The equilibrium state in (classical or quantum) statistical mechanics

can be defined as a state maximizing the entropy under some given

conditions.

In quantum mechanics, we represent a state by the density

matrix K; the entropy of this state is defined by the formula

S = −TrK logK. Denoting the Hamiltonian by H, we can say that

the equilibrium state maximizes the entropy for a given mean energy

E = Tr HK. (More precisely, we can say that this is an equilibrium

state of the canonical ensemble or a Gibbs state.) Let us check that

this state has the form

e−βH

Z, (13.1)

where β can be interpreted as the inverse temperature β = 1T and

Z = Tr e−βH is called the partition function or the statistical sum.

If the spectrum of H is discrete, then Z =∑e−βEi where Ei are the

energy levels. In the case when the ground state is non-degenerate,

we see that the equilibrium state tends to the ground state and its

entropy tends to zero when T → 0.

Note that (13.1) makes sense only if Z is finite (the operator e−βH

belongs to the trace class).

331

Page 379: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 332

332 Mathematical Foundations of Quantum Field Theory

To prove (13.1), we use the method of Lagrange multipliers, reduc-

ing our problem to finding the stationary points of the functional

L = −TrK logK − βTr(HK − E)− ζ(TrK − 1).

Calculating the variation of L, we obtain

δL = −Tr logKδK − Tr δK − β Tr HδK − ζ Tr δK,

hence at the stationary point,

− logK − 1− βH − ζ = 0.

This is equivalent to (13.1) with Z = e1+ζ .

Note that in the calculation of variation, we have used the formula

δTrφ(K) = Trφ′(K)δK, (13.2)

where φ′ stands for the derivative of the function φ. It is easy to

check this formula in the case when φ(K) = Kn; to treat the general

case, we can represent φ as a limit of polynomials.

The partition function Z does not have any physical meaning, but

there are many physical quantities that can be conveniently expressed

in terms of Z. For example, differentiating the definition of Z with

respect to β, we obtain

E = H = −∂ logZ

∂β.

The expression for the entropy in equilibrium state is

S = βE + logZ.

Introducing the notion of free energy by the formula F = E − TS,

we obtain

F = −T logZ.

If the Hamiltonian depends on a parameter λ and H(λ) = H +

λA+ · · · then using (13.2), we can write

Z(λ) = Z + (−β)λTrAe−βH + · · · .

Page 380: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 333

Methods of Quantum Field Theory in Statistical Physics 333

Here, · · · stands for higher order terms with respect to λ. Note that

A =TrAe−βH

Z

is the mean value (expectation value) of the observable A in the

equilibrium state. We will also use the notation 〈A〉 for the expecta-

tion value. The expectation values 〈A1 · · ·An〉 or, more generally,

〈A1(t1) · · ·An(tn)〉 are called correlation functions (here, Ai are

observables and Ai(t) are the corresponding Heisenberg operators).

If we want to emphasize that the correlation functions are calculated

for the equilibrium state with the temperature T = 1β , we will use

the notation 〈A1(t1) · · ·An(tn)〉β.If A and B are two observables, then

〈A(t)B〉β = 〈BA(t+ iβ)〉β. (13.3)

This equation, called the Kubo–Martin–Schwinger (KMS) condition,

can be proved by simple formal manipulations if the Hilbert space

H is finite-dimensional. In the infinite-dimensional case, one should

check that the correlation function 〈BA(t)〉β can be extended

analytically to the strip 0 ≤ =t ≤ β (here, = stands for imaginary

part).

It is easy to see that

A = −T ∂ logZ

∂λ=∂F

∂λ(13.4)

(the derivatives are calculated at the point λ = 0).

If the Hamiltonian depends linearly on a set of parameters

λ1, . . . , λk, we can calculate the correlation functions by differenti-

ating the free energy F . For example, if H = H0 +λ1A1 + · · ·+λkAk,

we have

∂2F

∂λi∂λj= 〈AiAj〉 − 〈Ai〉〈Aj〉. (13.5)

(The derivatives are calculated at the point λi = 0.) The RHS

of (13.5) is called a truncated correlation function. Higher truncated

correlation functions can be defined as higher derivatives of F. (See

Section 10.1 for the definition of truncated correlation functions in

terms of correlation functions.)

Page 381: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 334

334 Mathematical Foundations of Quantum Field Theory

If we have several integrals of motion, we can consider the

equilibrium state fixing the expectation values of all these integrals

and looking for the maximal value of entropy under these conditions.

For example, we can fix the expectation values of energy and of the

number of particles N . Then we can repeat the above consideration

by adding one more Lagrange multiplier. We obtain the following

formula for the equilibrium state:

e−β(H−µN)

Z. (13.6)

Note that we have denoted the new Lagrange multiplier as −βµto agree with standard notation (µ can be identified with chemical

potential).

13.1.1 Examples

Let us consider the multidimensional harmonic oscillator as an

example. Introducing the creation and annihilation operators

obeying CCR, we can represent the Hamiltonian in the form

H =∑

i ωia+i ai + C, where C =

∑ωi/2 is the energy of the ground

state. We shift the energy scale assuming that the ground state has

zero energy: C = 0. Then the energy levels are∑niωi where ni is a

non-negative integer. It is easy to calculate the partition function

Z = Π1

1− e−βωi,

the free energy

F =1

β

∑log(1− e−βωi),

and the mean value of energy

H =∑

ωini,

where ni = 1eβωi−1

.

The same formulas work for non-interacting identical bosons. In

the case of non-interacting identical fermions, we have very similar

formulas. The creation and annihilation operators, in this case, obey

CAR and the energy levels are given by the formula∑ωini where

Page 382: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 335

Methods of Quantum Field Theory in Statistical Physics 335

ni = 0, 1. The formulas for the partition function, free energy, and

mean energy are given by

Z = Π(1 + e−βωi),

F = − 1

β

∑log(1 + e−βωi),

H =∑

ωini,

where ni = 1eβωi+1

.

13.2 Equilibrium states of translation-invariant

Hamiltonians

In momentum representation, one can write a translation-invariant

Hamiltonian in the form

H =∑m,n

∫Hm,n(k1, . . . , km | l1, . . . , ln)

× a+(k1) · · · a+(km)a(l1) · · · a(ln)dmkdnl, (13.7)

where Hm,n(k1, . . . , km | l1, . . . , ln) contains a factor δ(k1 + · · · + km− l1 − · · · − ln). (We assume that the arguments belong to RD.)

Even in the simplest case, when H =∫ω(p)a+(p)a(p)dp, the

formula for the equilibrium state makes no sense (the integral for

the partition function diverges). Moreover, as we have seen, the

formula (13.7) in general does not define an operator on Fock

space. However, the Hamiltonian with volume cutoff HΩ specifies an

operator in Fock space under some mild conditions and we can define

an equilibrium state and correlation functions. In particular, we

can consider the correlation functions βwΩn (k1, ε1, t1, . . . , kn, εn, tn) =

〈aε1k1(t1) · · · aεnkn(tn)〉β generalizing Wightman functions. (Wightman

functions are correlation functions for T = 0.)

We can define correlation functions in infinite volume βwn =

limβ wΩn by taking the limit Ω → ∞ (removing the volume cutoff).

As in Section 8.2, the limit is understood in the sense of generalized

functions. Under some mild conditions, one can prove that the

correlation functions in infinite volume obey KMS condition (13.3)

Page 383: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 336

336 Mathematical Foundations of Quantum Field Theory

(one should verify that these functions can be analytically extended

to the strip 0 ≤ =t ≤ β).

Note that the limit at hand does not always exist, or, better, there

exist many limits, i.e. many equilibrium states in infinite volume. The

limiting procedure we have used can be slightly modified in many

ways: we can use various volume cutoffs or add to the Hamiltonian

HΩ terms that tend to zero as Ω tends to ∞. If the limit exists,

we say that it gives correlation functions in the equilibrium state

(more precisely, in one of equilibrium states.) We can repeat with

minimal changes the considerations of Sections 8.1 and 8.2. Again,

as in Section 8.2, we can construct for every β, a Hilbert space H =

H(β), a dense subset D, a vector Φ, operators H, P, and operator

generalized functions aε(k, t) in such a way that

βwn(k1, ε1, t1, . . . , kn, εn, tn) = 〈aε1(k1, t1) · · · aεn(kn, tn)Φ,Φ〉

(this is the Reconstruction theorem). The only difference is that the

spectral condition should be replaced by the KMS condition.

The operator generalized functions aε(k, t) obey the Heisenberg

equations. This allows us to say that the objects we have constructed

specify an operator realization of the equilibrium state. (The defini-

tion of an operator realization is the same as in Section 8.1, but

the condition that Φ is a ground state is replaced by the KMS

condition.)

13.3 Algebraic approach to quantum theory

Quantum mechanics can be formulated in terms of an algebra

of observables. The starting point of this formulation is a unital

associative algebra A over C (the algebra of observables). One

assumes that this algebra is equipped with antilinear involution

A→ A∗. One says that a linear functional ω on A specifies a state if

ω(A∗A) ≥ 0 (i.e. if the functional is positive). The space of states will

be denoted by C. If ω(1) = 1, we say that the state is normalized.

The probability distribution ρ(λ) of a real observable A = A∗ in

a normalized state ω should obey the relation ω(An) =∫λnρ(λ)dλ.

(In general, this formula does not specify the probability distribution

Page 384: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 337

Methods of Quantum Field Theory in Statistical Physics 337

uniquely. However, if for every continuous function f of a real

argument, we can define an element f(A) of the algebra A, then

the probability distribution is specified uniquely by the formula

ω(f(A)) =∫f(λ)ρ(λ)dλ. In particular, f(A) is well defined for every

continuous function f if A is a C∗-algebra.)

In the standard exposition of quantum mechanics, the algebra

of observables consists of operators acting on a (pre-)Hilbert space.

Every vector x having a unit norm specifies a state by the formula

ω(A) = 〈Ax, x〉. (More generally, a density matrix K defines a state

ω(A) = TrAK.) This situation is in some sense universal: for every

state ω on A, one can construct a (pre-)Hilbert space H and a

representation of A by operators on this space in such a way that

the state ω corresponds to a vector in this space.

To construct H, one defines inner product on A by the formula

〈A,B〉 = ω(A∗B). The space H can be obtained from A by means of

factorization with respect to zero vectors of this inner product. The

inner product on A descends toH providing it with the structure of a

pre-Hilbert space. The state ω is represented by a vector Φ of H that

corresponds to the unit element of A. An operator of multiplication

from the left by an element C ∈ A descends to an operator C acting

on H; this construction gives a representation of A in the algebra

of bounded operators on H. The algebra of operators of the form C

where C ∈ A will be denoted by A(ω). (In our definition, H is a

pre-Hilbert space; taking its completion, we obtain a representation

of A by operators in Hilbert space H.) The above construction is

called Gelfand–Naimark–Segal (GNS) construction.

Although every state of the algebra A can be represented by a

vector in Hilbert space, in general, it is impossible to identify Hilbert

spaces corresponding to different states.

Time evolution in the algebraic formulation is specified by a one-

parameter group α(t) of automorphisms of the algebra A preserving

the involution. This group acts in the obvious way on the space of

states. If ω is a stationary state (a state invariant with respect to time

evolution), then the group α(t) descends to a group U(t) of unitary

transformations of the corresponding space H. The generator H of

U(t) plays the role of the Hamiltonian; it can be considered as a self-

adjoint operator in H. The vector Φ representing ω obeys HΦ = 0.

Page 385: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 338

338 Mathematical Foundations of Quantum Field Theory

We say that the stationary state ω is a ground state if the spectrum

of H is non-negative.

Every element B of the algebra A specifies two operators acting

on linear functionals on A:

(Bω)(A) = ω(BA), (13.8)

(Bω)(A) = ω(AB∗). (13.9)

Note that the first of these operators is denoted by the same

symbol as the element. The operator B commutes with the operator

B′. These operators do not preserve positivity, but the operator

BB does. This means that the operator sending ω into BB(ω)

acts on the space of states. If a state ω corresponds to a vector

Φ in a representation of A (i.e. ω(A) = 〈AΦ,Φ〉), then the state

corresponding to the vector BΦ is equal to BBω.

13.3.1 Quantum field theory and statistical

physics in Rd

Symmetries of quantum theory in the algebraic formulation are

automorphisms of the algebra A commuting with the involution and

the evolution automorphisms α(t). We are especially interested in the

case when among the symmetries are operators α(x, t), where x ∈ Rdand t ∈ R, that are automorphisms of A obeying α(x, t)α(x′, t′) =

α(x + x′, t+ t′). We will use the notation A(x, t) for α(x, t)A, where

A ∈ A.

These automorphisms can be interpreted as space–time transla-

tions. One should expect that starting with a formal translation-

invariant Hamiltonian, one can construct an algebra A and space–

time translations α(x, t) in such a way that the corresponding

equilibrium states can be identified with the equilibrium states of the

preceding section (this statement can be justified in the framework

of perturbation theory).

We say that the algebra A and space–time translations specify

a quantum theory in Rd. We will define particles as elementary

excitations of the ground state; the theory of these particles and their

collisions is a quantum field theory through an algebraic approach.

Page 386: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 339

Methods of Quantum Field Theory in Statistical Physics 339

The action of a translation group on A induces an action of this

group on the space of states. In relativistic quantum field theory,

one should have an action of the Poincare group on the algebra of

observables (hence on states).

Let us now consider a state ω that is invariant with respect to

the translation group. We will define (quasi-)particles as “elementary

excitations” of ω. To consider collisions of (quasi-)particles, we should

require that ω satisfies a cluster property in some sense.

The weakest form of cluster property is the following condition:

ω(A(x, t)B) = ω(A)ω(B) + ρ(x, t), (13.10)

where A,B ∈ A and ρ is small in some sense for x→∞. For example,

we can impose the condition that∫|ρ(x, t)|dx < c(t), where c(t) has

at most polynomial growth. Note that (13.10) implies asymptotic

commutativity in some sense: ω([A(x, t), B]) is small for x→∞.To formulate a more general cluster property, we introduce the

notion of correlation functions in the state ω :

wn(x1, t1, . . . ,xn, tn) = ω(A1(x1, t1) . . . An(xn, tn)),

where Ai ∈ A. These functions generalize Wightman functions of

relativistic quantum field theory. We consider the corresponding

truncated correlation functions wTn (x1, t1, . . . ,xn, tn).

We have assumed that the state ω is translation-invariant; it

follows that both correlation functions and truncated correlation

functions depend on the differences xi − xj , ti − tj . We say that

the state ω has the cluster property if the truncated correlation

functions are small for xi − xj → ∞. A strong version of the

cluster property is the assumption that the truncated correlation

functions tend to zero faster than any power of min ‖xi − xj‖. Then

its Fourier transform with respect to variables xi has the form

νn(p2, . . . ,pn, t1, . . . , tn)δ(p1 + · · · + pn), where the function νn is

smooth. We will later need a weaker form of cluster property that

can be formulated as a requirement that the function νn is three

times continuously differentiable with respect to p2, . . . ,pn.

Instead of cluster property, one can impose a condition of

asymptotic commutativity of the algebra A(ω). In other words, one

Page 387: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 340

340 Mathematical Foundations of Quantum Field Theory

should require that the commutator [A(x, t), B], where A,B ∈ A is

small for x → ∞. For example, we can assume that for every n the

norm of this commutator is bounded from above by Cn(t)(1+‖x‖)−n,

where the function Cn(t) has at most polynomial growth. (This

property will be called strong asymptotic commutativity.) Note

that in relativistic quantum field theory, in the Haag–Araki or the

Wightman formulation, strong asymptotic commutativity can be

derived from locality.

Let us show how one can define one-particle excitations of the

state ω and the scattering of (quasi-)particles.

The action of the translation group on A generates a unitary

representation of this group on the pre-Hilbert space H constructed

from ω. Generators of this representation P and −H are identified

with the momentum operator and the Hamiltonian. The vector in

the space H that corresponds to ω will be denoted by Φ. If Φ is

a ground state, we say that it is the physical vacuum. If ω obeys

the KMS condition ω(A(t)B) = ω(BA(t+ iβ)), we say that ω is an

equilibrium state with the temperature T = 1β .

13.3.2 Particles and quasiparticles

We say that a state σ is an excitation of ω if it coincides with ω at

infinity. More precisely, we should require that σ(A(x, t))→ ω(A) as

x → ∞ for every A ∈ A. Note that the state corresponding to any

vector AΦ, where A ∈ A is an excitation of ω; this follows from the

cluster property.

One can define a one-particle state (a one-particle excitation of

the state ω or elementary excitation of ω) as a generalized H-valued

function Φ(p) obeying PΦ(p) = pΦ(p), HΦ(p) = ε(p)Φ(p). (More

precisely, for some class of test functions f(p), we should have

a linear map f → Φ(f) of this class into H obeying PΦ(f) =

Φ(pf), HΦ(f) = Φ(ε(p)f), where ε(p) is a real-valued function

called dispersion law. For definiteness, we can assume that test

functions belong to the Schwartz space S(Rd).) Let us fix an element

B ∈ A such that BΦ = Φ(φ). (We assume that φ is smooth and does

not vanish anywhere.) Recall that we consider H as a pre-Hilbert

Page 388: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 341

Methods of Quantum Field Theory in Statistical Physics 341

space obtained by the GNS construction, therefore, every element of

H can be represented in the form BΦ.

Note that for every function g(x, t) ∈ S(Rd+1), we have

B(g)Φ = Φ(gf ), (13.11)

where B(g) =∫g(x, t)B(x, t)dxdt, gf (p) = g(p,−ε(p))f(p), and g

stands for the Fourier transform of g.

It follows that we can always assume that B =∫α(x, t)

A(x, t)dxdt, where α(x, t) ∈ S(Rd+1), A ∈ A. We will say that an

operator B satisfying this assumption and transforming Φ into one-

particle state is a good operator. For a good operator, B(x, t) =∫α(x′ − x, t′ − t)A(x′, t′)dx′dt′, hence this expression is a smooth

function of x and t.

We assume that Φ(f) is normalized (i.e. 〈Φ(f),Φ(f ′)〉 = 〈f, f ′〉).Note that it is possible that there exist several types of (quasi-)

particles Φr(f) with different dispersion laws. We say that the space

spanned by Φr(f) is one-particle space and denote itH1. If the theory

is invariant with respect to spatial rotations, then the infinitesimal

rotations play the role of components of angular momentum. If d = 3,

a particle with spin s can be described as a collection of 2s + 1

functions Φr(p) obeying PΦr(p) = pΦr(p), HΦr(p) = ε(p)Φr(p).

The group of spatial rotations acts in the space spanned by these

functions; this action is a tensor product of the standard action

of rotations of the argument and irreducible (2s + 1)-dimensional

representation. (Here, s is half-integer, the representation is two-

valued if s is not an integer.) In relativistic theory, an irreducible

subrepresentation of the representation of the Poincare group in Hspecifies a particle with spin.

13.3.3 Scattering

Let us assume that we have several types of (quasi-)particles defined

as generalized functions Φk(p) obeying PΦk = pΦk(p), HΦk(p) =

εk(p)Φk(p), where the functions εk(p) are smooth and strictly

convex. Take some good operators Bk ∈ A, obeying BkΦ = Φ(φk).

Define Bk(f, t), where f is a function of p as∫f(x, t)Bk(x, t)dx and,

Page 389: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 342

342 Mathematical Foundations of Quantum Field Theory

as earlier, f(x, t) is a Fourier transform of f(p)e−iεk(p)t with respect

to p.

Let us consider the vectors

Ψ(k1, f1, . . . , kn, fn | t1, . . . , tn) = Bk1(f1, t1) · · · Bkn(fn, tn)Φ.

(13.12)

We assume that f1, . . . , fn have compact support. Let us introduce

the notation vi(p) = ∇εki(p). The set of all vectors vi(p) with

fi(p) 6= 0 will be denoted Ui. We assume that the sets Ui (the closures

of Ui) do not overlap. This assumption will be called NO condition

in what follows.

Then assuming the strong cluster property or asymptotic com-

mutativity (precise formulation will be given below), one can prove

that the vector (13.12) has a limit as ti → ∞ or ti → −∞. The set

spanned by these limits will be denoted D+ or D−.Note that the assumption that the sets Ui do not intersect (NO

condition) can be omitted if the space–time dimension is≥ 4. In these

dimensions, we can drop the NO condition defining the sets D±.The existence of the limit of the vectors (13.12) allows us to

define Møller matrices. We introduce the space Has as a Fock

representation for the operators a+k (f), ak(f) obeying canonical

commutation relations (CCR).

We define Møller matrices S− and S+ as operators defined on Hasand taking values in H by the formula

Ψ(k1, f1, . . . , kn, fn | ±∞) = S±(a+k1

(f1φk1) · · · a+kn

(fnφkn)θ).

(13.13)

This formula specifies S± on a dense subset of the Hilbert space Has.These operators are isometric, hence they can be extended to Hasby continuity.

One can say that the vector

e−iHtΨ(k1, f1, . . . , kn, fn | ±∞)

= Ψ(k1, f1e−iεk1 t, . . . , kn, fne

−iεkn t | ±∞) (13.14)

describes the evolution of a state corresponding to a collection

of n particles with wave functions f1φ1e−iεk1 t, . . . , fnφne

−iεkn t as

t→ ±∞.

Page 390: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 343

Methods of Quantum Field Theory in Statistical Physics 343

Note that we use the notation ak(f) =∫f(p)ak(p)dp, a+

k (f) =

(ak(f))∗ =∫f(p)a+

k (p)dp, hence a+k (f) =

∫f(p)a+(p)dp.

The existence of the limit in (13.12) can be derived from the strong

cluster property if we impose an additional condition B∗kΦ = 0 or the

condition B∗kΦ = Φk(νk).

Let us sketch the proof that the limit in (13.12) exists, assuming

for simplicity that the times ti are equal: ti = t. Let us denote the

LHS of (13.12) as Ψ(t). It is sufficient to prove that the norm of the

derivative of this vector with respect to t is a summable function:∫‖Ψ(t)‖dt <∞.Note that ‖Ψ(t)‖2 can be expressed in terms of correlation

functions, hence in terms of truncated correlation functions. If we

assume the strong cluster property, then ‖Ψ(t)‖ tends to zero faster

than any power of |t|, this implies the existence of the limit and the

fact that ‖Ψ(±∞)−Ψ(t)‖ also tends to zero faster than any power.

If we are interested only in the existence of the limit, it is sufficient

to assume a weaker version of cluster property (three continuous

derivatives of the function νn). This is sufficient to prove that

‖Ψ(t)‖ < C|t|−32 . Every factor in an arbitrary term of the expansion

for ‖Ψ(t)‖2 has the form

Ik,l(t) = 〈( ˙Bi1(fi1 , t))

∗ · · · ( ˙Bik(fik , t))

× ˙Bj1(fj1 , t) · · ·

˙Bjl(fjl , t)〉

T ,

where the dots above the operators denote differentiation with

respect to time, which can enter in one of the first k operators and in

one of the last l operators. We can suppose without loss of generality

that k ≥ 1, l ≥ 1 (otherwise Ik,l = 0; this follows from the remark

that one-particle states are orthogonal to Φ.). If each factor in the

given term has k = l = 1, then the term is equal to zero because one

of the factors either has the form

〈( ˙Bi(fi, t))

∗ ˙Bj(fj , t)〉T = 〈( ˙

Bi(fi, t))∗ ˙Bj(fj , t)Φ,Φ〉,

or the form

〈(Bi(fi, t))∗Bj(fj , t)〉T = 〈(Bi(fi, t))∗Bj(fj , t)Φ,Φ〉,

and both expressions are zero.

Page 391: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 344

344 Mathematical Foundations of Quantum Field Theory

In all other cases, the factor tends to zero as t → ∞ faster than

any power of t. This follows from Lemma 10.2 of Section 10.1.

If the space–time dimension is ≥ 4, we can prove the existence

of the limit of vectors (13.12) without the NO condition; the proof

generalizes the considerations used in Section 10.1 for d = 3.

The definition of S± that we gave specifies these operators as

multi-valued maps (for example, we can use different good operators

in the construction and it is not clear whether we get the same

answer). However, we can check that this map is isometric and every

multi-valued isometric map is really single-valued (see Section 10.1).

In particular, this means that the definition does not depend on the

choice of good operators.

To prove that the map is isometric, we express the inner product

of two vectors of the form Ψ(t) in terms of truncated correlation

functions. Only two-point truncated correlation functions survive in

the limit t → ±∞. This allows us to say that the map is isometric

(see Section 10.1 for a more detailed proof in a slightly different

situation).

Let us define in- and out-operators by the formulas

ain(f)S− = S−a(f), a+in(f)S− = S−a

+(f),

aout(f)S+ = S+a(f), a+out(f)S+ = S+a

+(f).

(For simplicity of notations, we consider the case when we have only

one type of particles. If we have several types of particles, in- and

out-operators as well as the operators a+, a are labeled by a pair

(k, f), where f is a test function and k characterizes the type of

particle.) These operators are defined on the image of S− and S+

correspondingly. One can check that

a+in(f φ) = lim

t→−∞B(f, t), a+

out(f φ) = limt→∞

B(f, t). (13.15)

The limit is understood as a strong limit. It exists on the set of

vectors of the form (13.13) (with NO assumption for d < 3). The

proof follows immediately from the fact that taking the limit ti →∞

Page 392: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 345

Methods of Quantum Field Theory in Statistical Physics 345

in the vectors (13.12), we can first take the limit for i > 1 and then

the limit t1 →∞.The formula (13.15) can be written in the following way:

a+in(f φ)Ψ(f1, . . . , fn | −∞) = Ψ(f, f1, . . . , fn | −∞),

a+out(f φ)Ψ(f1, . . . , fn |∞) = Ψ(f, f1, . . . , fn |∞).

Similarly,

ain(f)Ψ(φ−1f, f1, . . . , fn | −∞) = Ψ(f1, . . . , fn | −∞), (13.16)

aout(f)Ψ(φ−1f, f1, . . . , fn |∞) = Ψ(f1, . . . , fn |∞). (13.17)

If the operators S+ and S− are unitary, we say that the theory has

a particle interpretation. In this case (and also in the more general

case when the image of S− coincides with the image of S+), we can

define the scattering matrix

S = S∗+S−.

The scattering matrix is a unitary operator in Has. Its matrix

elements in the basis |p1, . . . , pn〉 = 1n!a

+(p1) · · · a+(pn)θ (scattering

amplitudes) can be expressed in terms of in- and out-operators.

Smn(p1, . . . ,pm |q1, . . . ,qn)

= 〈a+in(q1) · · · a+

in(qn)Φ, a+out(p1) · · · a+

out(pm)Φ〉. (13.18)

Effective cross-sections can be expressed in terms of the squares

of scattering amplitudes.

Note that only when ω is a ground state can one hope that the

particle interpretations exists. In other cases, instead of a scattering

matrix and cross-sections, one should consider inclusive scattering

matrix and inclusive cross-sections (see the following).

Formula (13.18) is proved only for the case when all values of

momenta pi,qj are distinct. More precisely, this formula should

be understood in the sense of generalized functions and as test

functions we should take collections of functions fi(pi), gj(qj) with

Page 393: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 346

346 Mathematical Foundations of Quantum Field Theory

non-overlapping U(fi), U(gj). Let us write this in more detail:

Smn(f1, . . . , fm | g1, . . . , gn)

=

∫dmpdnq

∏fi(pi)

∏gj(qj)Smn(p1, . . . ,pm |q1, . . . ,qn)

= 〈a+in(g1) · · · a+

in(gn)Φ, a+out(f1) · · · a+

out(fm)Φ〉.

Using (13.15), we obtain

Smn(f1, . . . , fm | g1, . . . , gn)

= limt→∞,τ→−∞

〈B(fmφ−1, t)∗ · · · B(f1φ

−1, t)∗B(g1φ−1, τ)

· · · B(gnφ−1, τ))Φ,Φ〉

= limt→∞,τ→−∞

ω(B(fmφ−1, t)∗ · · ·B(f1φ

−1, t)∗B(g1φ−1, τ)

· · · B(gnφ−1, τ)),

where B(f, t)∗ =∫dxB∗(x, t)f(x, t).

Note that in the same way, we can obtain a more general formula

Smn(f1, . . . , fm | g1, . . . , gn)

= limt→∞,τ→−∞

〈Bm+1(f1φ−1m+1, τ)

· · · Bm+n(gnφ−1m+n, τ))Φ, B1(f1φ

−11 , t) · · ·Bm(fmφ

−1m , t)Φ〉

= limt→∞,τ→−∞

ω(Bm(fmφ−1m , t)∗ · · ·B1(f1φ

−11 , t)∗Bm+1(g1φ

−1m+1, τ)

· · ·Bm+n(gnφ−1m+n, τ)), (13.19)

where Bi are different good operators and BiΦ = Φ(φi).

13.3.4 Asymptotic behavior of 〈Q(x, t)Ψ,Ψ′〉

Our next goal is to calculate the asymptotic behavior as t → ∞ of

the expression 〈Q(x, t)Ψ,Ψ′〉 where Q ∈ A.

Page 394: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 347

Methods of Quantum Field Theory in Statistical Physics 347

We assume that the NO condition is satisfied and that vectors

Ψ,Ψ′ belong to D+. It is sufficient to consider the case

Ψ = limt→∞

Ψ(t),Ψ(t) = Ψ(f1, . . . , fn | t),

Ψ′ = limt→∞

Ψ′(t),Ψ(t) = Ψ(f ′1, . . . , f′m | t).

(To simplify the notations, we consider the case of one type of

particles.) We assume that the strong cluster property is satisfied.

Then the differences between Ψ and Ψ(t), and between Ψ′ and Ψ′(t)

are negligible for t → ∞ (we neglect terms tending to zero faster

than any power of t, more precisely terms that are less than Cnt−n

where Cn does not depend on x). We see that it is sufficient to study

the asymptotic behavior of

〈Q(x, t)Ψ(t),Ψ′(t)〉

= 〈Q(x, t)B(f1, t) · · ·B(fn, t)Φ, B(f ′1) · · ·B(f ′1, t) · · ·B(f ′m, t)Φ〉

= ω(B∗(f ′m, t) · · ·B∗(f ′1, t)Q(x, t)B(f1, t) · · ·Bn(fn, t)).

We decompose this expression in terms of truncated correlation

functions. It is clear that every non-negligible truncated function

should contain equal number of B’s and B∗’s. Using this remark

and (13.17), we obtain

〈Q(x, t)Ψ,Ψ′〉 = 〈Ψ,Ψ′〉M +

∫dpdp′〈aout(p)Ψ, aout(p

′)Ψ′〉N(p,p′)

+

∫dp〈aout(p)Ψ,Ψ′〉T1(p)

+

∫dp′〈Ψ, aout(p

′)Ψ′〉T2(p′) +R, (13.20)

where R is negligible,

M = ω(Q) = 〈QΦ,Φ〉, N(p,p′) = 〈Q(x, t)Φ(p),Φ(p′)〉,

T1(p) = 〈Q(x, t)Φ(p),Φ〉, T2(p′) = 〈Q(x, t)Φ,Φ(p′)〉.

Similar formulas can be written for t→ −∞.

Page 395: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 348

348 Mathematical Foundations of Quantum Field Theory

13.3.5 Scattering theory from asymptotic

commutativity

In this section, we derive the existence of limit in (13.12) from

asymptotic commutativity of the algebra A(ω). This approach is

useful in relativistic theory where asymptotic commutativity follows

from locality.

Let us assume that the operators Bk and Bl asymptotically

commute:

‖[Bk(x, t), Bl(x′, t)]‖ <C

1 + |x− x′|a,

where a > 1. We also suppose the condition of asymptotic com-

mutativity is satisfied for the spatial and time derivatives of these

operators.

Instead of these conditions, we can assume that Bk =∫gk(x, t)Ak(x, t)dxdt and

‖[Ak(x, t), Al(x′, t′)]‖ <C(t− t′)

1 + |x− x′|a,

where gk are functions from Schwartz space, C(t) is a function of at

most polynomial growth and a > 1.

The statement we need follows from the following fact:∫‖[Bki(fi, t),

˙Bkj (fj , t)]‖dt <∞. (13.21)

To check this, we note that Ψ(t) is a sum of n terms; every term

contains a product of several operators B and one operator˙B. The

estimate for the norm of Ψ(t) follows from (13.21) and (13.11) and

the remark that the norms of operators Bki(fi, t) are bounded. (We

change the order of operators in the summand under consideration in

such a way that˙B is from the right. Then we note that

˙B(f, t)Φ = 0

as follows from (13.11).)

Equation (13.21) remains to be proved. First of all, we note that

one can prove the estimate

|fk(x, t)| ≤ C|t|−D2 (13.22)

(the proof for d = 3 given in Section 10.2 works for any d).

Page 396: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 349

Methods of Quantum Field Theory in Statistical Physics 349

Let us consider the integral∫dxdx′fki(x, t)fkj (x

′, t)[Bki(x, t),˙Bkj (x

′, t)]. (13.23)

First of all, we consider the integral (13.23) over the domain Γ(t)

defined in the following way. Take non-overlapping compact sets G ⊂Rd, G′ ⊂ Rd. We say that the point (x,x′) belongs to Γ(t) if there

exist points v ∈ G,v′ ∈ G′ such that |v − xt | < C|t|−ρ, |v′ − x′

t | <C|t|−ρ where ρ = 1

2−ε. (We will apply our estimate in the case when

G is the set of points of the form ∇εki(p) where p belongs to the

support of fki and G′ is defined in similar way using the function

fkj .) For large t, the volume of this domain is less than t2d(1−ρ) and

the norm of the integrand is less than t−d × t−a. (We have used the

fact that the distance between x and x′ grows linearly with t in the

integration domain, hence the norm of the commutator in (13.23) is

less than t−a.) This allows us to say that the norm of the integral

does not exceed t−a+ε.

To estimate the integral (13.21) over the complement to Γ(t), we

consider the integral∫exp(−iε(p)t+ ipx)µ

(tρ(v(p)− x

t

))f(p)dp, (13.24)

where µ(x) is a smooth function equal to zero for |x| ≤ ν1 and equal

to one for |x| ≥ ν2. One can prove that for every r, the absolute value

of this integral does not exceed C|t|−(1−2ρ)r and for |x| > bt, it does

not exceed C|x|−r|t|−(1−2ρ)r.

The integral (10.24) is equal to f(x, t) if tρ(v(p)− xt ) ≥ ν2 for p ∈

supp f. This means that the estimate of this integral can be applied

to at least one of the factors fki(x, t), fkj (x′, t) in the integrand of

the integral (13.23) on the complement to the domain Γ(t). Using

this fact and the inequality∫dx|fk(x, t)| ≤ C|t|

d2 ,

we obtain that the norm of the integral (13.23) over the complement

to Γ(t) tends to zero faster than any power of |t| as t → ±∞.Combining this estimate with the estimate for the integral over Γ(t)

Page 397: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 350

350 Mathematical Foundations of Quantum Field Theory

and taking ε < a−1, we obtain the estimate for the norm of (13.23):

for large |t|, it is less than |t|−b where b > 1. The same arguments

work if the dot (time derivative) in (13.23) is removed.

To prove (13.21), we note that calculating˙B(f, t) we should take

into account that f depends on t. It is easy to check that

˙B(f, t) =

∫dxg(x, t)B(x, t) +

∫dxf(x, t)

˙B(x, t),

where g(p) = −iε(p)f(p). Using this expression, we obtain that the

commutator in (13.21) can be written as a sum of (13.23) and similar

expression with removed dots. Both summands do not exceed |t|−bwith b > 1 for large |t|. This implies (13.21).

Note that the proof based on asymptotic commutativity also

allows us to estimate the speed of convergence to the limit. In the

case of strong asymptotic commutativity, the difference between Ψ(t)

and Ψ(±∞) tends to zero faster than any power of t.

13.3.6 Green functions and scattering: LSZ

Let us start with scattering of elementary excitations of ground state

(of particles). In this case, the scattering matrix can be expressed in

terms of on-shell values of Green functions. The Green function in

translation-invariant stationary state ω is defined by the formula

Gn = ω(T (A1(x1, t1) · · ·Ar(xr, tr))),

where Ai ∈ A and T stands for time ordering. More precisely,

this is a definition of Green function in (x, t)-representation, taking

Fourier transform with respect to x, we obtain Green functions

in (p, t)-representation; taking in these functions inverse Fourier

transform with respect to t, we obtain Green functions in in (p, ε)-

representation. Due to translation-invariance of ω, we obtain that

in (x, t)-representation, the Green function depends on differences

xi − xj , in (p, t)-representation, it contains a factor δ(p1 + · · ·+ pr).

Similarly, in (p, ε), we have the same factor and the factor

δ(ε1 + · · ·+ εr). We omit both factors talking about poles of Green

functions.

Page 398: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 351

Methods of Quantum Field Theory in Statistical Physics 351

For n = 2 in (p, ε)-representation, G2 has the form

G(p1, ε1 |A,A′)δ(p1 + p2)δ(ε1 + ε2).

Poles of function G(p, ε |A,A′) correspond to particles, the depen-

dence of the position of the pole on p specifies the dispersion law

ε(p) (we consider poles with respect to the variable ε for fixed p).

This follows from Kallen–Lehmann representation (Section 8.1), but

this can be obtained also from the following considerations.

We will prove that the scattering amplitudes can be obtained

as on-shell values of Green functions (LSZ formula). To simplify

notations, we consider the case when we have only one single-

particle state Φ(p) with dispersion law ε(p). We assume that the

elements Ai ∈ A are chosen in such a way that the projection of

AiΦ on the one-particle space has the form Φ(φi) =∫φi(p)Φi(p)dp

where φi(p) is a non-vanishing function. We introduce the notation

Λi(p) = φi(p)−1.

Let us consider Green function

Gmn = ω(T (A∗1(x1, t1) · · ·A∗m(xm, tm)Am+1(xm+1, tm+1)

· · ·Am+n(xm+n, tm+n))

in (p, ε)-representation. It is convenient to change slightly the defini-

tion of (p, t)- and (p, ε)-representation changing the signs of variables

pi and εi for 1 ≤ i ≤ m (for variables corresponding to the operators

A∗i ). This convention agrees with the conventions in Section 9.2.

Multiplying the Green function in (p, ε)-representation by∏1≤i≤m

Λi(pi)(εi + ε(pi))∏

m<j≤m+n

Λj(pj)(εj − ε(pj)),

and taking the limit εi → −εi(pi) for 1 ≤ i ≤ m and the limit εj →εj(pj) for m < j ≤ m+n, we obtain on-shell Green function denoted

by σmn. We prove that it coincides with scattering amplitudes:

σm,n(p1, . . . ,pm+n) = Smn(p1, . . . ,pm |pm+1, . . . ,pm+n). (13.25)

First of all, we note that the on-shell Green function can be expressed

in terms of the asymptotic behavior of the Green function in (p, t)-

representation: if for t→ ±∞ and fixed p, this behavior is described

Page 399: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 352

352 Mathematical Foundations of Quantum Field Theory

by linear combination of exponent eiλt, then the location of poles is

determined by the exponent indicators λ and the coefficients in front

of exponents determine the residues. Using this statement, one can

show that

limt→∞,τ→−∞

ω(Bm(fmφ−1m , t)∗ · · ·B1(f1φ

−11 , t)∗Bm+1(g1φ

−1m+1, τ)

· · ·Bm+n(gnφ−1m+n, τ)), (13.26)

where Bi are good operators can be expressed in terms of on-shell

Green function as∫dm+npf1(p1) . . . fm(pm)g1(−pm+1)

. . . gn(−pm+n)σm,n(p1, . . . ,pm+n).

Using (13.19), we obtain (13.25) in the case when Ai are good

operators. We will show that the general case can be reduced to

the case when operators Ai are good.

Let us suppose that one-particle spectrum does not overlap with

multi-particle spectrum. (We represent the space H as a direct sum

of one-dimensional space H0 spanned by Φ, one-particle space H1

and the space M called multi-particle space. Our condition means

that the joint spectrum of P and H in M (multi-particle spectrum)

does not overlap with the joint spectrum of P and H in H0 + H1.

If the theory has particle interpretation, then our condition means

that ε(p1 +p2) < ε(p1)+ε(p2).1 Then we can construct an operator

B transforming Φ into one-particle state (a good operator) using the

formula B =∫α(x, t)A(x, t)dxdt where A ∈ A and the projection

of AΦ onto one-particle state does not vanish. Namely, we should

assume that the support of α(p, ω) (of the Fourier transform of α)

does not intersect the multi-particle spectrum and does not contain 0.

Moreover, the operator we constructed obeys B∗Φ = 0.

1The physical meaning of this condition: the energy conservation law forbids thedecay of a particle. This condition is not always satisfied, however, stability of aparticle is always guaranteed by some conservation laws. Our considerations canbe applied in this more general situation.

Page 400: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 353

Methods of Quantum Field Theory in Statistical Physics 353

Now, we can apply this construction to the operators Ai and

verify that on-shell Green functions corresponding to good operators

Bi coincide with Green functions corresponding to the operators

Ai. (We note that the correlation functions of operators Bi can be

expressed in terms of correlation functions of operators Ai. As we

mentioned already, on-shell Green function can be expressed in terms

of asymptotic behavior as t → ±∞ of the Green function in (p, t)-

representation. It is easy to find the relation between this behavior

for the Green functions of operators Ai and the behavior for the

Green functions of operators Bi.)

For relativistic theories with a mass gap, one can derive asymp-

totic commutativity and cluster property from Haag–Araki or Wight-

man axioms. This proves the existence of scattering matrix. One can

prove that scattering matrix is Lorentz-invariant; let us sketch the

proof of this fact.

The operator B(f, t) used in our construction can be written as an

integral of f(x, t)B(x, t), where f(x, t) is a positive frequency solution

of the Klein–Gordon equation over the hyperplane t = constant . We

will define the operator B(f, ρ) integrating the same integrand over

another hyperplane ρ. One can replace the operators B(f, t) by the

operators B(f, ρ) in (13.12) and prove that the expression we obtained

has a limit if the hyperplanes ρi tend to infinity in time direction (for

example, if they have form αt+ax = constant and the constant tends

to infinity). This means that we can use B(f, ρ) in the definition of

Møller matrices and scattering matrix. The same arguments that were

used to prove that the limit does not depend on the choice of good

operators can be applied to verify that the new construction gives

the same Møller matrices. This implies Lorentz invariance because

the Lorentz group acts naturally on operators B(f, ρ).

13.3.7 Generalized Green functions; the inclusive

scattering matrix

Let us define generalized Green functions (GGreen functions) in the

state ω by the following formula where Bi ∈ A:

Gn = ω(MN),

Page 401: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 354

354 Mathematical Foundations of Quantum Field Theory

where

N = T (B1(x1, t1) · · ·Bn(xn, tn))

stands for chronological product (times decreasing) and

M = T opp(B∗1(x′1, t′1) · · ·B∗n(x′n, t

′n))

stands for antichronological product (times increasing).

One can give another definition of GGreen functions introducing

the operator

Q = T (B1(x1, t1) · · ·Bn(xn, tn)B1(x′1, t′1) · · · Bn(x′n, t

′n)),

where the operators Bi, Bi act on the space of linear functionals on

A. (Recall (13.8) and (13.9) that operators B and B act on linear

functionals defined on A; they transform ω(A) into ω(BA) and in

ω(AB∗) correspondingly.) It is easy to check that

Gn = (Qω)(1).

Let us define inclusive S-matrix as on-shell GGreen function. We

will show that in the case when the theory has particle interpretation,

inclusive cross-section can be expressed in terms of inclusive S-

matrix.

Recall that the inclusive cross-section of the process (M,N) →(Q1, . . . , Qm) is defined as a sum (more precisely, a sum of integrals)

of effective cross-sections of the processes (M,N) → (Q1, . . . , Qm,

R1, . . . , Rn) over all possible R1, . . . , Rn. If the theory does not have

particle interpretation, this formal definition of inclusive cross-section

does not work, but still the inclusive cross-section can be defined in

terms of probability of the process (M,N → (Q1, . . . , Qn+ something

else)) and expressed in terms of inclusive S-matrix.

Let us consider the expectation value

ν(a+out,k1

(p1)aout,k1(p1) · · · a+out,km

(pm)aout,km(pm)), (13.27)

where ν is an arbitrary state. This quantity is the probability density

in momentum space for finding m outgoing particles of the types

k1, . . . , kn with momenta p1, . . . ,pm plus other unspecified outgoing

particles. It gives inclusive cross-section if ν = ν(t) describes the

evolution of a state represented as a collection of incoming particles.

Page 402: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 355

Methods of Quantum Field Theory in Statistical Physics 355

It will be convenient to consider more general expectation value

ν(a+out,k1

(f1)aout,k1(g1) · · · a+out,km

(fm)aout,km(gm)), (13.28)

where fi, gi are test functions. This expression can be understood as

a generalized function

ν(a+out,k1

(p1)aout,k1(q1) · · · a+out,km

(pm)aout,km(qm));

we get (13.27) taking pi = qi.

As earlier, we assume that we have several types of (quasi-)particles

Φk(p) obeying PΦk = pΦk(p), HΦk(p) = εk(p)Φk(p), where the

functions εk(p) are smooth and strictly convex. Good operators

Bk ∈ A obey BkΦ = Φ(φk). The operator Bk(f, t), where f is a

function of p, defined as∫f(x, t)Bk(x, t)dx (as always f(x, t) is a

Fourier transform of f(p)e−iεk(p)t with respect to p).

Now we can calculate (13.28). First of all, we take as ν the state

corresponding to the vector (13.14). We are representing this vector

in terms of good operators Bi using the formula

Ψ(k1, f1, . . . , kn, fn |∞) = limt→−∞

Ψ(k1, f1, . . . , kn, fn | t)

= limt→−∞

Bk1(f1, t) · · · Bkn(fn, t)Φ.

The corresponding state ν considered as linear functional on A can

be expressed in terms of the state ω corresponding to Φ. (We should

use the remark that the state corresponding to the vector AΦ can

be written as AAω.) Expressing the out-operators by the formula

(13.15), we obtain the expression of (13.28) in terms of GGreen

functions on-shell.

13.4 L-functionals

In the approach of Section 13.2, for every equilibrium state, we should

construct a Hilbert space depending on the temperature, CCR or

CAR are represented in this space, the equilibrium state is described

by a vector in the space. In this section, we will describe a formalism

of L-functionals (positive functionals on Weyl algebra) that can be

used to describe the states corresponding to vectors and density

Page 403: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 356

356 Mathematical Foundations of Quantum Field Theory

matrices in all representations of CCR. This formalism can be used to

calculate physical quantities for equilibrium state in the framework

of an analog of Feynman diagram technique.

Note that there exists an obvious generalization of all these results

to the case of fermions. In this case, L-functionals are positive

functionals on the algebra with generators satisfying canonical

anticommutation relations (CAR) (Clifford algebra), α and α∗ in

the definition of L-functional are anticommuting variables.

Let us consider a representation of CCR in Hilbert space H. Here,

we understand CCR as relations

[ak, a+l ] = ~δkl, [ak, al] = [a+

k , a+l ] = 0,

where k, l run over a discrete set M. To a density matrix K (or

more generally, to any trace class operator in H), we can assign a

functional LK(α∗, α) defined by the formula

LK(α∗, α) = Tr e−αa+eα∗aK. (13.29)

Here, αa+ stands for∑αka

+k and α∗a for

∑α∗kak, where k runs

over M. This formula makes sense if αα∗ =∑|αk|2 < ∞. (This

follows from the remark that e−αa++α∗a = e~αα

∗/2e−αa+eα∗a is a

unitary operator in H). We can apply (13.29) also in the case when

K is an arbitrary operator of trace class (not necessarily a density

matrix).

One can say that LK is a generating functional of correlation

functions.

One can also consider a more general case when CCR are written

in the form

[a(k), a+(k′)] = ~δ(k, k′), [a(k), a(k′)] = [a(k)+, a(k′)+] = 0,

k, k′ run over a measure space M . We are using the exponential

form of CCR; in this form, a representation of CCR is specified

as a collection of unitary operators e−αa++α∗a obeying appropriate

commutation relations. Here, α(k) is a complex function on the

measure space M , the expressions of the form α∗a, αa+ can be

written as integrals∫α∗(k)a(k)dk,

∫α(k)a+(k)dk over M. In the

space of CCR a(k), a+(k′) become Hermitian conjugate generalized

Page 404: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 357

Methods of Quantum Field Theory in Statistical Physics 357

operator functions. We always assume that α is square integrable,

then the expression (13.29) is well defined.

Knowing LK , we can calculate 〈A〉L = TrAK where A is an

element of the unital associative algebra A generated by ak, a+l

(Weyl algebra). Hence, we can consider LK as a positive linear

functional 〈A〉L on the algebra A. (We consider A as an algebra

with involution +, positivity means that 〈A+A〉L ≥ 0. Recall that

positive linear functional ω on unital algebra with involution is called

a normalized state if ω(1) = 1.)

A functional L(α∗, α) satisfying the condition L(−α∗,−α) =

L∗(α, α∗) and positivity condition is called physical L-functional. We

say that such a functional is normalized if L(0, 0) = 1. Normalized

physical L-functionals are in one-to-one correspondence with states

on A. We will consider also the space L of all functionals L(α∗, α) (of

all L-functionals). Every trace class operator K in a representation

space of CCR specifies an element of L by the formula (13.29).

We can define an antilinear involution L → L on the space L of

functionals L(α∗, α) by the formula

(L)(α∗, α) = L∗(−α,−α∗). (13.30)

Physical L-functionals are invariant with respect to this involution.

It is easy to check that

LK = LK+ .

Every normalized vector Φ ∈ H specifies a density matrix

and hence a physical L-functional. Conversely, every physical

L-functional corresponds to a vector in some representation of CCR,

given by GNS construction (see the preceding section). One can

characterize this representation by the requirement that it contains

a cyclic vector Φ obeying L(A) = 〈AΦ,Φ〉. However, an L-functional

can be obtained from many density matrices in many representations

of CCR.

An action of Weyl algebra A on L can be specified by operators

b+(k) = ~c+1 (k)− c2(k), b(k) = c1(k)

obeying CCR. Here, c+i (k) are multiplication operators by α∗k, αk

and ci(k) are derivatives with respect to α∗k, αk. This definition is

Page 405: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 358

358 Mathematical Foundations of Quantum Field Theory

prompted by relations

La(k)K = b(k)LK , La+(k)K = b+(k)LK . (13.31)

Applying the involution L→ L, we obtain another representation of

A on L. It is specified by the operators

b+(k) = −~c+2 (k) + c1(k), b(k) = −c2(k),

obeying CCR and satisfying

LKa+(k) = b(k)LK , LKa(k) = b+(k)LK . (13.32)

More generally, for any A ∈ A, we have an operator acting in

L denoted by the same symbol and obeying A(LK) = LAK (the

left multiplication by A in A specifies the action of A on linear

functionals). We can define also an operator A using the formula

A(L) = A(L). If L = LK , then AL = LKA+ . It is easy to check that

A+B = A+ B, AB = AB, AB = BA. Operators A specify another

action of Weyl algebra on L that commutes with the original action;

one can say that the direct sum of two Weyl algebras acts on L.

Note that the space of physical L-functionals is not invariant with

respect to the operators A and A, however, it is invariant with respect

to the operators AA where A ∈ A. It is easy to check that

AALK = LAKA+ .

Let us consider a Hamiltonian H in a space of representation of

CCR. We will write H in the form

H =∑m,n

∑ki,lj

Hm,n(k1, . . . , km | l1, . . . , ln)a+k1· · · a+

kmal1 · · · aln .

(13.33)

There are two operators in L corresponding to H:

H =∑m,n

∑ki,lj

Hm,n(k1, . . . , km | l1, . . . , ln)b+k1 · · · b+kmbl1 · · · bln

(13.34)

Page 406: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 359

Methods of Quantum Field Theory in Statistical Physics 359

(we denote it by the same symbol) and

H =∑m,n

∑ki,lj

Hm,n(k1, . . . , km | l1, . . . , ln)b+k1 · · · b+kmbl1 · · · bln .

(13.35)

The equation of motion for the L-functional L(α∗, α) has the form

i~dL

dt= HL = HL− HL. (13.36)

(We introduced the notation H = H − H.) It corresponds to the

equation of motion for density matrices; this follows from the formula

HLK = LHK−KH+ .

Note that often the equations of motion for L-functionals make

sense even in the situation when the equations of motion in the

Fock space are ill defined. This is related to the fact that vectors

and density matrices from all representations of CCR are described

by L-functionals. This means that by applying the formalism of

L-functionals, we can avoid the problems related to the existence of

inequivalent representations of CCR. It is well known, in particular,

that these problems arise for translation-invariant Hamiltonians; in

perturbation theory, these problems appear as divergences related to

infinite volume. Therefore, in the standard formalism, it is necessary

to consider at first a Hamiltonian in finite volume Ω (to make volume

cutoff or, in another terminology, infrared cutoff) and to take the

limit Ω → ∞ in physical quantities (see Sections 8.2 and 13.2 for

more details). In the formalism of L-functionals, we can work directly

in infinite volume.

In general, the ground state of formal Hamiltonian (13.7) does

not belong to Fock space, but the corresponding L-functional is well

defined. (Note that we always assume that ultraviolet divergences are

absent.) Similarly, the equilibrium states for different temperatures

belong to different Hilbert spaces, but all of them are represented by

well-defined L-functionals.

In what follows, we consider (13.33) as a formal expression;

we assume that it is formally Hermitian. There is no necessity to

Page 407: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 360

360 Mathematical Foundations of Quantum Field Theory

assume that (13.33) specifies a self-adjoint operator in one of the

representations of CCR.

Let us make some remarks about adiabatic evolution in the

formalism of L-functionals. Let us take a Hamiltonian that depends

on time t, but is changing adiabatically (very slowly). More formally,

we can assume that the Hamiltonian depends on a parameter g and

g = h(at) where a → 0. Let us take a family of L-functionals L(g),

where L(g) is a stationary state of the Hamiltonian H(g), i.e.

H(g)L(g) = 0.

It is obvious that L(h(at)) obeys the equation of motion up to

terms tending to zero as a → 0. (Note that a similar statement

is wrong in the standard Hilbert space formulation of quantum

mechanics because the vector corresponding to a state is defined up to

phase factor. We can only say that for a smooth family of eigenvectors

Ψ(g), there exists a phase factor C(g) such that C(h(at))Ψ(h(at))

obeys the equation of motion up to terms tending to zero as a→ 0.)

13.4.1 Translation-invariant Hamiltonians in the

formalism of L-functionals; one-particle states

In what follows, we consider translation-invariant Hamiltonians.

We say that an L-functional σ is an excitation of a translation-

invariant L-functional ω if it coincides with ω at infinity. More

precisely, we should require that σ(Txα∗, Txα)) → ω(α∗, α) as

x→∞. Here, Tx stands for spatial translation.

We assume that the L-functional ω has cluster property. The

weakest form of cluster property is the requirement that 〈ATxB〉ω −〈A〉ω〈B〉ω tends to zero as x→∞ for A,B ∈ A.

If Φ is a vector corresponding to ω in GNS representation space

H, then the state corresponding to any vector AΦ where A ∈ A is

an excitation of ω; this follows from cluster property.

We can define elementary excitations (called particles if ω is a

ground state and quasiparticles if ω is a general translation-invariant

stationary state) in the same way as in Section 13.3.2.

Page 408: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 361

Methods of Quantum Field Theory in Statistical Physics 361

13.4.2 Quadratic Hamiltonians

Let us consider as an example the simplest translation-invariant

Hamiltonian

H0 =

∫ω(k)a+(k)a(k)dk, (13.37)

where k runs over Rd. It can be approximated by a Hamiltonian of

the form∑ωka

+k ak having a finite number of degrees of freedom.

The equilibrium state of the latter Hamiltonian can be represented

by density matrix Ω(T ) in the Fock space; it is easy to check that

akΩ(T ) = e−~ωkT Ω(T )ak, a

+k Ω(T ) = e

~ωkT Ω(T )a+

k . (13.38)

Applying (13.31) and (13.32), we obtain equations for the cor-

responding L-functional; taking the limit, we obtain for the

L-functional corresponding to the equilibrium state in infinite volume

c1(k)LT = e−~ω(k)T (−~c+

2 (k) + c1(k))LT , (13.39)

hence

c1(k)LT = −n(k)c+2 (k)LT , c2(k)LT = −n(k)c+

1 (k)LT , (13.40)

where

n(k) =~

e~ω(k)T − 1

. (13.41)

We obtain

LT = e−∫α∗(k)n(k)α(k)dk. (13.42)

If we are interested in equilibrium state for given density, we should

replace ω(k) with ω(k) − µ in (13.41) (here, µ stands for chemical

potential).

The Hamiltonian H governing the evolution of L-functionals can

be written in the form

H =

∫ω(k)b+(k)b(k)dk −

∫ω(k)b+(k)b(k)dk

=

∫~(ω(k)c+

1 (k)c+1 (k)dk − ω(k)c+

2 (k)c2(k))dk.

Page 409: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 362

362 Mathematical Foundations of Quantum Field Theory

It follows that the vector generalized functions Φ1(k) = c+1 (k)Φ,

Φ2(k) = c+2 (k)Φ in the space of GNS representation correspond-

ing to (13.42) are one-(quasi)particle excitations. (Note that in this

statement, n(k) in the formula (13.42) is an arbitrary function;

then the corresponding state is not an equilibrium, but it is still

a stationary translation-invariant state.)

Very similar considerations allow us to prove that the L-functional

corresponding to an equilibrium state of general quadratic Hamil-

tonian H = a+Ma+ + a+Na + aRa is Gaussian (has the form

eα∗Aα∗+α∗Bα+αCα).

13.4.3 Perturbation theory

Let us assume now that the Hamiltonian H is represented as a sum of

quadratic Hamiltonian H0 and interaction Hamiltonian Hint = gV.

Then in the formalism of L-functionals, one can introduce in the

standard way the evolution operator U(t, t0), the evolution operator

in the interaction picture S(t, t0) and the operator Sa, the analog

of adiabatic S-matrix. If Ha governs the evolution of L-functional

for the Hamiltonian H0 + h(at)Hint, then we denote by Ua(t, t0) the

operator transforming the L-functional at the moment t0 into the

L-functional at the moment t (the evolution operator). The operator

Sa(t, t0) is defined by the formula

Sa(t, t0) = ei~ H0tUa(t, t0)e−

i~ tH0t0 .

(We assume that h(t) is a smooth function equal to 1 in the

neighborhood of 0 and to 0 in the neighborhood of infinity. It is

increasing for negative t and decreasing for positive t.) The operator

Sa is defined as Sa(∞,−∞).

The perturbation theory for operators Sa(t, t0) can be constructed

in the standard way: we apply the formula

Sa(t, t0) = T exp

(− i~h(at)Hint(t)

).

(We use the notation A(t) = ei~ H0tAe−

i~ H0t.)

If we are interested only in the action of these operators on func-

tionals represented by polynomials of α, α∗ with smooth coefficients

Page 410: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 363

Methods of Quantum Field Theory in Statistical Physics 363

tending to zero at infinity (smooth functionals in the terminology of

Tyupkin (1973)), the diagram techniques can be described as follows.

The vertices come from − i~h(at)Hint. To find the propagator, we

calculate the T -product of two operators of the form b(+)i (t) and

express it in normal form with respect to the operators c+i , ci (i.e.

the operators ci are from the right). The propagator (that can be

considered as 4 × 4 matrix) is equal to the numerical part of this

expression. In other words, the propagator is given by the formula

〈T (b(k1, t1, σ1)b(k2, t2, σ2)〉L=1.

Here, b(k, 0, σ) is one of the operators b+, b, b+, b.

Let us define the adiabatic generalized Green functions (GGreen

functions) by the formula

Gan(k1, t1, σ1, . . . , kn, tn, σn)

= 〈T (b(k1, t1, σ1) · · · b(kn, tn, σn)Sa(∞,−∞))〉L=1 (13.43)

As usual, the perturbative expansion for these functions can be

constructed by the same rules as for adiabatic S-matrix, but the

diagrams have n external vertices.

Note that Sa(∞,−∞)1→ 1 and Sa(0,−∞)1→ L as a→ 0. Here,

L denotes the L-functional corresponding to the ground state of the

Hamiltonian H. (This follows immediately from similar statement

for the adiabatic evolution operators Ua(∞,−∞) and Ua(0,−∞)

and from the fact that the L-functional L = 1 corresponds to the

ground state of H0.) We obtain that the adiabatic GGreen function

Gan(k1, t1, σ1, . . . , kn, tn, σn) tends to the GGreen function

Gn(k1, t1, σ1, . . . , kn, tn, σn) = 〈T (b(k1, t1, σ1) · · ·b(kn, tn, σn))〉L(13.44)

as a→ 0.

(Here, we use the notation

b(k, t, σ) = S−1(t, 0)b(k, t, σ)S(t, 0) = U−1(t, 0)b(k, 0, σ)U(t, 0)

for the analog of Heisenberg operators.) This means that we can

construct the perturbation expansion for the GGreen function

Page 411: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 364

364 Mathematical Foundations of Quantum Field Theory

Gn(k1, t1, σ1, . . . , kn, tn, σn) taking the limit a → 0 in the diagrams

for Gan(k1, t1, σ1, . . . , kn, tn, σn). The only modification of diagrams is

in internal vertices: now, these vertices are governed by − i~Hint.

Similar procedure can be applied for the calculation of the action

of operators Sa(t, t0) on the space of functionals represented as a

product of a smooth functional and Gaussian functional Λ = eλ,

where λ is a quadratic expression in terms of α∗, α. We assume that

Λ is translation invariant and stationary with respect to the evolution

corresponding to the Hamiltonian H0, i.e. H0Λ = 0. (Here, H0 is a

translation-invariant quadratic Hamiltonian not necessarily of the

form (13.37).) It is easy to check imposing some non-degeneracy

conditions that there exist such operators c+i (k), ci(k), i = 1, 2

obeying CCR that Λ can be characterized as a functional satisfying

the conditions ci(k)Λ = 0, i = 1, 2,Λ(0, 0) = 0. Then the perturbative

expression for the action of Sa(t, t0) on the space under consideration

can be obtained by means of the diagram technique with propagators

described in the same way as for Λ = 1, the only difference is that

instead of normal form with respect to the operators c+i (k), ci(k),

we should consider normal form with respect to the operators

c+i (k), ci(k), i = 1, 2. Equivalently, we can define the propagator by

the formula

〈T (b(k1, t1, σ1)b(k2, t2, σ2)〉Λ.

Again, we can define adiabatic GGreen functions

Gan(k1, t1, σ1, . . . , kn, tn, σn)Λ

= 〈T (b(k1, t1, σ1) · · · b(kn, tn, σn)Sa(∞,−∞))〉Λ (13.45)

corresponding to Λ and introduce the diagram technique for their

calculation. The GGreen functions corresponding to Λ can be defined

either as limits of adiabatic GGreen functions as a → 0 or by the

formula

Gn(k1, t1, σ1, . . . , kn, tn, σn)Λ = 〈T (b(k1, t1, σ1) · · ·b(kn, tn, σn))〉Λ,(13.46)

where Λ = lima→0 Ua(0,−∞)Λ denotes the stationary state of

the Hamiltonian H that we obtain from the stationary state Λ of

Page 412: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 365

Methods of Quantum Field Theory in Statistical Physics 365

the Hamiltonian H0 adiabatically switching the interaction on. The

diagrams representing the functions (13.46) have n external vertices,

the internal vertices come from − i~Hint, the propagator is equal to

〈T (b(k1, t1, σ1)b(k2, t2, σ2)〉Λ.

In particular, we can take H0 of the form (13.37) and

Λ = e−∫α∗(k)n(k)α(k)dk. (13.47)

(All translation-invariant Gaussian functionals that are stationary

with respect to the Hamiltonian (13.37) have this form.) Then we

obtain the following formulas for the propagator:

〈T (b+(t)b(τ)〉Λ = θ(t− τ)ei~ω(k)(t−τ)n(k)

+ θ(τ − t)ei~ω(k)(t−τ)(n(k) + ~)r,

〈T (b+(t)b+(τ)〉Λ = ei~ω(k)(t−τ)(n(k) + ~),

〈T (b+(t)b(τ)〉Λ = 〈T (b+(t)b(τ)〉Λ,

〈T (b(t)b(τ)〉Λ = ei~ω(k)(t−τ)n(k).

All other entries vanish.

It follows from the above formulas that the diagrams we con-

structed coincide with the diagrams of Keldysh and TFD formalisms

(see Chu and Umezawa (1994) for review of both formalisms).

We have noted already that the L-functional corresponding to an

equilibrium state of quadratic Hamiltonian is Gaussian. Assuming

that the equilibrium state of the Hamiltonian H0 + gV can be

obtained from the equilibrium state of H0 by means of adiabatic

evolution, we can say that the diagram technique we have described

allows us to calculate the GGreen functions in the equilibrium state.

13.4.4 GGreen functions

We have constructed the diagram technique for GGreen functions.

As in the standard technique, we can express all diagrams in terms of

connected diagrams; moreover, connected diagrams can be expressed

in terms of 1 PI diagrams. (One says that a diagram is one particle

irreducible (1 PI) if it remains connected when we remove one of the

Page 413: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 366

366 Mathematical Foundations of Quantum Field Theory

edges. Calculating a 1 PI diagram, we do not take into account the

contributions of external edges.)

The contribution of a disconnected diagram is equal to the

product of the contributions of its components (up to some factor

taking into account the symmetry group of the diagram). The two-

point GGreen function

G2(k1, t1, σ1, k2, t2, σ2)Λ = 〈T (b(k1, t1, σ1)b(k2, t2, σ2))〉Λ (13.48)

obeys the Dyson equation

G2(k1, t1, σ1, k2, t2, σ2)Λ

= G2(k1, t1, σ1, k2, t2, σ2)Λ

+

∫dk′2dt

′2dσ

′2dk′′2dt′′2dσ

′′2G2(k1, t1, σ1, k

′2, t′2, σ′2)Λ

×M(k′2, t′2, σ′2, k′′2 , t′′2, σ′′2)Λ

×G2(k′′2 , t′′2, σ′′2 , k2, t2, σ2)Λ (13.49)

connecting it with the propagator and self-energy operator (mass

operator) M (the integration over discrete parameters is understood

as summation). The generalized mass operator M is defined as a sum

of 1 PI diagrams. The Green functions and mass operator can be

regarded as kernels of integral operators, hence the Dyson equation

can be represented in operator form

GΛ2 = GΛ

2 +GΛ2M

ΛGΛ2

or

(GΛ2 )−1 = (GΛ

2 )−1 +MΛ. (13.50)

It is useful to write this equation in (k, ε)-representation. (Here,

ε stands for the energy variable.) Due to translational invariance

in this representation, the operators entering Dyson equation are

operators of multiplications by a matrix function of (k, ε).We identify

the operators with these matrix functions. The (quasi-)particles are

related to the poles of the matrix function GΛ2 (k, ε). Recall that

we assume that the Hamiltonian H is represented as a sum of

Page 414: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 367

Methods of Quantum Field Theory in Statistical Physics 367

quadratic Hamiltonian H0 =∫ω(k)a+(k)a(k)dk and interaction

Hamiltonian Hint = gV. The poles of the GGreen function for

g = 0 (of the propagator) are located at the points ±ω(k) (we set

~ = 1); the dependence of these poles on g can be found in the

framework of the perturbation theory; the location of these poles will

be denoted ±ω(k | g). (Note that we cannot apply the perturbation

theory directly, but we can use it to find zeros of the RHS of (13.50).

The function ω(k | g) can be regarded as energy of (quasi-)particle.

Only in the ground state, one can hope that this function is real

(thermal quasiparticles are in general unstable).

13.4.5 Adiabatic S-matrix

The Dyson equation can also be written for adiabatic GGreen

functions; they can be used to describe the asymptotic behavior

of these functions for a → 0. As we have noted, the adiabatic

GGreen functions tend to GGreen functions as a → 0, but they do

not converge uniformly. However, the adiabatic self-energy operator

converges uniformly; this allows us to analyze the asymptotic behav-

ior of GGreen functions. (The same is true for conventional Green

functions). The reason for the uniform convergence is the fact that

matrix function M in (k, t)-representation tends to zero as t → ∞.Conventional adiabatic Green functions were analyzed by Likhachev

et al. (1970), the same method was applied by Tyupkin (1973) to

obtain the approximation for adiabatic GGreen functions. These

results were used to obtain the renormalized scattering matrix from

adiabatic scattering matrix. Note that all these considerations are

based on the assumption that the functions ω(k | g) are real, therefore

rigorously they can be applied only to the scattering of particles

(of elementary excitations of the ground state). Nevertheless, they

make sense as approximate formulas if the quasiparticles are almost

stable (we should assume that the collision time is much less than

the lifetime of quasiparticles and choose a in such a way that 1a

the lifetime of quasiparticles, but 1a the collision time).

The following statements were derived by Tyupkin (1973) from

the results of Schwarz (1967) and Likhachev et al. (1970) in the

framework of perturbation theory.

Page 415: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 368

368 Mathematical Foundations of Quantum Field Theory

The scattering matrix in the formalism of L-functionals can be

defined as an operator on the space of smooth L-functionals by the

formula

S = lima→0

VaSaVa, (13.51)

where Sa = Sa(∞,−∞) stands for the adiabatic S-matrix,

Va = exp i

∫ra(k)(c+

1 (k)c1(k)− c+2 (k)c2(k))dk

and the function ra is chosen from the requirement that S acts

trivially on one-particle states. (One can give an explicit expression

for ra in terms of one-particle energies ω(k | g); namely ra(k) =1a

∫ 0−∞(ω(k |h(τ))− ω(k))dτ.)

One can prove the existence of the limit in (13.51) in the

framework of perturbation theory. (One should impose the condition

ω(k1 + k2) < ω(k1) + ω(k2). This condition means that one-particle

spectrum does not overlap with multi-particle spectrum.)

The conventional renormalized S-matrix was related by Likhachev

et al. (1970) to the adiabatic S-matrices. (To obtain the S-matrix,

one should multiply the adiabatic S-matrix in finite volume by factors

similar to Va, take the limit when the volume tends to infinity, and

then take the limit a → 0.) Using the methods of Likhachev et al.

(1970), one can relate S to the conventional renormalized S-matrix S;

we obtain SLK = LSKS−1 . Using this formula, one can express

inclusive cross-section in terms of S (see Section 13.4.6).

13.4.6 Scattering of (quasi-)particles; inclusive

cross-section

This subsection is independent of the rest of this section (we use only

the definition of L-functional).

Let us start with the situation of quantum field theory when the

standard scattering matrix S is well defined as an operator acting

in the Fock space of asymptotic states. (Strictly speaking, we should

denote the operators acting in this space as ain(k), a+in(k), but we use

shorter notations a(k), a+(k).) We assume that the scattering matrix

Page 416: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 369

Methods of Quantum Field Theory in Statistical Physics 369

as well as Møller matrices S−, S+ are unitary. Considering the Fock

space as a representation of CCR, we assign an L-functional LK to

a density matrix K in the Fock space using the formula (13.47).

We define the scattering matrix in the space of L-functionals by the

formula

SLK = LSKS∗ . (13.52)

If the density matrix K corresponds to a vector Ψ, we can represent

the RHS of (13.52) as

Tr e−αa+eα∗aSKS∗

= Tr eα∗aSKS∗e−αa

+= 〈eα∗aSΨ, e−α

∗aSΨ〉

=∑n

∫dp1 · · · dpn〈eα

∗aSΨ|p1, . . . , pn〉〈e−α∗aSΨ|p1, . . . , pn〉,

where |p1, . . . , pn〉 =√

1n!a

+(p1) · · · a+(pn)θ constitute an orthonor-

mal basis in Fock space. (Here, θ = | 0〉 stands for Fock vac-

uum.) The expression 〈a(k1) · · · a(km)SΨ|p1, . . . , pn〉 =

√(m+n)!n!

〈SΨ|k1, . . . , km, p1, . . . , pn〉 can be interpreted as the scattering

amplitude of the process Ψ→ (k1, . . . , km, p1, . . . , pn) (after dividing

by numerical factor). We see that the LHS is expressed in terms of

scattering amplitudes. In particular, writing it in the form∑m,m′

∫dk1 · · · dkmdk′1 · · · dk′m′

(−1)m

m!m′!α(k1)

· · ·α(km)α∗(k′1)α′m′(k′m′)σm,m′(k1, . . . , km|k′1, . . . , k′m′ |Ψ),

we obtain

σm,m′(k1, . . . , km|k′1, . . . , k′m′ |Ψ)

=∑n

∫dp1 · · · dpn

√(m+ n)!

√(m′ + n)!

n!

×〈SΨ|k1, . . . , km, p1, . . . , pn〉〈SΨ|k′1, . . . , k′m′ , p1, . . . , pn〉.

If m = m′ and ki = k′i, this expression is proportional to the inclu-

sive cross-section Ψ → k1, . . . , km. If the initial state Ψ has definite

Page 417: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 370

370 Mathematical Foundations of Quantum Field Theory

momentum q, then 〈SΨ | k1, . . . , km, p1, . . . , pn〉 is a product of delta-

function δ(∑ki +

∑pj − q) coming from momentum conservation

and a function that will be denoted ρm(k1, p1, . . . , pn |Ψ). Now,

σm,m′(k1, . . . , km|k′1, . . . , k′m′ |Ψ)

= σm,m′(k1, . . . , km|k′1, . . . , k′m′ |Ψ)

× δ(k′1 + · · ·+ k′m − (k1 + · · ·+ km)),

where

σm,m′(k1, . . . , km|k′1, . . . , k′m′ |Ψ)

=∑n

∫dp1 · · · dpn

√(m+ n)!

√(m′ + n)!

n!

× ρm(k1, . . . , km, p1, . . . , pn |Ψ)ρm′(k′1, . . . , k

′n, p1, . . . , pn |Ψ)

× δ(k1 + · · ·+ km + p1 + · · ·+ pn − q).

The inclusive cross-section is proportional to σm,m(k1, . . . , km|k1,

. . . , km|Ψ) in this case.

We will call S inclusive S-matrix. Let us show that this matrix

can be calculated in terms of GGreen functions (more precisely, in

terms of on-shell values of these functions). Let us take a density

matrix K in the representation space of CCR. We assume that

the momentum and energy operators (infinitesimal spatial and time

translations) act on this space and K is translation invariant. This

allows us to define Heisenberg operators a+(k, t), a(k, t) where k

is the momentum variable and t is the time variable. Then the

corresponding GGreen function can be defined as TrBAK. Here, A

denotes chronological product (T -product) of Heisenberg operators

(the times are decreasing) and B stands for antichronological product

T opp of Heisenberg operators (the times are increasing).

If the density matrix K corresponds a vector Φ, the GGreen

function can be represented in the form

〈AΦ, B∗Φ〉,

where B∗ stands for chronological product of Hermitian conjugate

operators.

Page 418: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 371

Methods of Quantum Field Theory in Statistical Physics 371

Let us now consider the case when Φ is the ground state. Then one

can obtain an expression of S in terms of GGreen functions that is

analogous to the LSZ formula. It will be derived from some identities

that were used in the proof of LSZ, namely, we can use the identity

[· · · [S, ain(k1, σ1)] · · · ain(kn, σn)]

= (−1)nS−S∗+

∫dt1 · · · dtnLn · · ·L1T (a(k1, t1, σ1) · · · a(kn, tn, σn)),

(13.53)

where S−, S+ are Møller matrices, the scattering matrix S is

represented in terms of in-operators ain(k, 1) = a+in(k), ain(k,−1) =

ain(k), operators a(k, t, σ) are Heisenberg operators a+(k, t), a(k, t),

the operators∫dtiLi in (k, ε)-representation can be interpreted

as “on-shell operators.” (To apply the operator∫dtiLi in (k, ε)-

representation, we should multiply by iΛ(ki, σi)(εi−ω(ki)) and take

the limit εi → ω(ki). Here, ω(k) stands for the location of the pole

of the two-point Green function and Λ can be expressed in terms

of the residue in this pole. Note that in the transition to (k, ε)-

representation, we are using direct Fourier transform for σ = −1

and inverse Fourier transform for σ = 1.)

The identity (13.53) can be obtained, for example, from (32.17)

of Section 9.2 (by means of conjugation with S−).

It follows from (13.53) that Green function defined as vac-

uum expectation value of chronological product T (a(k1, t1, σ1) · · ·a(kn, tn, σn)) is related to scattering amplitude: one should take the

Fourier transform with respect to time variables and “go on-shell”

in the sense explained above. This gives the LSZ formula (see the

preceding section for more general approach). We remark that these

considerations also go through in the case when instead of vacuum

expectation value 〈0|A|0〉 where |0〉 obeys ain(k)|0〉 = 0 (represents

physical vacuum), we can take matrix elements 〈0|A|p1, . . . , pn〉where |p1, . . . , pn〉 = 1√

n!a+

in(p1) · · · a+in(pn)|0〉. This remark allows us

to express “on-shell” GGreen functions as sesquilinear combinations

of scattering amplitudes; comparing this expression with the formula

for SLK , we obtain the expression of S in terms of GGreen functions

“on-shell” (an analog of LSZ formula). Indeed, we can consider

Page 419: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 372

372 Mathematical Foundations of Quantum Field Theory

GGreen function as vacuum expectation value of chronological

product A multiplied by antichronological product B. Using the fact

that |p1, . . . , pn〉 constitute a generalized orthonormal basis, we can

say that

〈0|BA|0〉 =∑n

∫dp1 · · · dpn〈0|B|p1, . . . , pn〉〈p1, . . . , pn|A|0〉.

(13.54)

This representation allows us to express on-shell GGreen functions

in terms of scattering amplitudes. (The antichronological product is

related to the chronological one by Hermitian conjugation.)

Let us write explicit expressions obtained this way. We represent

S in normal form

S =∑r,s

1

r!s!

∫dp1 · · · dprdq1 · · · qsσr,s(p1, . . . , pr | q1, . . . , qs)a

+in(p1)

· · · a+in(pr)ain(q1) · · · ain(qs).

We assume that all pi’s are distinct and all qj ’s are distinct, then the

coefficient functions in normal form coincide with scattering ampli-

tudes 〈Sa+in(q1) · · · a+

in(qs)θ | a+in(p1) · · · a+

in(pr)θ〉. Let us take σi = 1

for i ≤ m, σi = −1 for i > m in (13.53). Introducing the notation

qi = ki−m, we obtain that

〈p1, . . . , pn|LHS|0〉 =1√n!σm+n,l(p1, . . . , pn, k1, . . . , km | q1, . . . , ql).

Here, LHS stands for the LHS of (13.53). Now, we can apply (13.53)

and (13.54) to identify on-shell GGreen functions with matrix entries

of the inclusive scattering matrix S.

Our considerations used LSZ relations that are based on the

conjecture that the theory has particle interpretation. Moreover, the

very definition of S that we have applied requires the existence of

conventional S-matrix. However, using (13.51) as the definition of

the inclusive scattering matrix S, one can prove the relation between

S and on-shell GGreen functions analyzing diagram techniques for

these objects. This proof can also be applied in the case when the

theory does not have particle interpretation. Moreover, the same

Page 420: Mathematical Foundations of Quantum Field Theory

March 30, 2020 11:40 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-ch13 page 373

Methods of Quantum Field Theory in Statistical Physics 373

ideas can be applied to quasiparticles considered as elementary exci-

tations of translation-invariant stationary state. These excitations

are related to the poles of the two-point GGreen function; we can

define the inclusive scattering matrix as on-shell GGreen function or

generalizing (13.51). Of course, this definition makes sense only if the

quasiparticles are (almost) stable.

In the preceding section, we have analyzed the inclusive scattering

matrix in the framework of algebraic approach to quantum theory

(see also Schwarz (2019c)).

Page 421: Mathematical Foundations of Quantum Field Theory

b2530 International Strategic Relations and China’s National Security: World at the Crossroads

b2530_FM.indd 6 01-Sep-16 11:03:06 AM

This page intentionally left blankThis page intentionally left blankThis page intentionally left blankThis page intentionally left blank

Page 422: Mathematical Foundations of Quantum Field Theory

March 30, 2020 10:29 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99a-app page 375

Appendix

A.1 Hilbert spaces

Let us consider a setH equipped with two operations: addition of two

elements and multiplication of elements by a complex number. We

say that H is a linear space if the following conditions are satisfied:

(1) x+y = y+x; (2) (x+y)+z = x+(y+z); (3) λ(x+y) = λx+λy;

(4) λ(µx) = (λµ)x, (λ+µ)x = λx+µx; (5) 0 ·x = 0, 1 ·x = x (where

x, y, z ∈ H; λ, µ are complex numbers). The elements of the linear

space are called vectors. We say the space is spanned by a subset S if

every element of the space can be represented as a linear combination

of the elements of the subset S. The dimension of a linear space is

the minimal number of vectors spanning the space. If this number

is infinite, then we say that the space is infinite dimensional. The

symbol 0 denotes the vector satisfying the conditions x + 0 = x for

all x in H.

A function on H taking values in the set of complex numbers is

called a linear functional if f(λx+ µy) = λf(x) + µf(y).

We say that a linear space is pre-Hilbert if it is equipped with

the inner product 〈x, y〉, satisfying the following axioms: (1) 〈x, y〉 =

〈y, x〉; (2) 〈λx, y〉 = λ〈x, y〉; (3) 〈x+y, z〉 = 〈x, z〉+〈y, z〉; (4) 〈x, x〉 ≥0 and 〈x, x〉 = 0 only if x = 0 (here, x, y, z ∈ H and λ is a complex

number).

We say that a bijective map is isomorphic if it preserves all

operations defined in a pre-Hilbert space.

The number ‖x‖ =√〈x, x〉 is called the norm of vector x. The

vector x is normalized if ‖x‖ = 1. Two vectors are orthogonal if

〈x, y〉 = 0.375

Page 423: Mathematical Foundations of Quantum Field Theory

March 30, 2020 10:29 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99a-app page 376

376 Mathematical Foundations of Quantum Field Theory

One says that the sequence xn ∈ H convergences to the element

x ∈ H (denoted by x = limxn), if limn→∞ ‖x−xn‖ = 0. We say that

the sequence xn ∈ H weakly converges to the vector x ∈ H (denoted

x = wlimxn) if for all elements y ∈ H the sequence 〈xn, y〉 tends

to 〈x, y〉.A sequence is called a Cauchy sequence if limm,n→∞ ‖xm−xn‖= 0.

The set M ⊂ H is closed in H if every point x ∈ H that can be

represented as a limit of a sequence xn ∈ M is itself an element of

M . A set M is dense in H if every vector x ∈ H can be represented

as a limit of elements of M .

A pre-Hilbert space H is called Hilbert space if every Cauchy

sequence converges. Every pre-Hilbert space can be embedded in a

Hilbert space. In other words, for every pre-Hilbert space H, one

can construct a Hilbert space H, called the completion of H, and an

isomorphic map α of the space H onto a dense subset of the space

H (isomorphic embedding of H into H).

The completion H is unique in the following sense: if H1 and H2

are two completions and α1 and α2 are isomorphic embeddings of Hinto H1 and H2, then there exists an isomorphic map α of H1 onto

H2 satisfying αα1 = α2.

A pre-Hilbert space is called separable if it contains a countable

dense subset. We will consider only separable spaces.

A.2 Systems of vectors in a pre-Hilbert vector space

A subset M ⊂ H is called a linear subspace if every linear

combination λx+ µy with x, y ∈M is also in M .

A system of vectors ξa ∈ H is called orthonormal if 〈ξα, ξβ〉 = δαβ(we consider systems that are either finite or countable). A orthonor-

mal system is called an orthonormal basis if linear combinations of

elements of this system are dense in H.The quantities 〈x, ξa〉 are called the Fourier coefficients of x in the

orthonormal basis ξa. For every vector x ∈ H, the series∑

a〈x, ξa〉ξa(called the Fourier series of x in the orthonormal basis ξa) converges

to the vector x. The scalar product 〈x, y〉 can be expressed in terms

Page 424: Mathematical Foundations of Quantum Field Theory

March 30, 2020 10:29 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99a-app page 377

Appendix 377

of the Fourier coefficients as follows:

〈x, y〉 =∑α

〈x, ξα〉 〈y, ξα〉.

In particular, 〈x, x〉 =∑

α | 〈x, ξα〉 |2.

If H is a Hilbert space, then the sequence cα is a sequence of

Fourier coefficients of vector x if and only if∑

α |cα|2 <∞.

A.3 Examples of function spaces

1. The space l2n, also denoted as Cn, is the n-dimensional space of

rows of n complex numbers equipped with the inner product

〈x, y〉 =

n∑i=1

xiyi

(here, x = (x1, . . . , xn) ∈ Cn, y = (y1, . . . , yn) ∈ Cn).

2. The space l2 consists of sequences of complex numbers x =

(x1, . . . , xi, . . . ), satisfying the condition∑∞

i=1 |xi|2 < ∞. Addi-

tion and multiplication by a complex number are defined

coordinate-wise and the scalar product is defined as 〈x, y〉 =∑∞i=1 xiyi.

The spaces l2n and l2 are Hilbert spaces.

If H is a Hilbert space and ξα is an orthonormal basis for H,

then we can construct an isomorphic map to the space l2n (if His finite dimensional) and l2 (if H is infinite dimensional). The

isomorphism transforms a vector x to its sequence of Fourier

coefficients in the basis ξα, obtained as 〈x, ξα〉.3. The space C2(Er) consists of functions on an r-dimensional space

Er that are continuous and square integrable (i.e. the functions

satisfy the condition∫|f(ξ1, . . . , ξr)|2dξ1, . . . , dξr <∞). Multipli-

cation and addition of functions are defined in the standard way.

The scalar product of two functions f and g in C2(Er) is given by

〈f, g〉 =

∫f(ξ1, . . . , ξr)g(ξ1, . . . , ξr)dξ1, . . . , ξr

=

∫f(ξ)g(ξ)dξ.

Page 425: Mathematical Foundations of Quantum Field Theory

March 30, 2020 10:29 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99a-app page 378

378 Mathematical Foundations of Quantum Field Theory

The space C2(Er) is a pre-Hilbert space. Its completion is L2(Er).

4. The space C2(Er × S), where S denotes a finite set, consists

of functions f(ξ, s) that are continuous with respect to the

variable ξ and are square integrable (i.e. satisfying the condition∑s

∫|f(ξ, s)|2dξ < ∞). Here, ξ = (ξ1, . . . , ξr) ∈ Er; s ∈ S, dξ =

dξ1 . . . dξr. The scalar product is defined by

〈f, g〉 =∑s

∫f(ξ, s)g(ξ, s)dξ.

The function f(ξ, s) can be considered as a column of k functions

fs(ξ) = f(ξ, s) depending on the variable ξ ∈ Er (k is the number

of elements in S).

The completion of C2(Er × S) is denoted by the symbol

L2(Er × S).

Let us fix a family B of subsets of M that contains with every

two subsets A,B the subsets A ∪ B,A ∩ B, and A \ B (we call this

family a ring of subsets). Let us also assume that the set M is a

countable union of subsets belonging to the family B. A countably

additive measure µ on a family B is specified by an assignment of

non-negative numbers µ(A) to every set A ∈ B.

We require that for any countable sequence of sets A,A1, . . . ,

An, . . . from B we have µ(A) =∑∞

i=1 µ(Ai) if

(1) A = ∪∞i=1Ai,

(2) the sets A1, . . . , An, . . . are pairwise disjoint. (In other words, we

assume that the measure is countably additive.)

The setM , equipped with the family B and the countably additive

measure µ, specifies a measure space.

A set R ⊂ M (not necessarily belonging to B) is called a set of

measure zero if for every ε > 0, there exists a set A ∈ B, such that

R ⊂ A and µ(A) < ε. One says that a sequence of functions fn(x)

converges to a function f(x) almost everywhere on a set M if the

set of points x ∈M , such that the sequence fn(x) does not converge

to f(x), is a set of measure zero. In general, if a relation is satisfied

everywhere except a measure zero set, then we say that it is satisfied

almost everywhere.

Page 426: Mathematical Foundations of Quantum Field Theory

March 30, 2020 10:29 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99a-app page 379

Appendix 379

For functions defined on a measure space, one can define a notion

of Lebesgue integral. A function f defined on a set A ∈ B is

called a simple function, if the set of values of the function is finite

y1, . . . , yn and the sets Ai consisting of points where the function

f takes the value yi belong to the family B. The Lebesgue integral

of a simple function f is defined by the formula∫Af(x)dµ =

r∑i=1

yiµ(Ai).

A function f(x) that can be represented as a limit of a sequence of

simple functions converging almost everywhere is called measurable.

In what follows, we consider only measurable functions. Let us say

that two measurable functions are equivalent if their difference is

equal to zero almost everywhere. We do not distinguish equivalent

measurable functions.

If a measurable function f is bounded, then it can be represented

as a limit of a sequence of simple functions fn(x) that are bounded

from above by a constant, |fn(x)| ≤ C. The Lebesgue integral of the

function f on a set A ∈ B can be defined by∫Af(x)dµ = lim

n→∞

∫Afn(x)dµ.

One can prove that this limit always exists and does not depend on

the choice of the sequence fn(x).

If the function f is unbounded and non-negative, then one can

define its Lebesgue integral by the relation∫Af(x)dµ = lim

n→∞

∫Afn(x)dµ,

where fn(x) = f(x) for f(x) ≤ n and fn(x) = 0 for f(x) ≥ n.

If the set A does not belong to the family B, but can be

represented as A =⋃∞i=1Ai, where Ai ∈ B, then the Lebesgue

integral of a non-negative function f over A is defined by∫Af(x)dµ = lim

n→∞

∫A1∪···∪An

f(x)dµ.

Page 427: Mathematical Foundations of Quantum Field Theory

March 30, 2020 10:29 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99a-app page 380

380 Mathematical Foundations of Quantum Field Theory

(For any non-negative measurable function, the Lebesgue integral

exists, but is not necessarily finite.)

A measurable function is called Lebesgue summable on the set

A ⊂ M , if the Lebesgue integral of the function |f(x)| is finite. For

a summable real function, the Lebesgue integral is defined by the

relation ∫Af(x)dµ =

∫A|f(x)|dµ−

∫A

(|f(x)| − f(x))dµ.

If the function f(x) is complex, then the Lebesgue integral is defined

by∫A Re(f(x))dµ + i

∫A Im(f(x))dµ, where Re(f(x)), Im(f(x)) are

real and imaginary parts of f(x).

We note some important properties of the Lebesgue integral:

1. Let fn(x) be a sequence of functions converging almost everywhere

to the function f(x) and suppose there exists a function g(x)

dominating the functions fn(x) (i.e. |fn(x)| ≤ g(x)). Then

limn→∞∫A fn(x)dµ =

∫A f(x)dµ.

This theorem allows us to interchange the limit and Lebesgue

integration.

2. Let M1 and M2 be measure spaces. Then in the space M1 ×M2

consisting of the pairs (x1, x2), where x1 ∈M1 and x2 ∈M2, one

can construct a countably additive measure µ defined by µ(A1 ×A2) = µ1(A1)µ2(A2) (here, µi denotes the measure on Mi; the

measure µ is called the product measure of µ1 and µ2).

If the function f(x1, x2) is a function on the product M1 ×M2

that is summable with respect to the measure µ, then for almost

all x2 the function f(x1, x2) is a summable function of x1 ∈ M1.

The double integral of the function f (i.e. the integral over the

measure µ) can be expressed as the repeated integral∫A1×A2

f(x1, x2)dµ =

∫A2

dµ2

∫A1

f(x, x2)dµ1

(Fubini’s theorem).

Using the Lebesgue integral, one can introduce a notion of measure

for some subsets that do not belong to the family B. Namely, for

Page 428: Mathematical Foundations of Quantum Field Theory

March 30, 2020 10:29 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99a-app page 381

Appendix 381

every subset A ⊂ M , one can construct a function χA(x) that is

equal to 1 for x ∈ A and 0 for x 6∈ A and set µL(A) =∫M χA(x)dµ.

This measure µL is called the Lebesgue measure. It is defined for

the set A if the function χA is measurable (then the set A is called

measurable).

For every measure set M , we can construct a space L2(M) of

square-integrable functions1 (i.e. measurable functions such that the

Lebesgue integral of the function |f2(x)| over the set M is finite).

The scalar product on the space L2(M) can be defined by means of

the Lebesgue integral2

〈f, g〉 =

∫f(x)g(x)dµ.

One can prove that the space L2(M) is a Hilbert space.

For the Euclidean space Er, we take the standard volume as a

measure. Then, for functions that are integrable in the standard

sense, the conventional integral (Riemann integral) coincides with

the Lebesgue integral. The space L2(Er) is the completion of the

space C2(Er).

If S is a finite set, then a measure of its subset A is by definition

the number of points in A. The space L2(S) is isomorphic to the

space l2k where k is the number of points in S. In the set Er × S,

we define the measure as the product of measure in Er and S. The

space L2(Er×S) constructed with this measure is the completion of

the space C2(Er × S).

A.4 Operations with Hilbert spaces

The direct sum H1 + H2 of Hilbert spaces H1 and H2 is a Hilbert

space, whose elements consist of pairs (h1, h2) where h1 ∈ H1, h2 ∈H2, with addition, multiplication by a complex number, and the

1Two equivalent measurable functions specify the same element of the space ofL2(M).2Integrating over the whole measure space M , we use the notation

∫φ(x)dµ

instead of the notation∫Mφ(x)dµ.

Page 429: Mathematical Foundations of Quantum Field Theory

March 30, 2020 10:29 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99a-app page 382

382 Mathematical Foundations of Quantum Field Theory

scalar product defined by

(h1, h2) + (h′1, h′2) = (h1 + h′1, h2 + h′2),

λ(h1, h2) = (λh1, λh2),⟨(h1, h2), (h′1, h

′2)⟩

=⟨h1, h

′1

⟩+⟨h2, h

′2

⟩.

Let us assume that subspaces H1 and H2 of the Hilbert space Hsatisfy the following conditions: (1) every two vectors h1 and h2 are

orthogonal and (2) every h ∈ H can be represented as a sum h =

h1 + h2 of two vectors h1 ∈ H1 and h2 ∈ H2. Then one says that

the Hilbert space H is represented as a direct sum of subspaces H1

and H2. In this case, one can construct a natural isomorphism of

the direct sum H1 + H2 and the space H, transforming the vector

(h1, h2) ∈ H1 +H2 into the vector h1 + h2 ∈ H.

The direct sum∑

iHi of a countable sequence of Hilbert spaces

H1, . . . ,Hi, . . . is defined as the space of sequences (h1, . . . , hi, . . . ),

satisfying the condition∑

i ‖hi‖2 < ∞ (here, hi ∈ Hi). Linear

combination and scalar product are defined by

λ(h1, . . . , hi, . . . ) + λ′(h′1, . . . , h′i, . . . )

= (λh1 + λ′h′1, . . . , λhi + λ′h′i, . . . ),⟨(h1, . . . , hi, . . . ), (h

′1, . . . , h

′i, . . . )

⟩=∑i

⟨hi, h

′i

⟩.

The Hilbert space H is said to be a tensor product of Hilbert spaces

H1 and H2, if there exists a bilinear map α(h1, h2) mapping H1×H2

to H and having the following properties:

(1) 〈α(h1, h2), α(h′1, h′2)〉 = 〈h1, h

′1〉 · 〈h2, h

′2〉;

(2) the set of linear combinations of vectors of the form α(h1, h2) is

dense in H. We say that a map α(h1, h2) from H1 ×H2 into His bilinear if it is linear with respect to h1 for a fixed h2 and is

linear with respect to h2 for fixed h1.

The tensor product is defined by the above properties up to a natural

isomorphism. (If H and H′ are two tensor products of H1 and H2

Page 430: Mathematical Foundations of Quantum Field Theory

March 30, 2020 10:29 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99a-app page 383

Appendix 383

and α1 and α2 are corresponding bilinear maps, then there exists an

isomorphism λ of Hilbert spaces H and H′, such that λα = α′.)

The tensor product of spaces H1 and H2 will be denoted by the

symbol H1 ⊗H2 and the vector α(h1, h2) by the symbol h1 ⊗ h2.

It is easy to check that the tensor product of the spaces l2m and

l2n is isomorphic to the space l2mn. The space l2mn can be realized as

a space of m × n matrices. Then the bilinear map α assigns to the

vectors (x1, . . . , xm) ∈ l2m and (y1, . . . , yn) ∈ l2n the matrix A with

the elements Aij = xiyj .

The tensor product of n copies of space H is called the nth tensor

power of the space and is denoted by ⊗Hn. The formal definition is

via taking repeated tensor products

⊗Hn = (. . . ((H⊗H)⊗H) · · · ⊗ H).

Every permutation π of the indices (1, . . . , n) naturally corresponds

to an isomorphic map ρπ of the space ⊗Hn on itself.

The subspace Hns of the space ⊗Hn consisting of vectors that

satisfy the condition ρπx = x for all permutations π is called an

nth symmetric power of H. The nth antisymmetric power of Hna of

the space H is defined as the subspace consisting of vectors obeying

ρπx = (−1)γ(π)x for all π (where γ(π) stands for the parity of the

permutation π).

Note that

L2(M1)⊗ L2(M2) = L2(M1 ×M2)

(the measure in the space M1 × M2 is defined as the product of

measures in M1 and M2). The bilinear map α in the definition of the

tensor product sends the functions f1 ∈ L2(M1), f2 ∈ L2(M2) to the

function f(x1, x2) = f1(x1)f2(x2) ∈ L2(M1 ×M2). The symmetric

(antisymmetric) power of the space L2(M) can be realized as the

space of symmetric (antisymmetric) square integrable functions of

the variables x1, . . . , xn ∈M .

A.5 Operators on Hilbert spaces

We consider linear operators acting on a Hilbert space H1 and taking

values in a Hilbert space H2 as maps from dense subspace D ⊂

Page 431: Mathematical Foundations of Quantum Field Theory

March 30, 2020 10:29 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99a-app page 384

384 Mathematical Foundations of Quantum Field Theory

H1 into H2, satisfying the condition of linearity (i.e. the operator

A sends every vector x ∈ D to the vector Ax ∈ H2 and satisfies

A(λ1x1 +λ2x2) = λ1Ax1 +λ2Ax2 where x1, x2 ∈ D). The domain of

definition for the operator A is denoted by DA. The set of vectors in

H2 that can be written as Ax for x ∈ H1 is called the range of A or

the image of A.

We say that the operator A is bounded if there exists a number

K, such that ‖Ax‖ ≤ K‖x‖ for all x ∈ DA. The operator norm of

A is defined as supx‖Ax‖‖x‖ . If the operator A is bounded, then it can

be continuously extended to the whole space H1 (i.e. there exists a

unique bounded operator on H1 that coincides with A on DA).

The multiplication operator by the number λ will be denoted by

the same symbol. In particular, the identity operator, considered as

a multiplication operator by 1, will be denoted by the symbol 1.

A linear combination C = λ1A1 +λ2A2 of operators A1 and A2 is

defined by Cx = λ1A1x+λ2A2x. Operator multiplication F = A1A2

of operators A1 and A2 is defined as composition Fx = A1(A2x).

The linear combination of operators A1 and A2 acting from H1

into H2 is defined if the intersection of their domains of definition

DA1 and DA2 is everywhere dense in H1. Similarly, the product of

operators A1 and A2 is defined if A2 maps H0 into H1, A1 maps H1

into H2 and the set of x ∈ DA2 for which A2x ∈ DA1 is dense in H0.

If the range of an operator A, transforming H1 into H2, is dense

in H2 and the operator is injective (only 0 ∈ H1 goes to 0 ∈ H2),

then we can define the inverse operator A−1, for which A−1y = x, if

y = Ax.

If the operator A transforms H1 into H2 and the operator B

transforms H2 into H1, then the two are called conjugate if 〈Ax, y〉 =

〈x,By〉 for all x ∈ DA, y ∈ DB.

If two operators B1 and B2 are both conjugate with the operator

A and both are defined on the vector y ∈ H2, then B1y = B2y (this

follows from the density of DA in H1). We will denote by A∗ the

conjugate operator to A that is defined for all y ∈ H2 for which there

exists z ∈ H1 that satisfies 〈Ax, y〉 = 〈x, z〉 (in other words, A∗ has

the maximum domain of definition of all operators conjugate to A).

Page 432: Mathematical Foundations of Quantum Field Theory

March 30, 2020 10:29 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99a-app page 385

Appendix 385

Note that we demand that an operator is defined on a dense set,

therefore, the operator A∗ does not always exist.

If the operator A is bounded, then the operator A∗ always exists

and is bounded, with ‖A∗‖ = ‖A‖.The operator U transforming H1 into H2 is called an isometry

if it preserves inner products: 〈Ux,Uy〉 = 〈x, y〉. For an operator

to be an isometry, it is necessary and sufficient for it to satisfy

U∗U = 1. An isometry is called a unitary operator if it is surjective

(the range coincides with H2). The condition for unitarity can be

written in the form UU∗ = U∗U = 1. A unitary operator defines

an isomorphism between H1 and H2, while an isometry defines an

isomorphism between H1 and a subspace of H2.

If H1 = H2 = H, then operators on H and their associated

conjugate operators are defined on the same space H. In this case,

we define the following notions. An operator A is called Hermitian if

it is conjugate to itself (i.e. 〈Ax, y〉 = 〈x,Ay〉 for all x, y ∈ DA). An

operator A is called self-adjoint if A∗ = A.

If A is Hermitian, B is self-adjoint, and DA ⊂ DB and Ax = Bx

for all x ∈ DA, then the operator B is called a self-adjoint extension

of A. If a Hermitian operator has only one self-adjoint extension,

then it is called essentially self-adjoint.

Self-adjoint operators are important in quantum mechanics

because they correspond to physical quantities. Usually, in physics

books, only Hermitian operators are discussed; the notion of the self-

adjoint operator is identified with the notion of the Hermitian opera-

tor. This does not cause many issues, since most Hermitian operators

encountered in physics turn out to be essentially self-adjoint. Note,

however, that if a Hermitian operator is not essentially self-adjoint,

then a choice of self-adjoint extension in physics problems leads to

different results (quite often this choice corresponds to a choice of

boundary conditions.)

An operator A is called bounded from below if there exists a

constant K such that 〈Ax, x〉 ≥ K 〈x, x〉 for any vector x ∈ DA. If

K = 0, then we call A positive semidefinite, and if for x 6= 0 the

expression 〈Ax, x〉 is positive, then we call A positive definite.

Page 433: Mathematical Foundations of Quantum Field Theory

March 30, 2020 10:29 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99a-app page 386

386 Mathematical Foundations of Quantum Field Theory

Hermitian operators that are bounded from below always have

self-adjoint extensions. One of these extensions is called the

Friedrichs extension Aµ; it can be defined for the operator A by

the following extremal property: if B is a self-adjoint extension of

A and λ is a number such that B + λ is positive definite, then Aµsatisfies

⟨(Aµ + λ)−1x, x

⟩≤⟨(B + λ)−1x, x

⟩. Friedrichs extension

Aµ is bounded from below if A is bounded from below (in fact, if

〈Ax, x〉 ≥ K 〈x, x〉 then 〈Aµx, x〉 ≥ K 〈x, x〉).The concrete operators we consider in this book are often given

as formal expressions that combine simpler operators (such as

multiplication and differentiation in the space L2(En) or annihilation

and creation operators on Fock space). For full rigor, one should

specify the domain of all operators considered. Instead, however, for

simplicity, we agree that operators in the space L2(En) are defined

on the space S(En) of rapidly decaying smooth functions (more

detailed definition of S(En) is given in A.6) and operators in the

Fock space F (L2(En)) are defined on the set of Fock states specified

by a sequence of functions fk(x1, . . . , xk) ∈ S(Enk) (we assume that

only a finite number of these functions are not equal to 0).

Saying that a formal operator expression specifies a self-adjoint

operator, we have in mind that this expression specifies a Hermitian

operator that is either essentially self-adjoint or is bounded from

below. In the latter case, the self-adjoint operator defined by the

expression at hand is the Friedrichs extension of the corresponding

Hermitian operator.

When the formal operator expression does not define a Hermitian

operator or specifies a Hermitian operator that does not have a self-

adjoint extension, then we say that this formal expression does not

specify a self-adjoint operator.

We say that the operator P is a projection operator if P = P ∗

and P 2 = P . Projections on a Hilbert space are in one-to-one

correspondence with the subspaces of the Hilbert space (every

projection corresponds to its range).

Let us consider the example of the operator a(x) of multiplication

by a measurable function a(x) in the space L2(M) where M is a

measure space (the operator a(x) transforms f(x) into the function

Page 434: Mathematical Foundations of Quantum Field Theory

March 30, 2020 10:29 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99a-app page 387

Appendix 387

a(x)f(x) and is defined on the functions f(x) ∈ L2(M) such that

the function a(x)f(x) is square integrable). It is easy to check that

(a(x))∗ =ˆ

a(x). It follows that the operator a(x) is (1) self-adjoint

if the function a(x) is real, (2) unitary if |a(x)| = 1, and (3) is a

projection if the function a(x) takes only the values 0 and 1.

Note that the operator a(x) can be considered on a smaller

domain; the adjoint operator does not depend on the choice of the

domain. It follows from this remark that for real a(x) the operator

a(x) is essentially self-adjoint on every domain.

The operator of multiplication is universal in some sense. Namely,

one can prove the following important theorem.

For every self-adjoint or unitary operator A in the space H, one

can find a measure space M and an isomorphism α of the space

H onto the space L2(M) that transforms the operator A into a

multiplication operator by a function a(x) (i.e. αAα−1 = α(x)). The

function a(x) is real if A is self-adjoint and |a(x)| = 1 if A is a unitary

operator.

The theorem allows us to define a function of a self-adjoint or a

unitary operator. Namely, if φ(λ) is a function of a real variable, then

we can define the operator φ(A), where A is a self-adjoint operator,

as the operator of multiplication by φ(a(x)) using the isomorphism α

described above (more precisely, if αAα−1 = a(x), then αφ(A)α−1 =

φ(a(x))). (We assume that the function φ(a(x)) is measurable; this

is always the case if the function φ(λ) is Borel measurable, i.e. it

is measurable with respect to the family B generated by open sets,

their countable intersections, and unions.)

Note the following obvious relations:

1. If h(λ) = γf(λ) + µg(λ); r(λ) = f(λ)g(λ); s(λ) = f(g(λ)), then

h(A) = γf(A) + µg(A); r(A) = f(A)g(A); s(A) = f(g(A)).

2. If B = U−1AU , where U is a unitary operator on the space H,

then φ(B) = U−1φ(A)U .

3. If φ(λ) is a bounded function and |φ(λ)| ≤ M , then φ(A) is a

bounded operator with ‖φ(A)‖ ≤M .

4. [φ(A)]∗ = φ(A); if the function φ(λ) is real, then the operator φ(A)

is self-adjoint, and if |φ(λ)| = 1, then φ(A) is a unitary operator.

Page 435: Mathematical Foundations of Quantum Field Theory

March 30, 2020 10:29 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99a-app page 388

388 Mathematical Foundations of Quantum Field Theory

In particular, if A is self-adjoint, then the operator exp(itA) is

unitary.

If φ is a measurable function on the circle |z| = 1, and the operator

A is unitary, then the operator φ(A) can be defined precisely in the

same way and has similar properties.

If A is a self-adjoint operator, it is convenient to consider

projections Eµ = eµ(A), where eµ(λ) is a function equal to 1 for

λ ≤ µ and 0 for λ > µ. Projections Eµ are called spectral projections

of the operator A. It follows from the relation 1 =∫deµ that

A =

∫ ∞−∞

µdEµ

(this formula is called the spectral decomposition of the operator

A; for finite a and b, the integral∫ ba µdEµ is defined as the limit of

integral sums ∑µk(Eµk+1

− Eµk),

where µ0 = a < µ1 < · · · < µn = b; ∆µk = µk+1 − µk → 0.

The integral∫∞−∞ µdEµ is defined as the strong limit

∫ ba µdEµ where

a→∞, b→ −∞.

We say that a real number λ does not belong to the spectrum of

a self-adjoint operator A, if for some ε > 0 we have Eλ+ε = Eλ−ε.

It is easily checked that the spectrum of the operator a(x) of

multiplication by the function a(x) in the space L2(M) coincides

with the range of a(x). Recall that we do not distinguish equivalent

measurable functions, therefore, the range of measurable functions

should be defined in such a way that the set does not change when we

replace a function by an equivalent function. We say that the number

λ belongs to the range of the function a(x) if for every function b(x)

differing from a(x) only on a set of measure 0 and every ε > 0, there

exists an x ∈M such that λ− ε < b(x) < λ+ ε.

Using this statement and the previously defined isomorphism

transforming a self-adjoint operator into an operator of multiplica-

tion by a function, it is easy to prove the following statements:

Page 436: Mathematical Foundations of Quantum Field Theory

March 30, 2020 10:29 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99a-app page 389

Appendix 389

1. If the number β doesn’t belong to the spectrum of a self-adjoint

operator H, then ∫exp(−iβa) exp(iHa)da = 0.

More precisely, for any x, y ∈ H∫exp(−iβa) 〈exp(iHa)x, y〉 da = 0.

2. Let us suppose that for every smooth finite function χ(ω) with

support on (a, b) we have∫χ(t) exp(iHt)dt = 0,

where

χ(t) =

∫exp(−iωt)χ(ω)dω.

Then the spectrum of the operator H lies outside the interval (a, b).

Two operators A and B, defined on the whole space H, are called

commuting if AB = BA. The same definition can be used in the

case where operators A and B are defined on the same set D and

transform it into itself. However, for self-adjoint operators, one should

use another definition of commuting operators. One says that two

self-adjoint operators A and B in the Hilbert space H commute if

φ(A)ψ(B) = ψ(B)φ(A) for all bounded functions φ(λ), ψ(λ) (if this is

the case, we say that AB = BA). Note that the operators φ1(A) and

φ2(A) commute (here, φ1(λ) and φ2(λ) are arbitrary real functions

and A is a self-adjoint operator).

One can prove the following theorem.

For any family of self-adjoint commuting operatorsA1, . . . , An, . . .

in a Hilbert space H, one can find a measure space M and an isomor-

phism α between the spaceH and the space L2(M) transforming each

of these operators into the operator of multiplication by a function

(i.e. αAiα−1 = ai(x)).

As in the case of one operator, this theorem allows us to

define functions of families of operators. Namely, if A1, . . . , An are

Page 437: Mathematical Foundations of Quantum Field Theory

March 30, 2020 10:29 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99a-app page 390

390 Mathematical Foundations of Quantum Field Theory

commuting self-adjoint operators and f(λ1, . . . , λn) is a measurable

function of n real variables, then f(A1, . . . , An) is defined as the

operator corresponding to the operator of multiplication by the

function f(a1(x), . . . , an(x)) when we apply the isomorphism α.

If ∆ is an open set in n-dimensional space, we define e∆(x1,

. . . , xn) to be the function that is equal to 1 if (x1, . . . , xn) ∈ ∆

and 0 otherwise.

The projection operators E∆ = e∆(A1, . . . , An) are called the

spectral projections of the family of commuting self-adjoint operators

A1, . . . , An.

We say that the point (λ1, . . . , λn) of an n-dimensional space

belongs to the spectrum (more precisely, the joint spectrum) of the

family of commuting self-adjoint operators A1, . . . , An, if for every

neighborhood U of the point (λ1, . . . , λn), the spectral projection

EU = eU (A1, . . . , An) does not equal 0.

If the operators A1, . . . , An can be realized in the space L2(M)

as operators of multiplication by the functions a1(x), . . . , an(x), then

the joint spectrum of these operators consists of all points of the form

(a1(x), . . . , an(x)) (more precisely, the point (λ1, . . . , λn) belongs to

the spectrum if and only if for all functions bi(x) equivalent to the

functions ai(x) and any neighborhood U of the point (λ1, . . . , λn)

there exists a point x ∈M such that (b1(x), . . . , bn(x)) ∈ U).

Let us discuss the question of the convergence of a sequence of

operators. We define three types of convergence, assuming for brevity

that all operators are bounded.

A sequence of operators An converges to the operator A uniformly

(or in norm), if ‖A−An‖ → 0.

A sequence of operators An converges to the operator A strongly

if limAnx = Ax for any vector x.

A sequence of operators An converges to the operator A weakly

if lim 〈Anx, y〉 = 〈Ax, y〉 for any vectors x and y.

Convergence in norm will be denoted by the symbol A = nlimAnor An =⇒ A, strong convergence will be denoted by the symbol

A = slimAn or An → A, and weak convergence will be denoted by

A = wlimAn or An A.

Page 438: Mathematical Foundations of Quantum Field Theory

March 30, 2020 10:29 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99a-app page 391

Appendix 391

It is clear that convergence in norm implies strong convergence

and strong convergence implies weak convergence.

Note the following statements.

(a) if for a dense set of vectors x the limit slimAnx exists and the

sequence ‖An‖ is bounded, then the sequence of operators An is

strongly convergent;

(b) if the limit lim 〈Anx, y〉 exists for all x from a dense subset X

and for all y from a dense subset Y and the sequence ‖An‖ is

bounded, then the sequence An is weakly convergent;

(c) if An → A,Bn → B and the sequence ‖An‖ is bounded, then

AnBn → AB (the analog of this statement for weak convergence

does not hold); if An → A and the sequence ‖A−1n ‖ is bounded,

then A−1n → A−1;

(d) if An is a sequence of unitary operators and An =⇒ A, then

the operator A is unitary; if An → A, then the operator A is

isometric; if An A, then we can only say that ‖A‖ ≤ 1;

(e) if An A, then A∗n A∗ (for strong convergence, the analogous

statement does not hold; however, if An is a sequence of unitary

operators that strongly converges to the unitary operator A, then

A∗n = A−1n → A∗ = A−1).

Very often, it is useful to consider a family of operators transforming

a fixed linear subspace D that is dense in the Hilbert subspace

into itself (in other words, the operators are defined on D and

their range is contained within D); the family of operators having

this property will be denoted ND. The family ND is closed with

respect to linear combination and multiplication of operators. If for

an operator A ∈ ND, there exists an operator B ∈ ND satisfying the

condition 〈Ax, y〉 = 〈x,By〉 for all x ∈ D, y ∈ D, then we say that

the operator B is D-conjugate to the operator A and we denote it

by A+.

The strong limit of a sequence A = slimAn of operators An ∈ NDis defined as the operator satisfying the condition Ax = limAnx for

all x ∈ D; the weak limit A = wlimAn is defined by the condition

〈Ax, y〉 = lim 〈Anx, y〉 for any x, y ∈ D.

Page 439: Mathematical Foundations of Quantum Field Theory

March 30, 2020 10:29 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99a-app page 392

392 Mathematical Foundations of Quantum Field Theory

A.6 Locally convex linear spaces

A seminorm on a linear space R is a non-negative function p(x) that

satisfies the conditions p(λx) = |λ|p(x), p(x+y) ≤ p(x)+p(y) for all

x, y ∈ R and for any real number λ. A seminorm that vanishes only

for x = 0 is called a norm.

A locally convex topology on a linear space R is specified by a

system A of seminorms pα(x) that satisfies the following conditions:

(1) for every two seminorms pα, pβ ∈ A, one can find a seminorm pγsuch that pγ ≥ pα, pγ ≥ pβ; (2) for every point x ∈ R, one can find a

seminorm pα ∈ A that does not vanish at the point x. A linear space

equipped with a locally convex topology is called a locally convex

topological linear space or simply a locally convex space, and the

system A is called the defining system of seminorms. If the defining

system consists of a single seminorm (which is automatically a norm),

then the space is called a normed space.

The function f(x) on a locally convex space R is continuous in

the topology of this space if for every ε > 0 and every point x0 ∈ R,

there exists a seminorm pα in the defining system of seminorms for

R and a δ > 0 such that for any point x satisfying the condition

pα(x − x0) < δ, we have |f(x) − f(x0)| < ε. As an example of a

locally convex space, we can consider the space S(Er) of smooth

functions of r variables, having all derivatives tending to zero faster

than any polynomial.

The topology on the space S(Er) is specified by the system of

seminorms

‖φ‖α,β = supx∈Er

|x(α)D(β)φ(x)|

(here, α = (α1, . . . , αr), β = (β1, . . . , βr) are arbitrary sets of

non-negative integers, x(α) = xα11 . . . xαrr , D(β) = ∂β1+···+βr

∂xβ11 ...∂xβrr

).

We say that two systems of seminorms on a linear space are

equivalent (i.e. they specify the same topology) if every function that

is continuous in one system is necessarily continuous in the other. In

other words, one can consider different defining systems of seminorms

on the same locally convex space; among these systems, there exists

Page 440: Mathematical Foundations of Quantum Field Theory

March 30, 2020 10:29 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99a-app page 393

Appendix 393

a maximal system consisting of all continuous seminorms. Instead of

using the precise term “a seminorm that is continuous with respect

to the locally convex topology in R”, we will use the shorter term “a

seminorm in R”.

A sequence ξn of elements of the locally convex space R converges

to the element ξ if for every seminorm p in R the sequence p(ξn− ξ)tends to zero. A sequence ξn is called Cauchy if for every seminorm

p in R we have limm→∞,n→∞ p(ξm − ξn) = 0. In a complete locally

convex space, every Cauchy sequence is convergent.

A.7 Generalized functions (distributions)

Physicists define generalized functions f(x) as symbols that only

make sense under the sign of an integral∫f(x)φ(x)dx, where φ(x) is

a “good” function. The mathematical formalization of this definition

is given as follows.

Let M be a measure space and R any linear subspace of the

space of measurable functions on M (the elements of R are called

test functions). A generalized function on M (more precisely, a

generalized function with numerical values) is defined as a linear

functional f(φ) on the space R. A generalized function on the space

M is denoted by f(x) and we define∫f(x)φ(x)dx = f(φ). Note that∫

f(x)φ(x)dx should be understood as another way of writing the

number f(φ).

It is often convenient to consider the space of test functions R

endowed with a topology, and generalized functions as continuous

linear functionals on R. We will consider the topology on the space

of test functions only in one situation — when defining generalized

functions of moderate growth.

Clearly, the definition of generalized functions depends on the

choice of the space of test functions R. The set of generalized

functions, corresponding to a given space of test functions R, we

denote by R′. The set R′ is naturally equipped with the structure of

a linear space. If R ⊂ L2(M), then any function g ∈ L2(M) defines

a linear functional on R by the formula g(φ) =∫M g(x)φ(x)dµ.

Moreover, this integral also makes sense for other functions g.

In other words, under certain conditions, a conventional function

Page 441: Mathematical Foundations of Quantum Field Theory

March 30, 2020 10:29 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99a-app page 394

394 Mathematical Foundations of Quantum Field Theory

generates a generalized function, which will be denoted by the same

symbol.3

Generalized functions of n variables x1, . . . , xn ∈ M will be

defined as functionals f(φ1, . . . , φn), depending on n elements

φ1, . . . , φn ∈ R and linear in every argument (in other words, gener-

alized functions of n variables are defined as linear functionals on the

tensor product of n copies of the space R). We will denote a general-

ized function of n variables by the symbol f(x1, . . . , xn) and will write

that f(φ1, . . . , φn) =∫f(x1, . . . , xn)φ(x1) . . . φ(xn)dx1dx2 . . . dxn.

The set of generalized functions of n variables will be denoted by R′n.

Let us consider some examples of generalized functions.

1. Let C(Er) be the space of continuous functions on the Euclidean

space Er. By the symbols δ(x,a) and δ(x − a), we will denote

the generalized function (the functional on the space C(Er)) that

maps a function φ(x) ∈ C(Er) to φ(a), the function’s value at

the point a (the point a ∈ Er plays the role of a parameter).

In other words, the function δ(x − a) (δ-function) satisfies the

relation∫δ(x−a)φ(x)dx = φ(a). It is clear that for any subspace

R ⊂ C(Er), a δ-function is a generalized function from the set R′.

2. Let Cn(Er) be the space of n-times continuously differentiable

functions on the space Er. The generalized function δ(α)(x−a) ∈(Cn(Er))′ is defined by the relation∫

δ(α)(x− a)φ(x)dx = (−1)|α|(D(α)(φ))(a)

(here, α = (α1, . . . , αk) is a collection of non-negative integers

satisfying |α| = α1 + · · ·+ αk ≤ n, D(α)(φ) = ∂|α|

∂xα11 ...∂xαnn

φ).

3. Let us assume that a space R is contained in the space L2(M),

then we can define a generalized function of two variables

3Let us emphasize that while∫f(x)φ(x)dx is understood as another way of

writing the number f(φ) and is defined if φ is a test function and f is a generalizedfunction, in this section we use the notation

∫f(x)φ(x)dµ for the Lebesgue

integral of the function f(x)φ(x). Clearly, the Lebesgue integral is defined whenthe function f(x)φ(x) is summable. For integrals in the Euclidean space Er inboth cases, we use the same notation

∫f(x)φ(x)dx.

Page 442: Mathematical Foundations of Quantum Field Theory

March 30, 2020 10:29 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99a-app page 395

Appendix 395

δ(x, y) ∈ R′2 by means of the relation∫δ(x, y)φ(x)ψ(y)dxdy =

∫φ(x)ψ(x)dµ.

4. Every operator A in the space L2(M) specifies a generalized

function of two variables A(x, y) ∈ (DA)′2 by the formula∫A(x, y)φ(x)ψ(y)dxdy =

⟨Aψ, φ

⟩,

and the function A(x, y) is called the kernel of the operator A.

The operator A can be represented as an integral operator with

the kernel A(x, y). In other words, we have

(Aψ)(x) =

∫A(x, y)ψ(y)dy. (A.1)

The equation (A.1) should be understood as an abbreviation of

the following relation:∫(Aψ)(x)φ(x)dµ =

∫A(x, y)ψ(y)φ(x)dxdy.

5. Considering numerical generalized functions on the space Er, we

restrict ourselves to generalized functions of moderate growth,

defined as continuous functionals on the space S(Er) of smooth

functions having derivatives that tend to zero at infinity faster

than any power of the norm of x. The topology in this space is

described in A.6.

The linear space of generalized functions of moderate growth is

denoted by S ′(Er).

Every numerical function a(x) that grows at infinity no faster than

a polynomial (i.e. for large enough x, we have |a(x)| ≤ C‖x‖n) and

is summable in any bounded domain specifies a generalized function

a(φ), belonging to the space S ′, by the formula

a(φ) =

∫a(x)φ(x)dx

(the integral is understood as a Lebesgue integral).

Page 443: Mathematical Foundations of Quantum Field Theory

March 30, 2020 10:29 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99a-app page 396

396 Mathematical Foundations of Quantum Field Theory

Let us define some operations on generalized functions.

A generalized function f ∈ S ′ is called the limit of a sequence

fn ∈ S ′ of generalized functions, if for every function φ ∈ S,

the limit of the sequence fn(φ) is equal to f(φ) (in other words,

lim∫fn(x)φ(x)dx =

∫f(x)φ(x)dx).

The derivative fi = ∂f∂xi

of a generalized function f ∈ S ′ is defined

by the formula

fi(φ) = f

(− ∂φ∂xi

)or, in a different notation, by the formula∫

∂f(x)

∂xiφ(x)dx = −

∫f(x)

∂φ

∂xidx.

The Fourier transform of a generalized function f is defined by the

relation

f(φ) = f(φ), where φ(x) = (2π)−r2

∫exp(i 〈k, x〉)φ(k)dk.

Note that the derivative of a function f ∈ S ′ and the Fourier

transform of this function always exist and belong to the space

S ′ (this is obvious because together with the function φ ∈ S, the

functions ∂φ∂xi

and φ also belong to S).

The following important statement is called the kernel theorem.

Let us consider the functional f(φ, ψ) of φ ∈ S(Er), ψ ∈ S(En),

belonging to the space S ′(Er) for fixed ψ and S ′(En) for fixed φ.

Then, there exists one and only one functional f ∈ S ′(Er+n), obeying

f(φ⊗ ψ) = f(φ, ψ)

(the symbol φ⊗ ψ denotes the function

φ(x1, . . . , xr)ψ(y1, . . . , yn) ∈ S(Er+n)).

Vector generalized functions and operator generalized functions are

defined analogously to numerical generalized functions.

Namely, to define a vector generalized function, one should

construct for every function φ, from the space of test functions R,

a vector f(φ), from a linear space F , that linearly depends on φ.

Page 444: Mathematical Foundations of Quantum Field Theory

March 30, 2020 10:29 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99a-app page 397

Appendix 397

(In other words, a vector generalized function is an operator defined

on the space R and taking values in the linear space F .)

To define an operator generalized function, we should construct

for every function φ, from the space of test functions R, an operator

A(φ) acting linearly on elements in the space H1 and transforming it

into the space H2 (we assume that all operators A(φ) have the same

domain D). One can say that an operator generalized function is an

operator defined on the space of test functions R, taking values on

the linear space of operators with domain D ⊂ H1 and taking values

in H2.

For vector and operator generalized functions, we will use the

same notation as for numerical generalized functions

f(φ) =

∫f(x)φ(x)dx.

To describe some examples of vector generalized functions, we use

the following general constructions.

Let us suppose that a linear space L is embedded in a linear

space K and on the measure space M we have specified a function

f(x) with values in K. Let us also assume that for every test function

φ ∈ R there exists an integral∫f(x)φ(x)dµ and this integral4 belongs

to the space L. Then, it is clear that the correspondence f(φ) =∫f(x)φ(x)dµ can be considered as a generalized function with values

in L, generated by the conventional function f(x) with values in K.

We will use the same notation f(x) for this generalized function.

Let us consider for example the case where x ∈ Er, R = L2(Er) =

L,K = S ′(Er). Consider the functions δa(x) = δ(x−a) and φa(x) =

(2π)−r/2 exp(i 〈a, x〉) from K = S ′(Er) that depend on the parameter

a ∈ Er (i.e. we consider the functions on Er with values in S ′(Er)).If f(a) ∈ R = L2(Er), then

∫f(a)δa(x)da = f(x) ∈ L = L2(Er)

and∫f(a)φa(x)da = f(x) ∈ L = L2(Er). In other words, we have

constructed vector generalized functions δa(x), φa(x) with values in

L2(Er).

4The integral in this formula can be understood in any sense; it is importantonly that it is linear with respect to φ.

Page 445: Mathematical Foundations of Quantum Field Theory

March 30, 2020 10:29 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99a-app page 398

398 Mathematical Foundations of Quantum Field Theory

Let us suppose that D is a linear subspace in H1 and the space

H2 is embedded in the space H2. Further, let us assume that to every

point x of a measure space M , we have constructed an operator A(x)

with domain D and taking values in H2. Finally, if we assume that

for every function φ, the operator A(φ) =∫A(x)φ(x)dµ transforms

D into H2, then the conventional operator function A(x) generates a

generalized operator function A(φ) with values in the set of operators

transforming D into H2.

In particular, if M = Er, R = L2(Er), D = H1 = Fn, H2 =

Fn+1, where Fn is the space of symmetric (antisymmetric) square-

integrable functions f(x1, . . . ,xn) depending on variables x1, . . . ,

xn ∈ Er, and H2 = S ′(Er(n+1)) is the space of generalized functions

of moderate growth, depending on the variables x1, . . . ,xn+1 ∈Er, then the operator function a+

n (x) on En, where a+n (x) is an

operator transforming the function f(x1, . . . ,xn) ∈ Fn into the

function√n+ 1P [f(x1, . . . ,xn)δ(xn+1 − x)] ∈ H2, specifies an

operator generalized function, acting from Fn to Fn+1. This is clear

because the operator∫φ(x)a+(x)dx transforms Fn into Fn+1 (in the

above formula, P denotes the operator of symmetrization (antisym-

metrization)). This operator generalized function is considered in

Section 3.2.

If the space of test functions R is dense in L2(M), then the vector

generalized function f(x) with values in H is called normalized (or,

more precisely, δ-normalized) if the following condition is satisfied:

〈f(x), f(y)〉 = δ(x, y)

(to make this equation precise, we should integrate it, multiplying

by test functions φ(x), ψ(y) and writing it in the form⟨∫f(x)φ(x)dx,

∫f(y)ψ(y)dy

⟩=

∫δ(x, y)φ(x)ψ(y)dxdy

=

∫φ(x)ψ(x)dµ

or, in the shorter form 〈f(φ), f(ψ)〉 = 〈φ, ψ〉).This means that for a generalized function f that is δ-normalized,

we can construct an isometric map of the space R into H, sending

Page 446: Mathematical Foundations of Quantum Field Theory

March 30, 2020 10:29 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99a-app page 399

Appendix 399

the function φ ∈ R into the vector f(φ) ∈ H. This map can be

extended continuously to an isometric map of the space L2(M) into

H, with the extended map denoted by the same symbol f . If linear

combinations of vectors f(φ) are dense in H, then the isometric map

of L2(M) into H is an isomorphism (i.e. it is a unitary map onto H).

A δ-normalized generalized vector function f having this property

is called a generalized basis. In this case, the statements given above

imply that every vector a ∈ H can be expanded with respect to the

generalized basis (i.e. in other words a has a unique representation

in the form

a = f(φ) =

∫φ(x)f(x)dx,

where φ ∈ L2(M)). If a = f(φ), then the function φ(x) can be written

in the form

φ(x) = 〈a, f(x)〉 .

The generalized vector functions δa and φa defined above can be

considered as generalized bases.

If A is an operator on the space H, then the generalized function

〈x|A|y〉 = 〈Af(y), f(x)〉

is called the matrix of the operator A in the generalized basis f(x).

Note the following useful statement: let us consider a sequence

An of operators, assuming that the norms ‖An‖ form a bounded

sequence and the matrices 〈x|An|y〉 converge to the matrix 〈x|A|y〉,

then the operatorsAn weakly converge to the operatorA

[saying that

the matrices 〈x|An|y〉 converge to 〈x|A|y〉, we have in mind that for

all functions φ, ψ ∈ R, where R is a dense subspace of L2(M), the

following relation holds:

lim

∫〈x|An|y〉φ(x)ψ(y)dxdy =

∫〈x|A|y〉φ(x)ψ(y)dxdy

].

Let us fix the space of test functions R and a dense linear subspace D

in the Hilbert spaceH. We will consider only the generalized operator

functions that transform D into itself (i.e. in other words, we assume

Page 447: Mathematical Foundations of Quantum Field Theory

March 30, 2020 10:29 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99a-app page 400

400 Mathematical Foundations of Quantum Field Theory

that the domain of the operator A(φ), where φ is a test function,

coincides with D and all vectors A(φ)a, where a ∈ D, are contained

in D).

The set of all operator functions of this kind will be denoted

by MD.

We call generalized operator functions A ∈ MD, B ∈ MD

conjugate if for any a ∈ D, b ∈ D

〈A(x)a, b〉 = 〈a,B(x)b〉

(more precisely, if 〈A(φ)a, b〉 =⟨a,B(φ)b

⟩). In this case, we will use

the symbol B = A+.

The product of generalized operator functions A ∈MD, B ∈MD

is defined as the generalized operator function

C(x, y) = A(x)B(y).

To be more precise, the generalized operator function C maps a test

function α(x, y) of the form

α(x, y) =∑v

φv(x)ψv(y), (A.2)

where φv, ψv ∈ R, into the operator

C(α) =

∫(α(x, y)A(x)B(y)dxdy =

∑v

A(φv)B(ψv).

In some cases, the generalized operator function C(x, y) can be

defined on a broader class of test functions than functions in the

form (A.2). As an example, consider the following statement (the

operator analog for the kernel theorem).

Let us suppose that the space of test functions is the space

S(Er). Let us fix in the set MD a subset RD, consisting of

operator generalized functions A ∈ MD, satisfying the following

conditions: (a) the functional 〈A(φ)a, b〉 for any a, b ∈ D is continuous

in the topology of the space S (in other words, the numerical

generalized functions 〈A(x)a, b〉 belong to the space S ′); (b) there

exists an operator generalized function A+ ∈ MD conjugate to A.

Page 448: Mathematical Foundations of Quantum Field Theory

March 30, 2020 10:29 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99a-app page 401

Appendix 401

If A1, . . . , Ap ∈ RD, then to every function ψ ∈ S(Epr) we can assign

an operator

C(ψ) =

∫ψ(x1, . . . , xp)A1(x1) . . . Ap(xn)dpx, (A.3)

defined on the set D. Namely, we should take

C(ψ) =s∑

k=1

A1(φ(k)1 ) . . . Ap(φ

(k)p )

in the case when

ψ =

s∑k=1

φ(k)1 (x1) . . . φ(k)

p (xp), (A.4)

and in general, we should represent ψ as limit of a sequence ψn of

functions of the form (A.4), converging to ψ in the topology of S and

take

C(ψ) = slimC(ψn). (A.5)

The existence of this limit follows from the following estimates:

‖C(ψm)a− C(ψn)a‖2 = 〈C(ψm)a− C(ψn)a,C(ψm)A− C(ψn)a〉

=⟨(C+(ψm)C(ψm) + C+(ψn)C(ψn)

−C+(ψm)C(ψn)− C+(ψn)C(ψm))a, a⟩

=

∫ρm,n(x1, . . . , xp|y1, . . . , yp)

⟨A+p (xp)

. . . A+1 (x1)A1(y1) . . . Ap(yp)a, a

⟩dpxdpy,

(A.6)

where

ρm,n(x1, . . . , xp|y1, . . . , yp) = ψm(x1, . . . , xp)ψm(y1, . . . , yp)

+ψn(x1, . . . , xp)ψn(y1, . . . , yp)

−ψn(x1, . . . , xp)ψ(y1, . . . , yp)

−ψm(x1, . . . , xp)ψn(y1, . . . , yp).

Page 449: Mathematical Foundations of Quantum Field Theory

March 30, 2020 10:29 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99a-app page 402

402 Mathematical Foundations of Quantum Field Theory

It is easy to see that for m→∞, n→∞, the function ρm,n tends to

zero in the topology of the space S(E2pr).

Using the kernel theorem, one can conclude from this fact that the

expression A.6 tends to zero for m → ∞, n → ∞. This means that

the sequence C(ψn)a is a Cauchy sequence, hence it is convergent

for every a ∈ D. This proves that the sequence of operators C(ψn)

is strongly convergent; hence, we can define the operator C(ψ) by

means of the formula (A.5).

Let us consider now the set D′ consisting of linear combinations

of vectors of the form C(ψ)a, where a ∈ D, and C(ψ) is an operator

of the form (A.3), constructed by means of the operator analog of the

kernel theorem. Let us show that the expression of the form (A.3)

specifies an operator with domain D′ by the formula

C(ψ)(C ′(ψ′)a) = (C(ψ)C ′(ψ′))a,

where

C ′(ψ′) =

∫ψ′(x1, . . . , xq)A

′1(x1) . . . A′q(xq)d

qx,

C(ψ)C ′(ψ′) =

∫ψ(x1, . . . , xp)ψ

′(x′1, . . . , x′q)A1(x1)

. . . Ap(xp)A′1(x′1) . . . A′q(x

′q)d

pxdqx′.

To check that this definition is correct, we should verify that the

vector C(ψ)d does not depend on the representation of the vector d

in the form d = C ′(ψ′)a. We should also check that the constructed

operator is linear. The proof is based on the following remark.

Assume that the vector ξn has the form

ξn =

s(n)∑i=1

C(i)n (φ(i)

n )ai,

ξn → 0, and the sequence

ηn =

s(n)∑i=1

(C(ψn)C(i)n (φ(i)

n ))ai,

Page 450: Mathematical Foundations of Quantum Field Theory

March 30, 2020 10:29 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99a-app page 403

Appendix 403

is convergent. Then ηn → 0.

[Here, ai ∈ D,φ(i)

n ∈ S, ψn ∈ S, ψn → ψ

in the topology of S, A(i)k ∈ RD,

C(i)n (φ(i)

n ) =

∫φ(i)n (y1, . . . , yp(i))A

(i)1 (y1) . . . A

(i)p(i)(yp(i))d

p(i)y.

]This is true because for every vector ζ ∈ D

| 〈ηn, ζ〉 | = |⟨ξn, C

+(ψn)ζ⟩| ≤ ‖ξn‖‖C+(ψn)ζ‖ → 0.

A.8 Eigenvectors and generalized eigenvectors

We call a vector φ ∈ H an eigenvector of the operator A, acting on the

spaceH, if Aφ = λφ and φ 6= 0. The number λ is called an eigenvalue,

corresponding to the eigenvector φ. The set Hλ of eigenvectors for a

given eigenvalue λ is in fact a linear subspace; the dimension of this

space is called the multiplicity of the eigenvalue λ. An eigenvalue

with multiplicity one is called non-degenerate or simple.

One says that the operator A has discrete spectrum if linear

combinations of its eigenvectors are dense in the space H.

The eigenvalues of a self-adjoint operator are real; the eigenvectors

corresponding to distinct eigenvalues of a self-adjoint operator are

orthogonal.

If a self-adjoint operator has discrete spectrum, then the spectrum

of this operator is the closure of its set of eigenvalues. The spectral

projection Eµ is equal to a sum of projections on the linear subspace

Hλ, where λ ≤ µ.

Suppose that M is a measure space, R is a space of test functions

defined on M , and f ∈ R′ is a generalized vector function of the

variable x ∈ M with values in H. Then the function f is called a

generalized eigenfunction of an operator A, acting on the space H, if

Af(x) = a(x)f(x).

More precisely, the above equation means that

A

∫f(x)φ(x)dx =

∫f(x)a(x)φ(x)dx,

Page 451: Mathematical Foundations of Quantum Field Theory

March 30, 2020 10:29 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99a-app page 404

404 Mathematical Foundations of Quantum Field Theory

i.e. Af(φ) = f(aφ). The last equation must be satisfied in the cases

when it makes sense (i.e. when φ ∈ R and aφ ∈ R).

The generalized vector function φa = (2π)−r/2 exp(iax), defined

in (A.6), turns out to be a generalized eigenfunction for the operators1i∂∂x1

, . . . , 1i∂∂xr

(and more broadly for differential operators with

constant coefficients):

1

i

∂xkφa = akφa.

Note the following important statement.

For any commuting self-adjoint operators A1, . . . , An, acting on

the Hilbert space H, we can find a generalized basis of functions that

are eigenfunctions for each of the operators (more precisely, we can

find a generalized basis f that is δ-normalized (〈f(x), f(y)〉 = δ(x, y))

and satisfies the equation Aif(x) = ai(x)f(x) for every i = 1, . . . , n).

The proof of this statement is easily obtained if we recall that

there exists an isomorphism α of the space H and the space

L2(M), by which the operators Ai are transformed into operators of

multiplication by the function ai(x). Indeed, the generalized vector

function f can be represented by the formula

f(φ) =

∫f(x)φ(x)dx = α−1φ,

where φ ∈ L2(M).

A.9 Group representations

Suppose that for every group element g of a group G, there is a

corresponding unitary operator Tg on the Hilbert space H, such

that the identity operator corresponds to the unit of the group

and the multiplication of operators corresponds to the multiplication

operation in the group:

Te = 1, Tgh = TgTh. (A.7)

We call this correspondence a unitary representation of the group G.

A subspace H1 ⊂ H is called invariant if every operator Tgtransforms H1 into itself. A unitary representation H is called

Page 452: Mathematical Foundations of Quantum Field Theory

March 30, 2020 10:29 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99a-app page 405

Appendix 405

irreducible if the space H has no non-trivial (differing from the zero

subspace and the full space H) invariant subspace.

Two unitary representations Tg and T ′g on the spaces H and H′are equivalent if there exists an isomorphism α between the spaces

H and H′ satisfying

T ′gα = αTg.

For any self-adjoint operator A, we can construct a unitary

representation of the group of real numbers E1 by assigning to every

number t ∈ E1 the operator Tt = exp(iAt) (we use addition as the

group operation on E1 so that equation (A.7) takes the form

T0 = 1, Tt+τ = TtTτ ).

Conversely, given a unitary representation Tt of the group E1 such

that for any two vectors a, b ∈ H, the function 〈Tta, b〉 is continuous

or at least measurable, then we can find a self-adjoint operator A such

that Tt = exp(iAt); this operator can be obtained by the formula

A = 1i limt→0

Tt−1t (Stone’s theorem).

A family of operators satisfying the conditions of Stone’s theorem

is called a one-parameter group of unitary operators and the operator

A is called the generator (or an infinitesimal generator) of the group.

A Poincare group P is the group of affine transformations of

Minkowski space

x′ = Λx+ a, (A.8)

preserving the space–time interval and the sign of x0.

(Minkowski space is to be understood as a four-dimensional space

with a scalar product (x, y) = x0y0 − 〈x,y〉 = x0y0 − x1y1 − x2y2 −x3y3. The space–time interval is defined as

√(x− y, x− y). We write

vectors from Minkowski space as x = (x0,x), where x0 ∈ E1,

x ∈ E3.)

The transformation (A.8), viewed as an element of the group P,

will be denoted by (Λ, a).

Let Tg be a unitary representation of the group P such that for any

vectors a, b ∈ H the function 〈Tga, b〉 is continuous in g (we will only

consider unitary representations satisfying this continuity condition).

Page 453: Mathematical Foundations of Quantum Field Theory

March 30, 2020 10:29 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99a-app page 406

406 Mathematical Foundations of Quantum Field Theory

The group P contains a subgroup of translations (1, a), consisting

of transformations x′ = x + a. This subgroup is commutative. By

Stone’s theorem, the unitary operator T(1,a), corresponding to the

translations (1, a), can be represented in the form

T(1,a) = exp[−i(a0H − a1P1 − a2P2 − a3P3)]

= exp[−i(a0H − aP)],

where H,P1, P2, P3 are commuting self-adjoint operators.

The operator H is called the energy operator and the vector

operator P = (P1, P2, P3) is the momentum operator.

It is easy to check that the joint spectrum of the operators H,P

is invariant with respect to homogenous Lorentz transformations.

One of the simplest representations of the group P can be written

in the following way.

Let us consider the set Um in four-dimensional space specified

by the relation p20 − p2 = m2, p0 ≥ 0. To every function f ∈

L2(E3), we will assign a function f on the space Um defined by the

formula f(p0,p) = f(p)√p0. This assignment can be viewed as an

isomorphism between the space L2(E3) and some space of functions

on Um. This space of functions on Um is denoted by L2(Um); the

inner product on L2(Um) is defined by the formula

⟨f , g⟩

=

∫f(p0,p)g(p0,p)

dp

p0.

On the space L2(Um), we define a unitary representation of the group

P by the formula

(U(Λ, a)f)(p) = exp(i(a, p))f(Λ−1p) (A.9)

(to check that the formula (A.9) defines a unitary representation,

observe that the expression p−10 dp defines a Lorentz-invariant mea-

sure on Um). The resulting representation is continuous; the spectrum

of the operators H,P on the space L2(Um) coincides with the set Um.

By the isomorphism between L2(E3) and L2(Um), demon-

strated above, there exists equivalent unitary representation V (Λ, a)

Page 454: Mathematical Foundations of Quantum Field Theory

March 30, 2020 10:29 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99a-app page 407

Appendix 407

on L2(E3):

(V (Λ, a)f) = U(Λ, a)f .

Let us describe (up to equivalence) all the irreducible representa-

tions of the group P in which the energy operator H is non-negative

and not equal to zero.

For every non-negative number m, there exists an infinite series

of unitary irreducible representations, where the joint spectrum of

the operators H,P in these representations is the set Um. The

representation of this series will be denoted by the symbols (m,σ),

where σ is a non-negative integer for positive m and any integer for

m = 0.

A representation of the type (m,σ) can be realized via unitary

operators on the space L2(E3 × S), where S is a set consisting of

2σ + 1 elements for positive m and of one element for m = 0. Here,

the energy and momentum operators on L2(E3×S) are given by the

formulas

Hf(p, s) =√

p2 +m2f(p, s), (A.10)

Pf(p, s) = pf(p, s). (A.11)

For example, the representation described above is a representation

of type (m, 1).

Let us note that in addition to the single-valued unitary rep-

resentations of P described here, there exist two-valued unitary

representations. Irreducible two-valued representations, where the

energy operator is non-negative and is not identically zero, can be

described in the same way as single-valued (with the difference that

σ must be a half-integer).

Page 455: Mathematical Foundations of Quantum Field Theory

b2530 International Strategic Relations and China’s National Security: World at the Crossroads

b2530_FM.indd 6 01-Sep-16 11:03:06 AM

This page intentionally left blankThis page intentionally left blankThis page intentionally left blankThis page intentionally left blank

Page 456: Mathematical Foundations of Quantum Field Theory

March 30, 2020 10:34 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99b-bib page 409

Bibliography

Araki, H. and Haag, R. (1967). Collision cross sections in terms oflocal observables, Communications in Mathematical Physics 4, 2,pp. 77–91.

Belavin, A. A., Polyakov, A. M., and Zamolodchikov, A. B. (1984).Infinite conformal symmetry in two-dimensional quantum field theory,Nuclear Physics B 241, 2, pp. 333–380.

Berezin, F. (2012). The Method of Second Quantization, Vol. 24 (Elsevier).Bethe, H. A. (1947). The electromagnetic shift of energy levels, Physical

Review 72, 4, p. 339.Bogolyubov, N., Logunov, A., and Todorov, I. (1969). Fundamentals of the

Axiomatic Approach in Quantum Field Theory, Nauka: Moscow [InRussian].

Bogolyubov, N., Medvedev, A., and Polivanov, M. (1956). Problems of theTheory of Dispersion Relations (Princeton University).

Bogolyubov, N. and Parasyuk, O. (1955). A theory of multiplication ofcausative singular functions, Doklady Akademii Nauk SSSR 100,pp. 25–28.

Chu, H. and Umezawa, H. (1994). A unified formalism of thermal quantumfield theory, International Journal of Modern Physics A 9, 14,pp. 2363–2409.

Dyson, F. J. (1949). The radiation theories of Tomonaga, Schwinger, andFeynman, Physical Review 75, 3, p. 486.

Faddeev, L. (1975). Hadrons from leptons, Tech. rep., VA Steklov Mathe-matics Institute.

Faddeev, L. D. (1963). Mathematical questions in the quantum theory ofscattering for a system of three particles, Trudy MatematicheskogoInstituta imeni VA Steklova 69, pp. 3–122.

409

Page 457: Mathematical Foundations of Quantum Field Theory

March 30, 2020 10:34 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99b-bib page 410

410 Mathematical Foundations of Quantum Field Theory

Fateev, V. A. and Shvarts, A. (1973). On axiomatic scattering theory,Theoretical and Mathematical Physics 14, 2, pp. 112–124.

Fedoryuk, M. V. (1971). The stationary phase method and pseudodifferen-tial operators, Russian Mathematical Surveys 26, 1, p. 65.

Feynman, R. P. (2005). Space-time approach to non-relativistic quantummechanics, in Feynman’s Thesis — A New Approach To QuantumTheory (World Scientific: Singapore), pp. 71–109.

Haag, R. (1958). Quantum field theories with composite particles andasymptotic conditions, Physical Review 112, 2, p. 669.

Haag, R. and Kastler, D. (1964). An algebraic approach to quantum fieldtheory, Journal of Mathematical Physics 5, 7, pp. 848–861.

Hepp, K. (1965). On the connection between the LSZ and Wightmanquantum field theory, Communications in Mathematical Physics 1,2, pp. 95–111.

Hepp, K. (1966). Proof of the Bogoliubov–Parasiuk theorem on renormal-ization, Communications in Mathematical Physics 2, 1, pp. 301–326.

Hepp, K. and Epstein, H. (1971). Analytic Properties of Scattering Ampli-tudes in Local Quantum Field Theory [Russian translation].

Hunziker, W. and Sigal, I. M. (2000). The quantum n-body problem,Journal of Mathematical Physics 41, 6, pp. 3448–3510.

Jaffe, A. (2000). Constructive quantum field theory, Mathematical Physics,pp. 111–127.

Jost, R. (1965). The General Theory of Quantized Fields, in Lectures inApplied Mathematics, Vol. IV (American Mathematical Society).

Kato, T. (2013). Perturbation Theory for Linear Operators, Vol. 132(Springer Science & Business Media).

Kuroda, S. T. (1959). On the existence and the unitary property of the scat-tering operator, Il Nuovo Cimento (1955–1965) 12, 5, pp. 431–454.

Lehmann, H., Symanzik, K., and Zimmermann, W. (1955). Zur vertexfunk-tion in quantisierten feldtheorien, Il Nuovo Cimento (1955–1965) 2,3, pp. 425–432.

Likhachev, V., Tyupkin, Y. S., and Shvarts, A. (1970). The adiabatic Smatrix and quasiparticles, Theoretical and Mathematical Physics 2,1, pp. 1–20.

Likhachev, V., Tyupkin, Y. S., and Shvarts, A. (1972). Adiabatic theoremin quantum field theory, Theoretical and Mathematical Physics 10, 1,pp. 42–55.

Osterwalder, K. and Schrader, R. (1973). Axioms for EuclideanGreen’s functions, Communications in Mathematical Physics 31, 2,pp. 83–112.

Polyakov, A. (1974). Spectrum of particles in quantum field theory, JETPLetters 20, pp. 430–433.

Page 458: Mathematical Foundations of Quantum Field Theory

March 30, 2020 10:34 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99b-bib page 411

Bibliography 411

Robertson, W. et al. (1980). Topological Vector Spaces, Vol. 53 (CambridgeUniversity Press Archive).

Ruelle, D. (1962). On asymptotic condition in quantum field theory,Helvetica Physica Acta 35, 3, p. 147.

Schwarz, A. (2019a). Geometric approach to quantum theory, arXiv:1906.04939.

Schwarz, A. (2019b). Inclusive scattering matrix and scattering of quasi-particles, arXiv:1904.04050.

Schwarz, A. (2019c). Scattering in algebraic approach to QFT inclusivescattering matrix, arXiv:1910.00000.

Schwarz, A. S. (1967). A new formulation of the quantum theory, DokladyAkademii Nauk 173, 4, pp. 793–796.

Schwarz, A. S. and Tyupkin, Y. S. (1987). Measurement theory andthe Schroedinger equation, in Quantum Field Theory and QuantumStatistics: Essays in Honour of the Sixtieth Birthday of ES Frad-kin. V. 1.

Schwinger, J. (1958). Selected Papers on Quantum Electrodynamics(Courier Corporation).

Sigal, I. M. and Soffer, A. (1987). The n-particle scattering problem:Asymptotic completeness for short-range systems, Annals of Math-ematics, 126, 1, pp. 35–108.

Soffer, A. (2006). Soliton dynamics and scattering, International Congressof Mathematicians, Vol. 3, pp. 459–471.

Streater, R. F. and Wightman, A. S. (2016). PCT, Spin and Statistics, andAll That (Princeton University Press).

Takhtadzhyan, L. A. and Faddeev, L. D. (1974). Essentially nonlinear one-dimensional model of classical field theory, Theoretical and Mathemat-ical Physics 21, 2, pp. 1046–1057.

Tao, T. (2009). Why are solitons stable? Bulletin of the American Mathe-matical Society 46, 1, pp. 1–33.

Tiupkin, I. S., Fateev, V. A., and Shvarts, A. (1975). Classical limit ofscattering matrix in quantum field theory, Akademii Nauk SSSRDoklady, Vol. 221, pp. 70–73.

Tomonaga, S.-i. (1946). On a relativistically invariant formulation of thequantum theory of wave fields. Progress of Theoretical Physics 1, 2,pp. 27–42.

Tyupkin, Y. S. (1973). On the adiabatic definition of the S matrix in theformalism of L functionals, Theoretical and Mathematical Physics 16,2, pp. 751–756.

Tyupkin, Y. S., Fateev, V., and Shvarts, A. (1975). On the existence ofheavy particles in gauge field theories, Pis’ma v Zhurnal Ehksperi-mental’noj i Teoreticheskoj Fiziki 21, 1, pp. 91–93.

Page 459: Mathematical Foundations of Quantum Field Theory

March 30, 2020 10:34 Mathematical Foundations of Quantum Field Theory 9in x 6in 11122-x99b-bib page 412

412 Mathematical Foundations of Quantum Field Theory

Tyupkin, Y. S. and Shvarts, A. (1972). On the adiabatic change ofa stationary state, Theoretical and Mathematical Physics 10, 2,pp. 172–175.

Velo, G. and Wightman, A. S. (2012). Constructive Quantum Field TheoryII, Vol. 234 (Springer Science & Business Media).

Wightman, A. S. (1956). Quantum field theory in terms of vacuumexpectation values, Physical Review 101, 2, p. 860.

Page 460: Mathematical Foundations of Quantum Field Theory

April 1, 2020 14:21 Mathematical Foundations of Quantum Field Theory 9in x 6in 11222-x99c-index page 413

Index

adiabatic, viii, xxxi, 43, 45, 52, 101,150, 166, 170, 253, 254, 309,362–365, 367, 368

angular momentum, xxxiv, 9–12, 341

asymptotically commutative, 194,233, 242, 292–294, 313

CAR, xxxvi, xxxviii, xxxix, 30, 75,76, 79, 82–86, 88, 93, 97, 98,109–113, 120, 123, 133, 156, 163,179, 334, 355, 356

CCR, xxviii, xxix, xxxiv, xxxvi,xxxviii, xxxix, 30, 69, 75–77, 87,90, 93, 95, 97, 98, 109–113, 120,126, 127, 133, 141, 156, 163, 179,181, 193, 277, 281–285, 290, 327

correlation function, xxvi, xxvii, xl,333, 335, 336, 343

cross section, xix, xlii, 58–61, 162,236–238, 345, 354, 368–370

decoherence, xxxi, xxxii, xliii

dressed particle, xix, 150

dressing operator, 153, 178, 229, 232,233, 294, 311

entropy, 331, 332, 334

equilibrium, xxix, xxxii, xl, xli, 181,331–336, 338, 340, 355

formal Hamiltonian, xxxviii, xxxix,126, 127, 275, 277, 327, 359

generalized function, xxii, xxxviii, 30,393–400

generalized Green function, xix, 353,363

GNS (Gelfand–Naimark–Segal)construction, xxv, xxvi, 337, 341,357, 360, 362

Green function, xxvi, xxvii, 111–114,116, 118–122, 128, 161, 327,350–353, 366, 370

Heisenberg equation, xxxiii, 6, 7, 18,84, 121, 138, 139, 143, 278–280,285, 290, 305

inclusive, xix, xlii, 345, 353, 368–370,372, 373

KMS (Kubo–Martin–Schwinger)condition, xli, 333, 335, 336, 340

LSZ(Lehmann–Symanzik–Zimmerman)formula, 163, 350, 351, 371, 372

measurable function, ix, 259,379–381, 386, 388, 390, 393

measure, xxiii, xxxii, 4, 5, 378, 380,381, 383, 388, 406

measure space, ix, xlv, xlvi, 4,378–380, 386, 387, 389, 393, 397,398, 403

413

Page 461: Mathematical Foundations of Quantum Field Theory

April 1, 2020 14:21 Mathematical Foundations of Quantum Field Theory 9in x 6in 11222-x99c-index page 414

414 Mathematical Foundations of Quantum Field Theory

momentum, xxi, xxxiv, xli, xliii, xlvi,3, 4, 9–11, 14–17, 27, 29, 34–36,58–60, 62–64, 123, 124, 126, 138,142, 172, 184, 233, 275, 301, 335,407

Møller matrix, 73, 149, 196, 203, 240,249, 253, 274, 279, 307–309, 311

operator realization, viii, xxxix,126–129, 138, 139, 143–146, 153,154, 166, 184–188, 275, 277–279,286–290, 292, 293, 325, 336

oscillator, xxxv, 17, 18, 20, 21, 77,334

partition function, 331, 332, 334, 335perturbation theory, viii, xvii, xviii,

xxxvii, xxxix, 39, 44, 45, 65, 86,114, 129, 131, 170, 277, 301, 303,305–309, 311, 327, 328, 338, 359

Schrodinger equation, 1, 2, 7, 46, 63,280

soliton, xliii–xlv, 188, 189spin, 11–16, 23, 57, 123, 316, 326, 341stationary, xxvi, xxix–xxxi, 6, 7, 15,

16, 18–22, 27, 39, 43, 44, 46, 63,119, 168, 178, 209, 306, 332, 360,373

symmetry breaking, 186, 187

translation-invariant Hamiltonian,viii, xxxix, 64, 65, 68, 123–126, 128,147, 275, 311,

truncated, xl, 197–199, 333

Wightman function, xxvii, xxxix, xl,107–113, 322