You are on page 1of 226

Path Integrals in Field Theory

Advanced Texts in Physics


This pro gram of advanced texts covers a broad spectrum of topics which are of
current and emerging interest in physics. Each book provides a comprehensive and
yet accessible introduction to a field at the forefront of modern research. As such,
these texts are intended for senior undergraduate and graduate students at the MS
and PhD level; however, research scientists seeking an introduction to particular
areas of physics will also benefit from the titles in this collection.

Springer-Verlag Berlin Heidelberg GmbH

ONLINE LlBRARY
Physics and Astronomy
http://www.springer.de
Ulrich Mosel

Path Integrals
in Field Theory
An Introduction

With 19 Figures

13
Professor Dr. Ulrich Mosel
Institut für Theoretische Physik
Universität Giessen
Heinrich-Buff-Ring 16
35392 Giessen, Germany

Library of Congress Cataloging-in-Publication Data


Mosel, Ulrich, 1943 - Path Integrals in Fields Theory: An Introduction 1 U. Mosel. p.cm. - (Advanced
texts in physics, Issn 1439-2674). Includes bibliographical referenees and index.
ISBN 978-3-540-40382-1 ISBN 978-3-642-18797-1 (eBook)
DOI 10.1007/978-3-642-18797-1 1. Path integrals. 2. Quantum field theory. I. Title. II. Series.
QCI74.52.P3M67 2003 530.14'3-dc21

ISSN 1439-2674
ISBN 978-3-540-40382-1
This work is subjeet to copyright. All rights are reserved, whether the whole or part of the material
is coneerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation, broad-
casting, reproduetion on microfilm or in any other way, and storage in data banks. Duplication of
this publieation or parts thereof is permitted only under tlte provisions of tlte German Copyright Law
of September 9, 1965, in its current version, and permission for use must always be obtained from
Springer-Verlag. Violations are liable for proseeution under the German Copyright Law.

http://www.springer.de
© Springer-Verlag Berlin Heidelberg 2004
Originally published by Springer-Verlag Berlin Heidelberg New York in 2004
The use of general descriptive names, registered names, trademarks, ete. in this publication does not
imply, even in the absence of a specifie statement, that such names are exempt from the relevant pro-
tective laws and regulations and therefore free for general use.
Typesetting: Data prepared by the author using a Springer TEX maero paekage
Cover design: design & production GmbH, Heidelberg
Printed on acid-free paper 54/3141/tr 54 32 10
To all my students
Preface

This is an introductory book to path integral methods in field theories. It is


aimed at graduate students and physicists who need a working knowledge of
field theory and its methods for applications in hadron, particle and nuclear
physics. While teaching field theory courses over the years I have found that
many books on field theory present the path integral methods used in only a
very recipe-like way. On the other hand, specialized books on path integrals
often contain many more details than are actually needed by non-specialists.
I hope that this book here fills the gap. It provides enough information to
actually follow all the arguments necessary for field theoretical developments
without, however, elaborating on the method as such and its mathemati-
cal intricacies. This book is – in a way – a technical companion to Fields,
Symmetries, and Quarks by the present author.
The reader of this book should have some knowledge of the relativistic
equations of motion of ’classical’ quantum theory, but no prior knowledge of
field theory is assumed. The material in this book can be covered in a one-
semester course with 3 hrs/week. It has evolved in many years of teaching
this subject. I am grateful to my students for many helpful questions and
comments and, in particular, to Frank Froemel for his help in preparing the
figures in this book.

Giessen Ulrich Mosel


June 2003
Contents

Part I Non-Relativistic Quantum Theory

1 The Path Integral in Quantum Theory . . . . . . . . . . . . . . . . . . . 3


1.1 Propagator of the Schrödinger Equation . . . . . . . . . . . . . . . . . . . 3
1.2 Propagator as Path Integral . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.3 Quadratic Hamiltonians . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.3.1 Cartesian Metric . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.3.2 Non-Cartesian Metric . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.4 Classical Interpretation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

2 Perturbation Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.1 Free Propagator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.2 Perturbative Expansion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.3 Application to Scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

3 Generating Functionals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3.1 Groundstate-to-Groundstate Transitions . . . . . . . . . . . . . . . . . . . 27
3.1.1 Generating Functional . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
3.2 Functional Derivatives
of Gs-Gs Transition Amplitudes . . . . . . . . . . . . . . . . . . . . . . . . . . 32

Part II Relativistic Quantum Field Theory

4 Relativistic Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
4.1 Equations of Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
4.1.1 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
4.2 Symmetries and Conservation Laws . . . . . . . . . . . . . . . . . . . . . . . 46
4.2.1 Geometrical Space–Time Symmetries . . . . . . . . . . . . . . . 47
4.2.2 Internal Symmetries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49

5 Path Integrals for Scalar Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . 53


5.1 Generating Functional for Fields . . . . . . . . . . . . . . . . . . . . . . . . . 53
5.1.1 Euclidean Representation . . . . . . . . . . . . . . . . . . . . . . . . . . 56
X Contents

6 Evaluation of Path Integrals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59


6.1 Free Scalar Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
6.1.1 Generating Functional . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
6.1.2 Feynman Propagator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
6.1.3 Gaussian Integration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
6.2 Interacting Scalar Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
6.2.1 Stationary Phase Approximation . . . . . . . . . . . . . . . . . . . 67
6.2.2 Numerical Evaluation of Path Integrals . . . . . . . . . . . . . 70
6.2.3 Real Time Formalism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72

7 Transition Rates and Green’s Functions . . . . . . . . . . . . . . . . . . 75


7.1 Scattering Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
7.2 Reduction Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
7.2.1 Canonical Field Quantization . . . . . . . . . . . . . . . . . . . . . . 77
7.2.2 Derivation of the Reduction Theorem . . . . . . . . . . . . . . . 78

8 Green’s Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
8.1 n-point Green’s Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
8.1.1 Momentum Representation . . . . . . . . . . . . . . . . . . . . . . . . 86
8.1.2 Operator Representations . . . . . . . . . . . . . . . . . . . . . . . . . 86
8.2 Free Scalar Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
8.2.1 Wick’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
8.2.2 Feynman Rules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
8.3 Interacting Scalar Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
8.3.1 Perturbative Expansion . . . . . . . . . . . . . . . . . . . . . . . . . . . 93

9 Perturbative φ4 Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
9.1 Perturbative Expansion of the Generating Function . . . . . . . . . 97
9.1.1 Generating Functional up to O(g) . . . . . . . . . . . . . . . . . . 98
9.2 Two-Point Function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
9.2.1 Terms up to O(g 0 ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
9.2.2 Terms up to O(g) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
9.2.3 Terms up to O(g 2 ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
9.3 Four-Point Function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
9.3.1 Terms up to O(g) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
9.3.2 Terms up to O(g 2 ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
9.4 Divergences in n-Point Functions . . . . . . . . . . . . . . . . . . . . . . . . . 110
9.4.1 Power Counting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
9.4.2 Dimensional Regularization of φ4 Theory . . . . . . . . . . . . 113
9.4.3 Renormalization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
Contents XI

10 Green’s Functions for Fermions . . . . . . . . . . . . . . . . . . . . . . . . . . . 125


10.1 Grassmann Algebra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
10.1.1 Derivatives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
10.1.2 Integration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128
10.2 Green’s Functions for Fermions . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
10.2.1 Generating Functional for Fermions . . . . . . . . . . . . . . . . . 134
10.2.2 Reduction Theorem for Fermions . . . . . . . . . . . . . . . . . . . 138
10.2.3 Green’s Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139

11 Interacting Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141


11.1 Feynman Rules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
11.1.1 Fermion Loops . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
11.2 Wick’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
11.3 Bosonization of Yukawa Theory . . . . . . . . . . . . . . . . . . . . . . . . . . 147
11.3.1 Perturbative Expansion . . . . . . . . . . . . . . . . . . . . . . . . . . . 150

Part III Gauge Field Theory

12 Path Integrals for QED . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157


12.1 Gauge Invariance in Abelian Free Field Theories . . . . . . . . . . . 157
12.2 Generating Functional . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
12.3 Gauge Invariance in QED . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162
12.4 Feynman Rules of QED . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164

13 Path Integrals for Gauge Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . 167


13.1 Non-Abelian Gauge Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
13.2 Generating Functional . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171
13.3 Gauge Fixing of L . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176
13.4 Faddeev–Popov Determinant . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178
13.4.1 Explicit Forms of the FP Determinant . . . . . . . . . . . . . . 180
13.4.2 Ghost Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182
13.5 Feynman Rules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184

14 Examples for Gauge Field Theories . . . . . . . . . . . . . . . . . . . . . . . 189


14.1 Quantum Chromodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
14.2 Electroweak Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190

Units and Metric . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 193


A.1 Units . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 193
A.2 Metric and Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 194
XII Contents

Functionals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 197
B.1 Definition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 197
B.2 Functional Integration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 197
B.2.1 Gaussian Integrals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 198
B.3 Functional Derivatives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201

Renormalization Integrals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 203

Gaussian Grassmann Integration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209

Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 211
Part I

Non-Relativistic Quantum Theory


1 The Path Integral in Quantum Theory

In this starting chapter we introduce the concepts of propagators and path


integrals in the framework of nonrelativistic quantum theory. In all these
discussions, and the following chapters on nonrelativistic quantum theory, we
work with one coordinate only, but all the results can be easily generalized
to the case of d dimensions.

1.1 Propagator of the Schrödinger Equation


We start by considering a nonrelativistic particle in a one-dimensional po-
tential V (x). The Schrödinger equation reads

h̄2 ∂ 2 ψ(x, t) ∂ψ(x, t)


Hψ(x, t) = − + V (x)ψ(x, t) = ih̄ . (1.1)
2m ∂x2 ∂t
This equation allows us to calculate the wavefunction ψ(x, t) at a later time,
if we know ψ(x, t0 ) at the earlier time t0 < t. For further calculations we
rewrite this equation into the following form
 

ih̄ − H ψ(x, t) = 0 . (1.2)
∂t

Next, we consider the function K (x, t; xi , ti ) which is defined as a solution


of the equation
 

ih̄ − H K (x, t; xi , ti ) = ih̄δ(x − xi )δ(t − ti ) . (1.3)
∂t

K is the “Green’s function” of the Schrödinger equation (K is also often


called the “propagator”) with the initial condition

K(x, ti + 0; xi , ti ) = δ(x − xi ) . (1.4)

The solution of the Schrödinger equation (1.2) can be written as



ψ(x, t) = K (x, t; xi , ti ) ψ(xi , ti ) dxi (1.5)

U. Mosel, Path Integrals in Field Theory


© Springer-Verlag Berlin Heidelberg 2004
4 1 The Path Integral in Quantum Theory

for t > ti (Huygen’s principle). Relation (1.5) can be proven by inserting the
lhs into the Schrödinger equation
 

ih̄ − H K (x, t; xi , ti ) ψ(xi , ti ) dxi
∂t

= ih̄ δ (t − ti ) δ (x − xi ) ψ(xi , ti ) dxi

= ih̄δ (t − ti ) ψ(x, ti ) = 0 for t > ti . (1.6)

Thus the ψ defined by (1.5) is indeed a solution of the Schrödinger equation


for all times t > ti . K (x, t; xi , ti ) is the probability amplitude for a transition
from xi , at time ti , to the position x, at the later time t. The restriction to
later times preserves causality.
We can find an explicit form for the propagator, if the solutions of the
stationary Schrödinger equation, ϕn (x), and the corresponding eigenvalues,
En , are known. Since the ϕn form a complete system, K can certainly be
expanded in this basis (for t ≥ ti )

an ϕn (x)e− h̄ En t Θ (t − ti ) .
i
K (x, t; xi , ti ) = (1.7)
n

Here the stepfunction Θ(t) = 0 for t < 0 and Θ(t) = 1 for t ≥ 0 takes
explicitly into account that we only propagate the wavefunction forward in
time. The expansion coefficients obviously depend on xi , ti

an = an (xi , ti ) . (1.8)

Because of the initial condition K(x, ti + 0; xi , ti ) = δ (x − xi ) we have


 i
δ (x − xi ) = an (xi , ti )ϕn (x)e− h̄ En ti . (1.9)
n

The lhs is time-independent; thus we must have

an (xi , ti ) = an (xi )e+ h̄ En ti ,


i
(1.10)

and consequently 
δ (x − xi ) = an (xi ) ϕn (x) . (1.11)
n
This can be fulfilled by
an (xi ) = ϕ∗n (xi ) (1.12)
(closure relation). Thus we have a representation of K (x, t; xi , ti ) in terms of
the eigenfunctions and eigenvalues of the underlying Hamiltonian

ϕ∗n (xi ) ϕn (x)e− h̄ En (t−ti ) .
i
K (x, t; xi , ti ) = Θ(t − ti ) (1.13)
n

It is easy to show that this propagator fulfills (1.3).


1.2 Propagator as Path Integral 5

In Dirac’s bra and ket notation this result can also be written as (for
t > ti )

ϕ∗n (xi ) ϕn (x)e− h̄ En (t−ti )
i
K (x, t; xi , ti ) =
n

n|xi e− h̄ En (t−ti ) x|n
i
=
n

n|e+ h̄ Ĥti |xi x|e− h̄ Ĥt |n
i i
=
n

= x|e− h̄ Ĥ(t−ti ) |xi  ≡ x|Û (t, ti ) |xi  .


i
(1.14)

Thus the propagator is nothing other than the time development operator

Û (t, ti ) = e− h̄ Ĥ(t−ti )
i
(1.15)

for t > ti in the x representation. It is also often written as

K (x, t; xi , ti ) = x|e− h̄ Ĥ(t−ti ) |xi  ≡ xt|xi ti 


i
(1.16)

for t > ti ; for t < ti it vanishes.


The notation here is that of the Heisenberg representation of quantum
mechanics. In this representation the physical state vectors are time-inde-
pendent and the operators themselves carry all the time-dependence, whereas
this is just the opposite for the Schrödinger representation. For example, for
the position operator x̂ in the Schrödinger representation with x̂|x = x|x
we obtain the time-dependent operator in the Heisenberg representation

x̂H (t) = e h̄ Ĥt x̂ e− h̄ Ĥt


i i
(1.17)

and
x̂H (t)|xt = x|xt (1.18)
with
|xt = e h̄ Ĥt |x .
i
(1.19)
The state |xt is thus the eigenstate of the operator x̂H (t) with eigenvalue x
and not the state that evolves with time out of |x; this explains the sign of
the frequency in the exponent.

1.2 Propagator as Path Integral

We start by dividing the time-interval between ti and t by inserting the time


t1 . The wavefunction is first propagated until t1 and then, in a second step,
until t
6 1 The Path Integral in Quantum Theory

ψ (x1 , t1 ) = K (x1 , t1 ; xi , ti ) ψ (xi , ti ) dxi (1.20)

ψ(x, t) = K(x, t; x1 , t1 )ψ (x1 , t1 ) dx1 .

Taking these two equations together we get


 
ψ(x, t) = K (x, t; x1 , t1 ) K (x1 , t1 ; xi , ti ) ψ(xi , ti ) dxi dx1 . (1.21)

Comparing this result with (1.5) yields



K (x, t; xi , ti ) = K (x, t; x1 , t1 ) K (x1 , t1 ; xi , ti ) dx1 . (1.22)

We can thus view the transition from (xi , ti ) to (x, t) as the result of a
transition first from (x, t) to all possible intermediate points (x1 , t1 ), which is
then followed by a transition from these intermediate points to the endpoint.
We could also say that the integration in (1.22) is performed over all possible
paths between the points (xi , ti ) and (x, t), which consist of two straight line
segments with a bend at t1 . This is illustrated in Fig. 1.1.

t1

ti
x
xi x
Fig. 1.1. Possible paths from xi to x, corresponding to (1.22)

We now subdivide the time interval further into (n + 1) equal parts of


length Δt = η. We then have in direct generalization of the previous result
 
K (x, t; xi , ti ) = . . . dx1 dx2 . . . dxn (1.23)

× [K (x, t; xn , tn ) K (xn , tn ; xn−1 , tn−1 ) . . . K (x1 , t1 ; xi , ti )] .


1.2 Propagator as Path Integral 7

The integrals run here over all possible paths between (xi , ti ) and (x, t) which
consist of (n + 1) segments with boundaries that are determined by the time
steps ti , t1 , . . . , tn , t.
We now calculate the propagator for a small time interval Δt = η from
tj to tj+1 . For this propagation we have according to (1.16)

K (xj+1 , tj+1 ; xj , tj ) = xj+1 |e− h̄ Ĥη |xj 


i
(1.24)
∼ i
= xj+1 |1 − Ĥη|xj 

i
= δ (xj+1 − xj ) − ηxj+1 |Ĥ|xj 
 h̄
1 iη
e h̄ p(xj+1 −xj ) dp − xj+1 |Ĥ|xj 
i
=
2πh̄ h̄
with the representation for the δ-function

 1 
δ(x − x ) = eik(x−x ) dk . (1.25)

We now assume that Ĥ is given by

Ĥ = T̂ (p̂) + V̂ (x̂) . (1.26)

Here T̂ , p̂, V̂ , x̂ are all operators; we assume that T (p̂) and V (x̂) are Taylor-
expandable. In this case, where the p- and x-dependencies separate, we can
also bring the last term in (1.24) into an integral form. We have

xj+1 |Ĥ|xj  = xj+1 |T̂ + V̂ |xj  . (1.27)

First, we consider the first summand



xj+1 |T̂ |xj  = dp dpxj+1 |p p |T̂ (p̂)|pp|xj  (1.28)

= dp dpxj+1 |p δ(p − p)T (p)p|xj 

= dpxj+1 |pT (p)p|xj  .

With the normalized momentum eigenfunctions


1
e h̄ px
i
x|p = √ (1.29)
2πh̄
we thus obtain

1
e h̄ p(xj+1 −xj ) T (p) dp .
i
xj+1 |T̂ (p̂)|xj  = (1.30)
2πh̄
8 1 The Path Integral in Quantum Theory

While there is an operator p̂ on the lhs of this equation there are only numbers
p on its rhs.
For the potential part an analogous transformation can be performed

xj+1 |V̂ (x̂)|xj  = V (xj ) δ (xj+1 − xj ) (1.31)



1
e h̄ p(xj+1 −xj ) dp V (xj ) .
i
=
2πh̄

Again, on the lhs the argument x̂ of V̂ is an operator, while the rhs of this
equation contains no operators.
In summary, we have for the propagator over a time-segment η

1
dp e h̄ p(xj+1 −xj )
i
K (xj+1 , tj+1 ; xj , tj ) =
2πh̄
   
iη 1 i
p(xj+1 −xj ) 1 i
p(xj+1 −xj )
− dp e h̄ T (p) + dp e h̄ V (xj )
h̄ 2πh̄ 2πh̄
  
1 iη
dp e h̄ p(xj+1 −xj ) 1 − H (p, xj )
i
=
2πh̄ h̄
  
−→ 1 dp exp
i
[p (x − x ) − ηH (p , x )] . (1.32)
η→0 j j j+1 j j j
2πh̄ h̄
Here H = T + V is a function of the numbers x and p and no longer an
operator! In the last step we have, therefore, renamed the integration variable
to pj to indicate that it may be viewed as the momentum of a classical particle
moving from xj to xj+1 between times tj and tj+1 .
We now insert (1.32) into (1.23), take the limit n → ∞, and obtain the
so-called Hamiltonian path integral

K (x, t; xi , ti ) (1.33)
⎧ ⎫
  n  
n ⎨i  n ⎬
dpl
= lim dxk exp [pj (xj+1 − xj ) − ηH (pj , xj )] ,
n→∞ 2πh̄ ⎩ h̄ ⎭
k=1 l=0 j=0

(with x0 = xi and xn+1 = x). The asymmetry in the range of the products
over x- and p-integrations comes about because with n intermediate steps
between xi and x there are n + 1 intervals and corresponding momenta.
The integrand here is, for finite n, a complex function of all the coordinates
x1 , x2 , . . . , xn and the momenta p1 , p2 , . . . , pn . In the limit n → ∞ it depends
on the whole trajectory x(t), p(t). Here we note that p is not the momentum
canonically conjugate to the coordinate x, but instead just an integration
variable.
In the limit n → ∞ we obtain for the exponent
n

[pj (xj+1 − xj ) − ηH (pj , xj )]
j=0
1.3 Quadratic Hamiltonians 9

n  
xj+1 − xj
= η pj − H (pj , xj )
j=0
η
t
−→ dt [p(t )ẋ(t ) − H(p(t ), x(t ))] . (1.34)
n→∞

ti

With this result we rewrite (1.33) in an abbreviated, symbolic form


t
  i
h̄ dt [p(t )ẋ(t )−H(p(t ),x(t ))]
K (x, t; xi , ti ) = Dx Dp e ti
, (1.35)
 
where Dx stands for dxk and Dp for dpl /(2πh̄). The integrals here are
limits of n-dimensional integrals over x and p for n → ∞, they are integrals
over all functions (paths) x(t) and p(t) and are defined by (1.33).
Equation (1.35) represents an important result. It allows us to calculate
the propagator and thus the solution of the Schrödinger equation in terms of
a path integral over classical functions.

1.3 Quadratic Hamiltonians


Even though the propagator (1.35) looks like a path integral over an expo-
nential function of the action, this is in general not the case, because

pẋ − H(p, x) = L(x, ẋ, p) (1.36)

is not equal to the classical Lagrange function since p is not the canonical mo-
mentum, as already stressed above. Therefore, in general one cannot express
the path integral (1.35) in terms of the action.
Such a simplification, however, is possible for a special p-dependence of
the Hamiltonian. If H depends at most quadratically on p, then the path
integration over the momentum p can be performed and the action appears
in the exponent. This will be discussed in the next 2 sections.

1.3.1 Cartesian Metric

In the last section we have made the special ansatz H = T (p)+V (x) in which
the momenta and coordinates are separated. For the special case, in which
H depends only quadratically on p with constant coefficient, e.g.

p2
H= + V (x) , (1.37)
2m
we can further simplify the path integral (1.33)
10 1 The Path Integral in Quantum Theory
 
n   n
dpl
K (x, t; xi , ti ) = lim dxk (1.38)
n→∞ 2πh̄
k=1 l=0
⎧  ⎫
⎨i  n
xj+1 − xj p2j ⎬
× exp η pj − − V (xj ) .
⎩ h̄ η 2m ⎭
j=0

Using the integral relation


 +∞ 
π b2 +c
e−ap
2
+bp+c
dp = e 4a (1.39)
−∞ a

for Gaussian integrals, discussed in more detail in App. B.2.1, we obtain by


performing the p-integration
  n+1
m 2
K (x, t; xi , ti ) = lim (1.40)
n→∞ 2πh̄iη
⎧   ⎫
 
n ⎨i  n 2 ⎬
m xj+1 − xj
× dxk exp η − V (xj ) .
⎩ h̄ 2 η ⎭
k=1 j=0

Thus in this special case (H = p2 /2m + V ) the propagator K is given (again


in abbreviated notation) by the so-called Lagrangian path integral
t
 i
h̄ L(x,ẋ)dt 
Dx e h̄ S[x(t)] ,
i
K (x, t; xi , ti ) = N Dx e ti
=N (1.41)

with the Lagrangian


m 2
L(x, ẋ) = ẋ − V (x) (1.42)
2
and the action  t
S[x(t)] = L(x(t ), ẋ(t )) dt . (1.43)
ti

There is a problem with the factor in front of the integral in (1.40)


  n+1
m 2
N= . (1.44)
2πh̄iη

The factor N is complex and becomes infinite for n → ∞, η → 0. We will see,


however, later, for example in Sect. 2.1, that the whole path integral leads to
a well-defined expression. In addition, this problem will be bypassed in later
developments where we show that only a normalized propagator, in which N
has been removed, is physically relevant.
The propagator K has thus been reduced to a one-dimensional path inte-
gral, which is only possible for Hamiltonians which are quadratic in p. This
1.3 Quadratic Hamiltonians 11

is a quite important result that we will use throughout all of the following
sections. Equation (1.41) shows that the propagator is given by the phase
exp h̄i S[x(t)] summed over all possible trajectories x(t) with fixed starting
and end points.
At this point we should realize that in going from (1.38) to (1.40) we have
integrated an oscillatory integrand (eif (p) ) over an infinite interval. This was
only possible by a mathematical trick: in applying the Gaussian integration
formula (B.18) we have in effect used the quantity iη in (1.38) as if it were
real. In other words: we have analytically continued expression (1.38) into the
complex plane by setting the time interval η → −iη  with η  real and positive.
This amounts to setting the time t → −it and, correspondingly, ẋ2 → −ẋ2 .
Then the action goes over into the Euclidean action.
 t 
m 2
SE = ẋ + V (x) dt (1.45)
ti 2
and the propagator becomes
 t
Dx e− h̄ SE .
1
KE (x, t; xi , ti ) = N (1.46)
ti

Thus (1.38) has now become a well-behaved integral. Equation (1.46) is rem-
iniscent of the partition function in statistical mechanics which is obtained
by summing the Boltzmann factor exp (−En /T ) over all possible states of
the system.
After performing the integration we have then effectively gone back to the
original definition of time. This analytical continuation is in general possible
only if no singularities are encountered while going to the real time vari-
able t. Also, one has to worry about phase ambiguities connected with the
appearance of the square root of a complex number.

1.3.2 Non-Cartesian Metric


The momentum integration is even possible for Hamiltonians of a more gen-
eral form. As an example we consider
1
H= f1 (x)p2 + f2 (x)p + f3 (x) . (1.47)
2m
Here a problem arises because the canonical quantization of such a Hamilto-
nian is ambiguous. This is so because the classical coordinates and momenta
commute, so that H can be brought into various forms that are classically
all equal, but differ after quantization because the operators p̂ and x̂ do not
commute. We, therefore, turn the question around and ask:
Given a path integral
t 
  i
h̄ dt (pẋ−H(p,x))
Dx Dp e ti (1.48)
12 1 The Path Integral in Quantum Theory

with the classical Hamiltonian (1.47), can this still be identified as a propa-
gator and, if so, for which Hamiltonian?
The answer to this question is given here without proof1
t
  i
h̄ dt [pẋ−H(p,x)]
i
Dx Dp e ti
= x |e− h̄ (t−ti )ĤW |x . (1.49)

Here ĤW is the “Weyl-ordered” Hamiltonian


1 1 2 
ĤW = p̂ f1 (x) + 2p̂f1 (x)p̂ + f1 (x)p̂2
2m 4
1
+ [p̂f2 (x) + f2 (x)p̂] + f3 (x) . (1.50)
2
In this form the momentum- and coordinate-dependent terms are symme-
trized. Note that the path integral in (1.49) is well determined because all
quantities on the lhs of this equation are classical, commuting quantities.
Explicitly the lhs of (1.49) reads
 
n n
dpl
K (x, t; xi , ti ) = lim dxk (1.51)
n→∞ 2πh̄
k=1 l=0
⎧   ⎫
⎨i 
n
p2j ⎬
× exp pj (xj+1 − xj ) − η f1 (xj ) + f2 (xj )pj + f3 (xj ) .
⎩ h̄ 2m ⎭
j=0

It can again be simplified by integrating over the momenta, as in the last


section. Using the Gaussian integral formula (1.39) again this gives
  n+1  n n
m 2
1
K (x, t; xi , ti ) = lim dxk  (1.52)
n→∞ 2πh̄iη f1 (xl )
k=1 l=0
⎧ ⎛  2
⎞⎫

⎨i  n (xj+1 −xj )
− f2 (xj ) ⎪

⎜m η ⎟
× exp η⎝ − ηf3 (xj )⎠ .

⎩ h̄ j=0 2 f1 (xj ) ⎪

The exponent (in round brackets) is


 2
xj+1 −xj
m η − f2 (xj )
(. . .) = − f3 (xj ) (1.53)
2 f1 (xj )
−→ m ẋ − 2ẋf2 (x) + f2 (x) − f (x) .
2 2
η→0 3
2 f1 (x)

1
The proof can be found in [2] and [3].
1.4 Classical Interpretation 13

This last expression is just the Lagrangian


m
L = g1 (x)ẋ2 + g2 (x)ẋ + g3 (x) (1.54)
2
(a kinetic term of this form appears, for example, when rewriting the kinetic
energy of a particle from cartesian into polar coordinates). For this L the
corresponding classical Hamiltonian is given by
m
H = pẋ − L = (mg1 ẋ + g2 ) ẋ − g1 ẋ2 − g2 ẋ − g3
2
 2
m m p − g 2
= g1 ẋ2 − g3 = g1 − g3
2 2 mg1
1 1 g2 (x) 1 g22 (x)
= p2 − p+ − g3 (x) . (1.55)
2m g1 (x) mg1 (x) 2m g1 (x)
With
1 g2 (x) g22 (x)
f1 (x) = , f2 (x) = − , f3 (x) = − g3 (x) (1.56)
g1 (x) mg1 (x) 2mg1 (x)
this is just the Hamiltonian (1.47) that we started out with.
We thus have for the complete propagator in this case
K (x, t; xi , ti ) (1.57)
⎡ ⎤
  n+1  n  n   n
m 2
i
= lim dxk g1 (xl ) exp ⎣ η L(xj , ẋj )⎦
n→∞ 2πh̄iη h̄ j=1
k=1 l=0
⎡ ⎤
  n n  
i 1 h̄
∼ dxk exp ⎣ η L(xj , ẋj ) + ln g1 (xj ) ⎦ (1.58)
h̄ j=1 2 iη
k=1

Thus, in this case the path integral is changed. The square root of a func-
tion that determines the metric of the system appearsin the integrand. Its
presence is easy to understand in terms of a rescaling g1 (x) x → x.

1.4 Classical Interpretation


The simple form (1.41) for the path integral allows a very physical interpre-
tation of the connection between classical and quantum mechanics. For the
classical path the variation of the action is, according to Hamilton’s principle,
equal to zero, i.e. the action is stationary2
t t
δS = L (xcl + δx, ẋ + δ ẋcl ) dt − L (xcl , ẋcl ) dt = 0 . (1.59)
ti ti
2
For an explanation of functionals and their derivatives see App. B.
14 1 The Path Integral in Quantum Theory

This implies that all the paths close to the classical path xcl (t) give about
equal contributions to the path integral. For each path (C1 ) somewhat more
removed from the classical one there will also be another one, (C2 ), whose
action differs from that on C1 just by πh̄. Then we have

e h̄ S(C1 ) = e h̄ S(C2 )+iπ = −e h̄ S(C2 ) ,


i i i
(1.60)

so that the contributions from these two paths cancel each other. Sizeable
contributions to the path integral thus come from paths close to the classical
one. Quantum mechanics then describes the fluctuations of the action in a
narrow range around the classical path.
This observation forms the basis for a semiclassical approximation. This
can be formulated by expanding the action functional S[x(t), ẋ(t)] in terms
of fluctuations δx around the classical path xcl (t). This gives
  2 
1 δ L 2 δ2 L δ2L 2
S[x, ẋ] = Scl + (δx) + 2 δx δ ẋ + (δ ẋ) + .....
2 δx2 δxδ ẋ δ ẋ2
≡ Scl + δ 2 S + . . . . (1.61)

Here all the derivatives have to be taken at the classical path. Because S is
stationary at the classical path, there is no first derivative in this equation.
The propagator (1.41) now becomes
  ( )
1i 2
K (x, t; xi , ti ) = N Dx e h̄ S = N e h̄ Scl Dx exp
i i
δ S + .... . (1.62)
2 h̄

Note that this result (without higher order terms) is exact for Lagrangians
that depend at most quadratically on x and ẋ. The second factor gives the
effects of quantum mechanical fluctuations around the classical path.
An interesting observation on the character of these fluctuations can be
made for the free case (V = 0). The main contribution to the integrand in
(1.40) for η → 0 comes from exponents
 2
η m xj+1 − xj
≈ 1, (1.63)
h̄ 2 η

i.e. from velocities vj ≈ 2h̄/(ηm) which diverge with η → 0. This implies

that the average displacement d within a time-step η is proportional to η,
so that d ∼ η (not d = v dt = v η !), just as for a random walk. The main
2 2 2 2 2 2

contribution to the path integral, and therefore to the quantum mechanical


fluctuation around the classical trajectory, thus comes from paths that are
continuous, but have no finite derivative.
2 Perturbation Theory

In this chapter we discuss first how to calculate the propagator of a free


particle and derive its analytic form. In most cases, however, with a potential
included the exact propagator cannot be calculated in closed form. Thus one
has to resort to perturbation theory which will also be developed in this
chapter.

2.1 Free Propagator


We start with the free propagator K0 , given by

K (x, t; xi , ti )0 = N Dx e h̄ S0
i

⎡ ⎤
  n+1 
+∞ n
 n  2
m 2
i m xj+1 − xj
= lim dxk exp ⎣ η ⎦ . (2.1)
n→∞ 2πh̄iη h̄ j=0 2 η
−∞ k=1

This path integral can be performed exactly. With (B.19) we obtain for the
free propagator
⎛ ⎞ 12
  n+1 n n
m ⎝
2
i π
K0 = lim  n ⎠ (2.2)
n→∞ 2πh̄iη m
(n + 1) 2h̄η
 
i m 2
× exp (x − xi ) ,
n + 1 2h̄η

since xn+1 = x, x0 = xi . With (n + 1)η = t − ti this becomes


 i m(x−xi )
2 
m m i mΔx2
K0 (Δx, Δt) = e h̄ 2(t−ti ) = e h̄ 2Δt (2.3)
2πh̄i (t − ti ) 2πh̄iΔt

with Δx = x−xi , Δt = t−ti . This is the propagator of a free particle for Δt ≥


0; for Δt < 0 it has to be supplemented by the condition K = 0. Because of
Galilei invariance and time homogeneity the free particle propagator depends
only on the space- and time-distances.

U. Mosel, Path Integrals in Field Theory


© Springer-Verlag Berlin Heidelberg 2004
16 2 Perturbation Theory

The last step, from (2.2) to (2.3), shows nicely how in the limit n → ∞
the infinite normalization factor combines with another equally ill defined
factor from the path integral to give a well defined product in (2.3).
Since a free particle has conserved momentum, it is advantageous to trans-
form K0 into the momentum representation

e− h̄ pΔx K0 (Δx, Δt) dΔx
1 i
K0 (p, Δt) = √
2πh̄
 
i m
e− h̄ pΔx e h̄ 2Δt Δx dΔx .
1 m i 2
= √ (2.4)
2πh̄ 2πh̄iΔt
We can now use the integral relation (B.18) in the form


+∞ 
π b2
e−aΔx +bΔx dΔx =
2
e 4a (2.5)
a
−∞

(with a = −im/(2h̄Δt) and b = −ip/h̄) to write


 
2πh̄Δt − i p Δt
2
1 m
K0 (p, Δt) = e h̄ 2m
2πh̄ iΔt −im
p 2

e− h̄ 2m Δt ,
1 i
= √ (2.6)
2πh̄
so that we obtain

1 i
K0 (x, t; xi , ti ) = √ e h̄ pΔx K0 (p, Δt) dp
2πh̄
 
 i p2
1 h̄ pΔx − 2m Δt
= e dp . (2.7)
2πh̄

Equation (2.7) is just the Fourier representation of the propagator (2.3).


The tacit boundary condition K = 0 for Δt < 0 can now explicitly be
taken into account by multiplying (2.7) with the stepfunction Θ(Δt). This
step function can be rewritten using the relation


+∞
1 eiωΔt
Θ(Δt) = dω (ε > 0) , (2.8)
2πi ω − iε
−∞

which follows directly from the residue theorem: for Δt > 0 the integral can
be closed in the upper half of the complex ω plane; Cauchy’s integral theorem
then gives 2πi in the limit ε → 0 so that Θ(t) = 1. If Δt < 0, on the other
hand, then the loop integration can only be closed in the lower half-plane
thus missing the pole at ω = +iε). Multiplying (2.7) with (2.8) gives
2.2 Perturbative Expansion 17
  
p2

+∞ i
pΔx − − h̄ω Δt
h̄ dp dω e h̄ 2m
K0 (x, t; xi , ti ) = . (2.9)
i 2πh̄ 2πh̄ ω − iε
−∞

The energy of a physical, free particle is linked to its momentum by the free
dispersion relation E = p2 /(2m); it is said to be on the energy shell . In order
to exhibit explicitly the deviation of the energy of a quantum-mechanical
particle from its on-shell value in the integral we now substitute

p2
E= − h̄ω (2.10)
2m
and obtain

dp dE i (pΔx−EΔt) ih̄
K0 (x, t; xi , ti ) = e h̄ p2
(2.11)
2πh̄ 2πh̄ E − 2m + iε

(the “−” sign coming from the substitution is cancelled by another sign ob-
tained by inverting the integration boundaries).
Since we will need these expressions later on in three space dimensions
we give them here in a straightforward generalization of (2.7) and (2.11)
 
 i  p2 
  h̄ p · (x − x) − 2m (t − t) d3p
K0 (x , t ; x, t) = e Θ(t − t) , (2.12)
(2πh̄)3
and

d3p dE i (p · Δx − EΔt) ih̄
K0 (x, t; xi , ti ) = 3
e h̄ p2
. (2.13)
(2πh̄) 2πh̄ E − 2m + iε

The integrand on the right-hand side of (2.11) and (2.13) is the free propa-
gator in the energy-momentum representation. Since p and E are independent
variables in the integral, we see that propagation also takes place at energies
E = p2 /2m. The classical dispersion relation does show up as a pole in the
propagator.

2.2 Perturbative Expansion


We now assume that the unperturbed particle moves freely and that the
perturbing interaction is given by V (x, t). We furthermore assume that H
has the special form H = p2 /2m + V (x, t), so that the propagator is given by

K(xf , tf ; xi , ti ) = N Dx e h̄ S
i
(2.14)

with
18 2 Perturbation Theory
 tf  tf m 
S= L(x, ẋ) dt = ẋ2 − V (x, t) dt . (2.15)
ti ti 2
Since the integrand here is a classical function, we have
 tf m 2 t
i
h̄ S
i
h̄ t 2 ẋ dt − h̄i tif V (x, t) dt
e =e i e . (2.16)

The second factor can now be expanded in powers of the potential, thus
yielding a perturbative expansion of the action
 tf  tf  tf 2
i
h̄ ti
V (x,t) dt
∼ i 1 1
e = 1− V (x, t) dt− V (x, t) dt +. . . . (2.17)
h̄ ti 2! h̄2 ti

When we substitute this expansion into the expression (2.14) we obtain


 *
Dx e h̄ S0
i
K(xf , tf ; xi , ti ) = N
   2 +
tf tf
i 1 1
× 1− V (x, t) dt − V (x, t) dt + ...
h̄ ti 2! h̄2 ti

= K0 + K1 + K2 + . . . , (2.18)

with S0 being the action of the free particle. Equation (2.18) represents a
sum of path integrals ordered in powers of the interaction.

First order propagator. We next determine the first-order propagator K1 .


According to (2.18) it is given by

 tf
i
h̄ S0
i
K1 (xf , tf ; xi , ti ) = − N Dx e V (x, t) dt

ti
  n+1
i m 2
=− lim (2.19)
h̄ n→∞ 2πh̄iη
⎛ ⎞

+∞ n
 n
 n
m
V (xk , tk ) η exp ⎝i (xj+1 − xj ) ⎠ .
2
× dxi
2h̄η j=0
−∞ i=1 k=1

Here the time-integral has been written as a sum. In a next step we now split
the sum in the exponent into two pieces, one running from j = 0 to j = k − 1
and the other from j = k to j = n and separate the corresponding integrals.
This gives
2.2 Perturbative Expansion 19

n 
i ⎢
K1 (xf , tf ; xi , ti ) = − lim η dxk ⎢
⎣V (xk , tk )
h̄ n→∞
k=1

⎛ ⎞
-
k−1
 m
i 2h̄η
2
(xj+1 − xj ) ⎟
⎜ k
×⎜
⎝N
2 dx1 dx2 · · · dxk−1 e j=0 ⎟
⎠ (2.20)

⎛ ⎞⎤
-
n
 m
i 2h̄η
2
(xj+1 − xj ) ⎟⎥
⎜ n−k+1
×⎜
⎝N
2 dxk+1 · · · dxn e j=k ⎟⎥ .
⎠⎦

The term in the first round bracket is nothing else than the propagator from
ti to tk (K0 (xk , tk ; xi , ti )), and that in the second bracket is that from tk to
t (K0 (x, t; xk , tk )). Thus we have
K1 (xf , tf ; xi , ti ) =

+∞ tf
i
− dx dt K0 (xf tf ; x, t) V (x, t)K0 (x, t; xi , ti ) . (2.21)

−∞ ti

The time integral over the interval from ti to tf can be extended to ∞ by


noting that
K0 (x, t; xi , ti ) = 0 for t < ti (2.22)
K0 (xf , tf ; x, t) = 0 for tf < t .
This gives
K1 (xf , tf ; xi , ti ) =

+∞ 
+∞
i
− dx dt K0 (xf , tf ; x, t) V (x, t)K0 (x, t; xi , ti ) . (2.23)

−∞ −∞

Higher order propagators. Similarly, the higher order terms in the perturba-
tion expansion (2.18) can be obtained. This yields finally
K (xf , tf ; xi , ti ) =

i
K0 (xf , tf ) − K0 (xf , tf ; x1 , t1 ) V (x1 , t1 ) K0 (x1 , t1 ; xi , ti ) dx1 dt1

 
1
− 2 K0 (xf , tf ; x1 , t1 ) V (x1 , t1 ) K0 (x1 , t1 ; x2 , t2 )

× V (x2 , t2 ) K0 (x2 , t2 ; xi , ti ) dx1 dt1 dx2 dt2
+ ··· . (2.24)
20 2 Perturbation Theory

Note that the time-integrations in (2.24) are effectively time-ordered because


of the implicit Θ (tf − ti ) function in each of the propagators, leading to
tf > t1 > t2 > · · · > ti . This fact explains why, for example, the factor 1/(2!)
in the second order term of (2.18) no longer appears in (2.24)
 2 
1 1
dtV (x, t) = V (x, t)V (x, t ) dt dt
2! 2!

1
= [V (x, t)V (x, t )Θ (t − t) + V (x, t)V (x, t )Θ(t − t )] dt dt
2!

= V (x, t )Θ (t − t) V (x, t) dt dt . (2.25)

The last line follows from a simple change of variables; the Θ function in it
is absorbed into the propagator in (2.24). A similar argument holds for all
the higher-order terms in the interaction. In each case the prefactor (1/n!)
stemming from the expansion of the exponential in (2.17) is cancelled by the
time ordering inherent in the propagators.
Equation (2.24) is the Born series for the propagator; it can be repre-
sented graphically as shown in Fig. 2.1. A time axis runs here from bottom
to top. The straight lines denote the free propagation of the particles (K0 )
and the dots stand for the interaction vertices (−iV /h̄) (the circles mark the
interaction range) and a space and time integration appears at each vertex.

xf tf

x2 t2
= + + + ...
x1 t1 x1 t1

xi ti

Fig. 2.1. Born series for the propagator

Bethe–Salpeter equation. The expansion (2.24) can be formally summed. This


can be seen as follows (in the obvious short-term notation)
2.2 Perturbative Expansion 21

K = K0 + K0 U K0 + K0 U K0 U K0 + · · · (2.26)
= K0 + K0 U (K0 + K0 U K0 + · · ·) with U = − h̄i V .

The expression in parentheses is just again K, so that we obtain the Bethe–


Salpeter equation
K = K0 + K0 U K . (2.27)
The Bethe–Salpeter equation is an integral equation for the full interacting
propagator K as can be seen most easily from its space-time representation

K (xf , tf ; xi , ti ) = K0 (xf , tf ; xi , ti ) (2.28)



i
− K0 (xf , tf ; x, t) V (x, t)K (x, t; xi , ti ) dx dt .

We can also represent the Bethe–Salpeter equation in a diagrammatic
way. If we denote the so-called “dressed propagator” K, that includes all the
effects of the interactions, by a double line

= K (x2 , t2 ; x1 , t1 ) , (2.29)
then this equation can be graphically represented as shown in Fig. 2.2. Each
graph in Fig. 2.2 finds a one-to-one correspondence in (2.27): the single lines
represent the free propagator, the double line the dressed propagator and
the dot stands for the interaction U . Comparison with (2.27) shows that the
factors in the second term of this equation have to be written from left to
right against the time-arrow in Fig. 2.2.
The Bethe-Salpeter equation can also be written in an equivalent form
for the interacting wavefunction

= +

Fig. 2.2. Bethe-Salpeter equation (2.27). The arrows indicate the time direction.
The double line denotes the dressed propagator (2.2), the single line the free prop-
agator and the dot the interaction vertex U = − h̄i V
22 2 Perturbation Theory

Ψ (xf , tf ) = K (xf , tf ; xi , ti ) Ψ (xi , ti ) dxi

= K0 (xf , tf ; xi , ti ) Ψ (xi , ti ) dxi

i
− K0 (xf , tf ; x, t) V (x, t)K (x, t; xi , ti ) Ψ (xi , ti ) dxi dx dt


= K0 (xf , tf ; xi , ti ) Ψ (xi , ti ) dxi (2.30)

i
− K0 (xf , tf ; x, t) V (x, t)Ψ (x, t) dx dt .

This constitutes an integral equation for the unknown wavefunction Ψ (x, t).

2.3 Application to Scattering


Let us now apply the results of the last section to a scattering process. In this
case the particle is free at t = −∞, then undergoes the scattering interaction
and then, at t = +∞, is free again.
We treat this problem as usual by adiabatically switching on and off the
interaction V (x, t). The initial condition (for t → −∞) for the wavefunction
then is
Ψin (x, t) = N ei(ki ·x−ωi t) . (2.31)
We choose here
√ a box normalization with periodic boundary conditions so
that N = 1/ V . The scattering state that evolves from this incoming state
is denoted by Ψ (+) (x, t). The superscript (+) indicates that the state evolves
forward in time, starting from Ψin at t = −∞; it thus fulfills the boundary
condition
Ψ (+) (x, t → −∞) = Ψin (x, t) . (2.32)
In a scattering experiment one looks at t → +∞ for a free scattered particle
with definite momentum; the corresponding final state is denoted by

Ψout (x, t) = N ei(kf ·x−ωf t) . (2.33)

The probability amplitude for the presence of Ψout in the scattered state Ψ (+)
is given by


Sf i = Ψout (xf , tf ) Ψ (+) (xf , tf ) d3xf for tf → ∞ . (2.34)

This is just the transition amplitude from the initial state i to the final state
f (the S-matrix).
Expansion (2.30) yields the Bethe–Salpeter equation for the scattering
wavefunction
2.3 Application to Scattering 23

Ψ (+) (xf , tf ) = K0 (xf , tf ; xi , ti ) Ψin (xi , ti ) d3xi (2.35)

i
− K0 (xf , tf ; x, t) V (x, t)K(x, t; xi , ti )Ψin (xi , ti ) d3xi d3x dt .

Inserting this into (2.34) gives for the S-matrix


Sf i = Ψout (xf , tf ) K0 (xf , tf ; xi , ti ) Ψin (xi , ti ) d3xi d3xf (2.36)

i ∗
− Ψout (xf , tf ) K0 (xf , tf ; x, t) V (x, t)

× K(x, t; xi , ti )Ψin (xi , ti ) d3xi d3x dt d3xf .

Since Ψin is a plane wave, we know that



φ (xf , tf ) = K0 (xf , tf ; xi , ti ) Ψin (xi , ti ) d3xi (2.37)

is also a plane wave state, since K0 is the free propagator so that no interac-
tion takes place. φ is actually the same wavefunction as Ψin , only taken at a
later time and space point

φ (xf , tf ) = N ei(ki ·xf −ωi tf ) . (2.38)

This means that the first integral in (2.36) can be easily evaluated


Ψout (xf , tf ) φ (xf , tf ) d3xf = δ 3 (kf − ki ) . (2.39)

We thus have

i
Sf i = δ 3 (kf − ki ) − d3xf d3x d3xi dt (2.40)

/ ∗ 0
× Ψout (xf , tf ) K0 (xf , tf ; x, t) V (x, t)K(x, t; xi , ti )Ψin (xi , ti ) .

The amplitude for a scattering process is given by the second term.


If K on the rhs of (2.40) is now represented by the Born series expan-
sion of the Bethe–Salpeter equation (2.26) this amplitude can be graphically
represented as shown in Fig. 2.3. This diagram can be translated into the
amplitude just given by writing for each straight-line piece

= K0 (x2 , t2 ; x1 , t1 ) (2.41)
xt t1 x2 t2
and for each interaction vertex
i
= − V (x, t) + integration over x, t . (2.42)
xt h̄
24 2 Perturbation Theory

xf tf

xt

xi ti

Fig. 2.3. First order scattering diagram, corresponding to the second term in (2.40)
with K = K0 . Time runs from left to right


The rules are completed by multiplying Ψin and Ψout at the corresponding
sides of the diagram and then integrating over the spatial variables of these
wavefunctions and over all intermediate times. The ordering of factors is such
that in writing the various factors from left to right one goes against the flow
of time in the figure.
These rules are illustrated for a second-order scattering process in Fig. 2.4.
In this figure the time runs from ti to tf from left to right. The corresponding
amplitude is given by
 (

A(2) = d3xi d3xf d3x1 d3x2 dt1 dt2 Ψout (xf , tf )
 
i
× K0 (xf , tf ; x2 , t2 ) − V (x2 , t2 ) K0 (x2 , t2 ; x1 , t1 )

  )
i
× − V (x1 , t1 ) K0 (x1 , t1 ; xi , ti ) Ψin (xi , ti ) . (2.43)

xf tf

x1 t1

x2 t2

xi ti

Fig. 2.4. Second order scattering diagram, corresponding to (2.43)


2.3 Application to Scattering 25

We now evaluate the first order scattering amplitude (cf. (2.40))


 
i ∗
A =−
(1)
Ψout (xf , tf ) K0 (xf , tf ; x, t) V (x, t)K0 (x, t; xi , ti )

× Ψin (xi , ti ) d3x dt d3xi d3xf (2.44)

somewhat further by using expression (2.6) for the free propagator (extended
to three dimensions)
 
 i  p2 
  1 h̄ p·(x −x)− 2m (t −t)
K0 (x , t ; x, t) = e d3p Θ(t − t) . (2.45)
(2πh̄)3

Inserting this into the expression for A(1) gives


 (
i 1 1
e− h̄ (pf ·xf −Ef tf )
i
A(1) = − d3
x dt d 3
xi d3
xf d3
p d 3
q
h̄ V (2πh̄)6
   
p 2 q 2
i
h̄ p·(xf −x)− 2m (tf −t) i
h̄ q·(x−xi )− 2m (t−ti )
×e V (x, t)e
)
h̄ (pi ·xi −Ei ti )
i
×e ,

where the time-ordering tf > t > ti is understood and we have abbreviated


Ei = p2i /(2m) and Ef = p2f /(2m).
We integrate first over xi and xf . This yields

d3xi −→ (2πh̄)3 δ 3 (pi − q) (2.46)

d3xf −→ (2πh̄)3 δ 3 (pf − p) .

Performing next the integrals over p and q gives for the amplitude

i
A(1) = − d3x dt
h̄V
  p2
  2
 
f p
i
−pf ·x− 2m (tf −t) i i (t−t )
pi ·x− 2m i
+ h̄i Ef tf h̄ h̄ − h̄i Ei ti
× e e V (x, t)e e

i
e h̄ (−pf ·x+Ef t) V (x, t)e h̄ (+pi ·x−Ei t) d3x dt .
i i
=− (2.47)
h̄V

The time-integration is performed next. For that we assume that V (x, t)


acts only during a finite, but long time-interval [−T, T ], in which it is time-
independent; at the boundaries of the interval it is adiabatically, i.e. without
any significant energy-transfer, turned on and off. We then obtain
26 2 Perturbation Theory

 T
T →∞
h̄ (Ef −Ei )t V (x)ei(ωf −ωi )t dt −→ V (x) 2πδ (ωf − ωi ) ,
i
dt V (x, t)e dt =
−T
(2.48)
with ωf − ωi = (Ef − Ei )/h̄. Thus A(1) becomes

i
2πδ (ωf − ωi ) e h̄ (pi −pf )·x V (x) d3x .
i
A(1) = − (2.49)
h̄V
This result is the well-known lowest-order Born-approximation.
3 Generating Functionals

In this section we consider the transition amplitude in the presence of an


external “source” J(t), so that the Hamiltonian reads

HJ (x, p) = H(x, p) − h̄J(t)x . (3.1)

A classical example is that of a harmonic oscillator with externally driven


equilibrium position x0 (t). Its Hamiltonian is

p2 1
Hx0 = + k(x − x0 (t))2
2m 2
= H − kx0 (t)x + O(x20 ) . (3.2)

With h̄J(t) = kx0 (t) and for small x0 we just have the form of (3.1). It is
evident that the states of this system will change as time develops, because
of its changing equilibrium position.
Suppose now that the system was in its groundstate at t → −∞ and
that x0 (t) acts only for a limited time period. We could then calculate the
transition probability for the system to be still in its groundstate at t → +∞
by using the techniques developed in sections 2.2 and 2.3. There we saw (cf.
(2.40) and (2.43)) that this probability is determined by matrix elements
of time-ordered products of the interaction, i.e. of the operator x̂ in the
present case. In this chapter we will show that these matrix elements can all
be generated once only the functional dependence of the probability for the
system to remain in its groundstate on an external source is known.

3.1 Groundstate-to-Groundstate Transitions


Generalizing the special example of the introduction we assume that an ar-
bitrary physical system is at first (at ti ) stationary and then changes under
the influence of an external source h̄J(t)x of finite duration. After the source
has been turned off, the groundstate of the system is still the same (up to a
phase), but the system may be excited.

U. Mosel, Path Integrals in Field Theory


© Springer-Verlag Berlin Heidelberg 2004
28 3 Generating Functionals

The propagator for this system, is quite generally, given by (cf. (1.16)

tf
  i
h̄ [pẋ−H(x,p)+h̄J(t)x] dt
xf , tf |xi , ti J = Dx Dp e ti
. (3.3)

The value of this propagator depends obviously on the source function J(t);
it is a functional of the source J(t) as indicated by the index J on the lhs
(see App. B).
In the following paragraphs we will now discuss this functional dependence
for tf → +∞ and ti → −∞. We will also show that it determines the
groundstate expectation values of time-ordered x̂ operators.
We start by calculating the propagator (3.3) of a system under the influ-
ence of the source h̄J(t)x. We first assume that the source is nonzero only
for a limited time between −T and +T

J(t) = 0 for |t| > T . (3.4)

We can then write for the propagator (ti < −T , tf > T )



xf tf |xi ti J = dx dx xf tf |x T x T |x −T J x −T |xi ti  . (3.5)

Note that the two outer propagators are taken with the sourceless Hamil-
tonian (J = 0) because of condition (3.4). They are given by, e.g.,

x −T |xi ti  = x|e− h̄ Ĥ(−T −ti ) |xi 


i


ϕn (x)ϕ∗n (xi ) e− h̄ En (−T −ti ) .
i
= (3.6)
n

Similarly we get

xf tf |x T  = xf |e− h̄ Ĥ(tf −T ) |x 


i


ϕn (xf )ϕ∗n (x )e− h̄ En (tf −T ) .
i
= (3.7)
n

Here the ϕn are eigenstates of the Hamiltonian Ĥ without a source; we assume


that the spectrum of Ĥ is bounded from below with eigenvalues En ≥ E0 .
The state with the lowest eigenvalue E0 is the groundstate of the theory with
its coordinate-space representation ϕ0 (x).
The dependence on ti and tf is now isolated, but now taking the limits
ti → −∞ and tf → +∞ is not straightforward because both times appear as
arguments of oscillatory functions. These functions oscillate the more rapidly
the higher the eigenvalues En are. We can thus expect that the dominant
contribution will come from the lowest eigenvalue E0 .
3.1 Groundstate-to-Groundstate Transitions 29

Wick Rotation. This can indeed be shown by by a mathematical trick, the


so-called Wick rotation. In this method one looks at the propagators (3.6)
and (3.7) as functions of ti and tf , respectively, and continues these variables
analytically from the real axis into the complex plane. Mathematically this
is achieved by replacing the physical Minkowski time t by a complex time τ
t −→ τ = te−iε . (3.8)
The direction of the rotation is mandated by the requirement that there
are no singularities of the integrand encountered in the rotation of the time-
axis. This is the case for
0≤ε<π . (3.9)
For larger angles the rhs of (3.6) and (3.7) develops an essential singularity
for ti/f → −/ + ∞.
This rotation corresponds to an analytic continuation in the variable ε;
the real, four-dimensional Minkowski space can be viewed as a subspace of
a four-dimensional complex space (or of a five-dimensional space with 3 real
space-coordinates and 1 complex time-coordinate). The substitution (3.8)
corresponds to a clockwise rotation of the time-axis out of the real subspace
into the larger complex space. For ε = +π/2 this corresponds to a rotation
t = +∞ → τ = −i∞, t = −∞ → τ = +i∞; just this special case is
also often called a Wick rotation. Physical results are obtained by evaluating
the relevant expressions in the enlarged, complex space and then, at the
end, rotating the result back to the real time-axis (which works only if no
singularities are encountered on this way).
We now apply this rotation to the expression

e− h̄ E0 ti x −T |xi ti  = ϕn (x)ϕ∗n (xi ) e h̄ En T e h̄ (En −E0 )ti
i i i
(3.10)
n

and thus perform the limit ti → −∞ on the rotated time axis

e− h̄ E0 τi x −T |xi τi 
i
lim (3.11)
τi →−∞(cos ε−i sin ε)

ϕn (x)ϕ∗n (xi ) e h̄ En T e+ h̄ (En −E0 )τi .
i i
= lim
τi →−∞(cos ε−i sin ε)
n

Because we have separated out the lowest frequency in the limit the last
factor vanishes for ε > 0 , except for n = 0. Thus we get

e− h̄ E0 τi x −T |xi τi  = ϕ0 (x)ϕ∗0 (xi ) e h̄ E0 T .


i i
lim (3.12)
τi →−∞(cos ε−i sin ε)

Analogously we also obtain

e+ h̄ E0 τf xf τf |x T  = ϕ0 (xf )ϕ∗0 (x ) e h̄ E0 T .


i i
lim (3.13)
τf →+∞(cos ε−i sin ε)

Both expressions, (3.12) and (3.13), are the analytical continuations of the
corresponding limits on the real time axis from ε = 0 to ε > 0. Since the
30 3 Generating Functionals

right hand sides of these equations do not depend on ε they can be continued
back to the real time axis (ε = 0) without any change. By means of the
Wick rotation we have thus been able to make the expression (3.12) and
(3.13) convergent; in this process we have found that only the groundstate
contributes to the transition probability.
We now insert these expressions into (3.5) and get

e h̄ E0 (τf −τi ) xf τf |xi τi J


i
lim (3.14)
τi →−∞(cos ε−i sin ε)
τf →∞(cos ε−i sin ε)

ϕ0 (xf )ϕ∗0 (xi ) e h̄ E0 2T dx dx ϕ∗0 (x ) x T |x −T J ϕ0 (x)
i
=

The integral in the last line can be rewritten, using ϕ0 (x) = x|0. This gives
 
dx dx ϕ0 (x ) x T |x −T J ϕ0 (x) = dx dx 0|x x T |x −T J x|0
 ∗  

= 0T |0 −T J , (3.15)
i.e. the groundstate-groundstate transition amplitude under the influence of
the source h̄Jx. Thus we obtain for (3.14)

e h̄ E0 (τf −τi ) xf τf |xi τi J


i
lim
τi →−∞(cos ε−i sin ε)
τf →∞(cos ε−i sin ε)

= ϕ0 (xf )ϕ∗0 (xi ) e h̄ E0 2T 0T |0 −T J ,


i
(3.16)
and by Wick-rotating back

e h̄ E0 (tf −ti ) e− h̄ E0 2T
i i

0T |0 −T J = lim xf tf |xi ti J . (3.17)


ti →−∞
tf →+∞
ϕ∗0 (xi )ϕ0 (xf )

With (in the limit ti → −∞, tf → +∞)

xf tf |xi ti J=0 = ϕ∗0 (xi )ϕ0 (xf )e− h̄ E0 (tf −ti )
i
(3.18)
(3.17) becomes
xf tf |xi ti J − i E0 2T
0T |0 −T J = lim e h̄ . (3.19)
ti →−∞
tf →+∞
xf tf |xi ti 0

This immediately gives for the free (J = 0) transition amplitude

0T |0 −T 0 = e− h̄ E0 2T ,
i
(3.20)
as it should.
Equation (3.17) implies that the groundstate-to-groundstate (gs-gs) tran-
sition amplitude is – up to a factor – given by a path integral from arbitrary
xi to arbitrary xf and thus does not depend on these quantities, if only the
corresponding times are taken to infinity.
3.1 Groundstate-to-Groundstate Transitions 31

Gs-gs transition amplitude. The gs-gs amplitudes 0T |0 −T J deserve some


explanation. Formally, they are given by
 +T
−i (H−h̄J(t)x)dt
0T |0 −T J = 0|e h̄ −T |0 . (3.21)

The ground state |0 is assumed to be unique, if there is no source J present,


and normalized to 1. It is the groundstate of the
 theory before and afterthe
 +T
action of the source. On the other hand, exp − h̄ −T (H − h̄J(t)x) dt |0
i

is the state that |0 has evolved into at t = +T under the influence of the
external source J(t).
The matrix element (3.21) is the probability amplitude of finding the
original groundstate in the time-evolved one. Since the source-free propaga-
tion just leads to a phase for this amplitude (3.20) we also write the full
probability amplitude in terms of a phase
i
0T |0 −T J = e h̄ (S[J]−E0 2T ) (3.22)

where we have taken out the source-free propagation contribution (3.20). This
phase is the quantity determined by (3.19).
To conclude these considerations we note that instead of using the Wick-
rotation to make the transition rates well behaved for very large times we
could also have added a small negative imaginary term −iεEn to all eigenval-
ues En . This would have given a damping factor to the oscillating exponen-
tials that becomes larger with n and thus would have led to a suppression of
all higher excitations. This becomes apparent by looking at expressions (3.6)
and (3.7). At the end of the calculation the limit ε → 0 would have to be
performed.

3.1.1 Generating Functional

The gs-gs transition rate is a functional of the source J(t) which, for T → +∞,
we denote by

xf tf |xi ti J − i E0 2T
W [J] = 0 +∞|0 −∞J = lim e h̄ . (3.23)
ti →−∞
tf →+∞
xf tf |xi ti J=0
T →∞

W [J] is called a generating functional for reasons that will become clear in
the next section.
In order to get rid of the phase that is already produced by a source-
free propagation (exp(−iE0 2T /h̄)) we define now a normalized generating
functional
W [J] 0 +∞|0 −∞J xf tf |xi ti J
Z[J] = = = lim (3.24)
W [0] 0 +∞|0 −∞J=0 ti →−∞
tf →+∞
xf tf |xi ti J=0
32 3 Generating Functionals

with Z[0] = 1. The functional Z[J] describes the processes relative to the un-
perturbed (J = 0) time-development. The numerator in (3.24) is a transition
amplitude and can therefore be written as a path integral

+∞
  i
h̄ [pẋ−H(p,x)+h̄J(t)x] dt
xf + ∞|xi − ∞J = Dx Dp e −∞
. (3.25)

If H is quadratic in p and of the form H = p2 /(2m)+V (x), the propagator


can be rewritten as (see Sect. 1.3)

+∞
 i
h̄ [L(x,ẋ)+h̄J(t)x] dt
xf + ∞|xi − ∞J = N Dx e −∞
. (3.26)

In the normalized functional Z[J] the (infinite) factor N cancels out because
it is independent of J
i W [J] 0 +∞|0 −∞J
Z[J] = e h̄ (S[J] = =
W [0] 0 +∞|0 −∞J=0

+∞
i
[L(x,ẋ)+h̄J(t)x] dt
 h̄

Dx e −∞
= (3.27)

+∞
i
[L(x,ẋ)] dt
 h̄

Dx e −∞

3.2 Functional Derivatives


of Gs-Gs Transition Amplitudes
In this section we will show – by using the methods outlined in App. B – that
the groundstate expectation value of a time-ordered product of interaction
operators – in this present case of the operators x̂(t) – can be obtained as
functional derivatives of the functional W [J] with respect to J.
We start with the definition of a path integral as a limit of a finite dimen-
sional integral (see (1.33))
 
n  
n
dpl
xf tf |xi ti J = lim dxk (3.28)
n→∞ 2πh̄
k=1 l=0
⎛ ⎞
n
i
× exp ⎝ [pj (xj+1 − xj ) − ηH (pj , xj ) + h̄Jj xj ]⎠ .
h̄ j=0

and calculate its functional derivative with respect to J. In order to become


familiar with functional derivatives we do this in quite some detail. Using the
3.2 Functional Derivatives of Gs-Gs Transition Amplitudes 33

definition (B.26) for the functional derivative we get (with the abbreviation
F (xj , pj ) = pj (xj+1 − xj ) − ηH(pj , x̄j ))
δxf tf |xi ti J
δJ(t1 )
*   n   n
1 dpl
= lim lim dxk (3.29)
ε→0 ε n→∞ 2πh̄
k=1 l=0
⎡ ⎛ ⎞
n
i
× ⎣exp ⎝ {F (xj , pj ) + h̄xj [Jj + εδ(tj − t1 )]}⎠
h̄ j=0
⎛ ⎞⎤⎫
i n ⎬
− exp ⎝ (F (xj , pj ) + h̄xj Jj )⎠⎦
h̄ j=0 ⎭
⎛ ⎞
  n   l n
dpl i
= lim dxk ix1 exp ⎝ [F (xj , pj ) + h̄xj Jj ]⎠ ,
n→∞ 2πh̄ h̄ j=0
k=1 l=0

which can be written as


tf
  i
[pẋ−H(x,p)+h̄J(t)x] dt
δxf tf |xi ti J h̄

=i Dx Dp x(t1 )e ti
. (3.30)
δJ(t1 )
We now want to relate this derivative of a classical functional to quan-
tum mechanical expressions and thus understand its physical significance and
meaning. In order to do so, we go back to the definition of the propagator
(1.16). Equation (1.23) then reads

xf , tf |xi ti  (3.31)



= dx1 . . . dxn xf tf |xn , tn xn tn |xn−1 tn−1  . . . x1 t1 |xi ti  .

Now, as in (3.30) for J = 0, we introduce one factor of x1 on the righthand


side of this equation

dx1 . . . dxn xf tf |xn tn  . . . x1 t1 |xi ti x1

= dx1 . . . dxn xf tf |xn tn  . . . x1 t1 |x̂(t1 )|xi ti  ; (3.32)

the last step is possible because |x1 t1  is an eigenstate of the x̂(t) operator
with eigenvalue x1 (cf. (1.18)). The last integral is obviously equal to

xf tf |x̂(t1 )|xi ti  , (3.33)

i.e. to a matrix element of the position operator. We thus have


34 3 Generating Functionals

tf
1   i
[pẋ−H(x,p)] dt
δxf tf |xi ti J 11 h̄

1 =i Dx Dp x(t1 )e ti

δJ(t1 ) J=0

= ixf tf |x̂(t1 )|xi ti  . (3.34)

For the case that H is separable in x and p and quadratic in p, this relation
reads
tf
1  i
L(x,ẋ) dt
δxf tf |xi ti J 11 h̄

1 = iN Dx x(t1 )e ti

δJ(t1 ) J=0

= ixf tf |x̂(t1 )|xi ti  . (3.35)

The functional derivative of the propagator with respect to the source thus
gives the transition matrix element of the coordinate x̂.
The higher order functional derivatives yield

δ n xf tf |xi ti J
(3.36)
δJ(t1 )δJ(t2 ) . . . δJ(tn )
tf
  i
h̄ [pẋ−H(x,p)+h̄J(t)x] dt
n
= (i) Dx Dp x(t1 )x(t2 ) . . . x(tn ) e ti
.

One might guess that


1
δ n xf tf |xi ti J 1
1 = in xf tf |x̂(t1 )x̂(t2 ) . . . x̂(tn )|xi ti  , (3.37)
δJ(t1 )δJ(t2 ) . . . δJ(tn ) 1J=0

but this equation is not quite correct. We see this by considering explicitly
the second derivative. We can proceed there exactly in the same way as for
the first. We have for the rhs of (3.36) in the case of the second derivative

tf
1   i
[pẋ−H(x,p)] dt
δ xf tf |xi ti J 11
2 h̄

=i 2
Dx Dp x(tα )x(tβ )e ti

δJ(tα )δJ(tβ ) 1 J=0



= i2 dxi . . . dxn xf tf |xn tn  · · · xl tl |x̂(tα )|xl−1 tl−1 

· · · xk tk |x̂(tβ )|xk−1 tk−1  · · · x1 t1 |xi ti  . (3.38)

Here we have assumed that tα > tβ , since each of the infinitesimal Green’s
functions propagates only forward in time. In this case (3.38) is indeed equal
to
i2 xf tf |x̂(tα )x̂(tβ )|xi ti  . (3.39)
3.2 Functional Derivatives of Gs-Gs Transition Amplitudes 35

However, if tα < tβ , then these two times appear in a different ordering on


the rhs of (3.38) and thus of the matrix element (3.39). The two cases can be
combined by introducing the time-ordering operator T

x̂(t1 )x̂(t2 ) t1 > t2
T [x̂(t1 )x̂(t2 )] = (3.40)
x̂(t2 )x̂(t1 ) t2 > t1 .

With the time-ordering operator we have

tf
1   i
[pẋ−H(x,p)] dt
δ xf tf |xi ti J 11
2 h̄

= i2
Dx Dp x(t )x(t )e ti

δJ(t1 )δJ(t2 ) 1J=0


1 2

= i2 xf tf |T [x̂(t1 )x̂(t2 )] |xi ti  . (3.41)

The same reasoning leads to the following result for higher-order deriva-
tives
 n 1
1 δ n xf tf |xi ti J 1
1
i δJ(t1 )J(t2 ) . . . δJ(tn ) 1J=0
tf
  i
h̄ [pẋ−H(x,p)] dt
= Dx Dp x(t1 )x(t2 ) . . . x(tn ) e ti

= xf tf |T [x̂(t1 )x̂(t2 ) . . . x̂(tn )] |xi ti  . (3.42)

This is the generalization of (3.41). We thus have


 n 1
1 δn xf tf |xi ti J 11
i δJ(t1 )δJ(t2 ) . . . δJ(tn ) xf tf |xi ti J=0 1J=0
 n 1
1 δn 1
= Z[J]11
i δJ(t1 )δJ(t2 ) . . . δJ(tn ) J=0
tf
i
[pẋ−H(x,p)] dt
  h̄

Dx Dp x(t1 )x(t2 ) . . . x(tn ) e ti

=
tf
i
[pẋ−H(x,p)] dt
  h̄

Dx Dp e ti

xf tf |T [x̂(t1 )x̂(t2 ) . . . x̂(tn )] |xi ti 


= , (3.43)
xf tf |xi ti 

where the limit ti → −∞, tf → +∞ is understood and all the times t1 , . . . , tn


lie in between these limits. If the Hamiltonian is quadratic in p and separates
in p and x, then we have
36 3 Generating Functionals

Dx x(t1 )x(t2 ) . . . x(tn ) e h̄ S[x(t)]
i
xf tf |T [x̂(t1 )x̂(t2 ) . . . x̂(tn )] |xi ti 
 = ,
xf tf |xi ti  Dx e h̄ S[x(t)]
i

(3.44)
where S[x(t)] is the action that depends functionally on the trajectory x(t).
We now rewrite this equation. The numerator of the lhs becomes (in the
limit ti → −∞, tf → +∞, indicated by the arrow)

xf tf |T [. . .]|xi ti J=0 −→ xf tf |00|T [. . .]|00|xi ti J=0


= 0|T [. . .]|0 ϕ0 (xf )e− h̄ E0 tf ϕ∗0 (xi )e+ h̄ E0 ti ,
i i
(3.45)

while the denominator can be written as (c.f. (3.18))

xf tf |xi ti  = xf |e− h̄ Ĥtf e+ h̄ Ĥti |xi  −→ ϕ0 (xf )e− h̄ E0 tf ϕ∗0 (xi )e+ h̄ E0 ti
i i i i

(3.46)
where we have used in both cases (1.19), inserted a complete set of states
and – through the limit of infinite times – projected out the groundstate. The
gs wavefunctions and the time-dependent exponentials cancel out so that we
obtain (for quadratic Hamiltonians)

Dx x(t1 )x(t2 ) . . . x(tn ) e h̄ S[x(t)]
i

0|T [x̂(t1 )x̂(t2 ) . . . x̂(tn )] |0 = 


Dx e h̄ S[x(t)]
i

 n 1
1 δn 1
= Z[J]11
i δJ(t1 )δJ(t2 ) . . . δJ(tn ) J=0
(3.47)

with  +∞
S[x(t)] = L(x(t), ẋ(t)) dt . (3.48)
−∞

This is a very important result. It shows that the groundstate expectation


value of a time-ordered product of position operators, the so-called correlation
function, can be obtained as a functional derivative of the functional Z[J]
defined in (3.24) and (3.27). Note that |0 is the groundstate of Ĥ, as can be
seen from (3.45). Thus the groundstate appearing on the lhs of (3.47) and
the propagator Z[J] are linked together: if Z[J] contains, for example, only
a free Hamiltonian, then |0 is the groundstate of a free theory. If, on the
other hand, Z[J] contains interactions, then |0 is the groundstate of the full
interacting theory.
Coming back to our example of the driven harmonic oscillator, discussed
at the start of this section, we see that the time-ordered groundstate ex-
pectation values of x̂ are just the matrix elements that would appear in a
time-dependent perturbation theory treatment of the groundstate of this sys-
tem. Thus, if all the correlation functions are known, then the perturbation
series expansion is also known.
Part II

Relativistic Quantum Field Theory


4 Relativistic Fields

In this chapter a few essential facts of classical relativistic field theory are
summarized. It will first be shown how to derive the equations of motion of
a field theory, for example the Maxwell equations of electrodynamics, from a
Lagrangian. Second, the connection between symmetries of the Lagrangian
and conservation laws will be discussed1 . This chapter follows very closely
the presentation in [4] which can also serve as a more expanded, yet still
quite easy to follow introduction to modern field theories of fundamental
interactions.

4.1 Equations of Motion


The equations of motion of classical mechanics can be obtained from a La-
grange function by using Hamilton’s principle that the action for a given
mechanical system is stationary for the physical space–time development of
the system.
The equations of motion for fields that determine their space–time depen-
dence can be obtained in an analogous way by identifying the field amplitudes
at a coordinate x with the dynamical variables (coordinates) of the theory.
Let the functions that describe the fields be denoted by

Φα (x) with xμ = (t, x) , (4.1)

where α labels the various fields appearing in a theory. The fields Φα (x, t)
play the same role as the generalized coordinates qi (t) in classical mechanics;
the analogy here is such that the fields Φα correspond to the coordinates q
and the points x to the classical indices i. The corresponding velocities are
given in a direct analogy by the time derivatives of Φα : ∂t Φα .
The Lagrangian L of the system is expressed as an integral over these
labels in terms of a Lagrange density L

L = L(Φα , ∂μ Φα ) d3 x (4.2)

1
The units, the metric and the notation used from now on are explained in App.
A.

U. Mosel, Path Integrals in Field Theory


© Springer-Verlag Berlin Heidelberg 2004
40 4 Relativistic Fields

where the spatial integration is performed over the volume of the system. The
requirement of Lorentz covariance for the equations of motion for the fields
that we are after requires that L is a Lorentz scalar. This can only be the
case if L not only depends on the time-derivative, but also on the derivatives
with respect to the first three coordinates; this explains the presence of the
four-gradients ∂μ Φα in (4.2).
The action S is then defined as usual by

t1 
S= L dt = L(Φα , ∂μ Φα ) d4 x (4.3)
t0 Ω

with the Lorentz-invariant four-dimensional volume element d4 x = d3 x dt.


The, in general finite, space–time volume of the system is denoted by Ω.
In order to derive the field equations from the action S by Hamilton’s
principle, we now vary the fields and their derivatives

Φα → Φα = Φα + δΦα
∂μ Φα → (∂μ Φα ) = ∂μ Φα + δ(∂μ Φα ) . (4.4)

This yields

δL = L(Φα , (∂μ Φα ) ) − L(Φα , ∂μ Φα )


∂L ∂L
= δΦα + δ(∂μ Φα )
∂Φα ∂(∂μ Φα )
∂L ∂L
= δΦα + ∂μ (δΦα ) . (4.5)
∂Φα ∂(∂μ Φα )

According to the Einstein convention a summation over μ is implicitly con-


tained in this expression. In going from the second to the third line differen-
tiation and variation can be commuted because both are linear operations.
The equations of motion are now obtained from the variational principle
 (   )
∂L ∂L ∂L
δS = − ∂μ δΦα + ∂μ δΦα d4 x
Ω ∂Φα ∂(∂μ Φα ) ∂(∂μ Φα )
=0 (4.6)

for arbitrary variations δΦα under the constraint that

δΦα (t0 ) = δΦα (t1 ) = 0 , (4.7)

where t0 and t1 are the time-like boundaries of the four-volume Ω. The last
term in (4.6) can be converted into a surface integral by using Gauss’s law.
For fields which are localized in space the surface integral over the spatial
surface vanishes if the surface is moved out to infinity; the surface integral
4.1 Equations of Motion 41

over the time-like boundaries vanishes because the fields are not to be varied
at t0 and t1 (cf. (4.7)).
Since the variations δΦα are arbitrary, the condition δS = 0 leads to the
equations of motion
 
∂ ∂L ∂L
− =0. (4.8)
∂xμ ∂(∂μ Φα ) ∂Φα
The relativistic equivalence principle demands that these equations have
the same form in every inertial frame of reference, i.e. that they are Lorentz
covariant. This is only possible if L is a Lorentz scalar, i.e. if it has the same
functional dependence on the fields and their derivatives in each reference
frame.
In a further analogy to classical mechanics, the canonical field momentum
is defined as
∂L ∂L
Πα = = . (4.9)
∂ Φ̇α ∂(∂0 Φα )
From L and Πα the Hamiltonian H is obtained as
 
H = H d3 x = (Πα Φ̇α − L) d3 x . (4.10)

The Hamiltonian H represents the energy of the field configuration.

4.1.1 Examples
The following sections contain examples of classical field theories and their
formulation within the Lagrangian formalism just introduced. We start out
with probably the best-known case of classical electrodynamics, then general-
ize it to a treatment of massive vector fields and then move on to a discussion
of classical Klein–Gordon and Dirac fields that will play a major role in the
later chapters of this book.

Electrodynamics
The best-known classical field theory is probably that of electrodynamics,
in which the Maxwell equations are the equations of motion. The two ho-
mogeneous Maxwell equations allow us to rewrite the fields in terms of a
four-potential
Aμ = (A0 , A) , (4.11)
defined via
B = ∇×A ,
∂A
E = −∇A0 − . (4.12)
∂t
Note that the 2 homogeneous Maxwell equations are now automatically ful-
filled. The two inhomogeneous Maxwell equations2
2
Here the Heaviside units are used with c = 1 and 0 = μ0 = 1.
42 4 Relativistic Fields

∇·E = ρ ,
∂E
∇×B− =j (4.13)
∂t
with external density ρ and current j can be rewritten as
∂Fμν
∂μ Fμν = = jν , (4.14)
∂xμ
with the four-current
j ν = (ρ, j) (4.15)
and the antisymmetric field tensor
∂Aν ∂Aμ
Fμν = − = ∂ μ Aν − ∂ ν A μ . (4.16)
∂xμ ∂xν
The tensor Fμν is the dyadic product of two four–vectors and thus a Lorentz-
tensor. Equation (4.14) is the equation of motion for the field tensor or the
four–vector field Aμ . Current conservation is expressed by the continuity
equation
∂ρ
+ ∇ · j = ∂0 j 0 + ∂i j i = ∂ν j ν = 0 (4.17)
∂t
It is easy to show that (4.14) can be obtained from the Lagrangian
1 1
L = − Fμν Fμν − j ν Aν = − F2 − j · A (4.18)
4 4
by using (4.8); the fields Aν here play the role of the fields Φα in (4.8). We
have
∂L
= −j ν ,
∂Aν
∂L 1
= − 2(+Fμν − Fνμ ) = −Fμν . (4.19)
∂(∂μ Aν ) 4
The last step is possible because F is an antisymmetric tensor. The equation
of motion is therefore
 
∂ ∂L ∂L ∂Fμν
− = − + jν = 0 , (4.20)
∂xμ ∂(∂μ Aν ) ∂Aν ∂xμ

in agreement with (4.14). It is now easy to interpret the two terms in L (4.18):
the first one gives the Lagrangian for the free electromagnetic field, whereas
the second describes the interaction of the field with charges and currents.
The two homogeneous Maxwell equations can also be expressed in terms
of the field tensor by first introducing the dual field tensor
1 μνρσ
F̃μν =  Fρσ ; (4.21)
2
4.1 Equations of Motion 43

here μνρσ is the Levi-Civita antisymmetric tensor which assumes the values
+1 or −1 according to whether (μνρσ) is an even or odd permutation of
(0,1,2,3), and 0 otherwise. In terms of F̃ the homogeneous Maxwell equations
read
∂μ F̃μν = 0 . (4.22)
Equations (4.14) and (4.22) represent the Maxwell equations in a mani-
festly covariant form. The Lagrangian (4.18) is obviously Lorentz-invariant
since it consists of invariant contractions of two Lorentz-tensors (the first
term) and two Lorentz-vectors (the second term). It is also invariant under
a gauge transformation

Aμ (x) −→ Aμ  (x) = Aμ (x) + ∂ μ (x) (4.23)

where (x) is an arbitrary, differentiable function of space-time x. The invari-


ance of L holds because F itself is gauge-invariant by construction and the
contribution of the interaction term to the action is gauge-invariant for an
external conserved current. The same then holds for the equation of motion
(4.20) and the directly observable fields E and B.
The gauge freedom can be used to impose constraints on the four compo-
nents of the vector field Aμ , for example, by requiring the covariant Lorentz
gauge-condition ∂μ Aμ (x) = 0 or the transverse gauge condition ∇ · A = 0.
In addition, for free fields the gauge freedom can be used, for example, to set
the 0th component of the four-potential equal to zero. Thus, a free electro-
magnetic field has only two degrees of freedom left.
Symmetry (Lorentz-invariance), gauge-invariance and simplicity (there
are no higher order terms in (4.18)) thus determine the Lagrangian of elec-
trodynamics.

Massive Vector Fields

Vector fields in which – in contrast to the electromagnetic field – the field


quanta are massive are described by the so-called Proca equation:

∂μ Fμν + m2 Aν = j ν . (4.24)

Operating on this equation with the four-divergence ∂ν gives, because F is


antisymmetric,
m2 ∂ν Aν = ∂ν j ν . (4.25)
For m = 0 and a conserved current (∂ν j ν = 0 (4.17)), this reduces the
equation of motion (4.24) to
 
2 + m2 Aν = j ν with ∂ν Aν = 0 . (4.26)

Thus for massive vector fields the freedom to make gauge transformations
on the vector field is lost. The condition of vanishing four-divergence of the
44 4 Relativistic Fields

field reduces the degrees of freedom of the field from 4 to 3. The space-like
components represent the physical degrees of freedom.
The Lagrangian that leads to (4.24) is given by
1 1
L = − F 2 + m2 A 2 − j · A . (4.27)
4 2

Klein–Gordon Fields

A particularly simple example is provided by the so-called Klein–Gordon field


φ that obeys the equation of motion
   
∂μ ∂ μ + m2 φ(x) = 2 + m2 φ(x) = 0 ; (4.28)

such a field describes scalar particles, i.e. particles without intrinsic spin. The
Klein–Gordon equation is solved by plane waves

φ(x, t) = e±i(ωt−k·x) = e±ikx (4.29)

with the four-vector



k = (ωk , k) with ωk = k 2 + m2 . (4.30)

The Lagrangian leading to (4.28) is given by


1 
L= (∂μ φ) (∂ μ φ) − m2 φ2 (4.31)
2
since we have
∂L
= ∂μφ (4.32)
∂ (∂μ φ)
and
∂L
= −m2 φ . (4.33)
∂φ
It is essential to note here that the Lagrangian density
 (4.31) that leads to
(4.28) is not unique; unique is only the action S = L d4x. For localized
fields for which the surface contributions vanish we can perform a partial
integration of the kinetic term in (4.28)
 
d4x (∂μ φ) (∂ μ φ) = − d4x φ2φ . (4.34)

Thus, taking the alternative Lagrangian


1  
L = − φ 2 + m2 φ (4.35)
2
leaves the action invariant. The Lagrangians (4.35) and (4.31) are equivalent.
Since both give the same action they also lead to the same equation of motion
(4.28).
4.1 Equations of Motion 45

An interesting case occurs if we consider two independent real scalar fields,


φ1 and φ2 , with the same mass m. The total Lagrangian is then simply given
by a sum over the Lagrangians describing the individual fields, i.e.
1  1 
L= (∂μ φ1 ) (∂ μ φ1 ) − m2 φ21 + (∂μ φ2 ) (∂ μ φ2 ) − m2 φ22 . (4.36)
2 2
On the other hand, we can also construct two complex fields from the two
real fields φ1 and φ2 , namely
1
φ = √ (φ1 + iφ2 ) (4.37)
2
and its complex conjugate. In terms of these the Lagrangian (4.36) can be
rewritten to

L = (∂μ φ) (∂ μ φ) − m2 φ∗ φ . (4.38)
The equivalent Lagrangian is then
 
L = −φ∗ 2 + m2 φ . (4.39)

Dirac Fields

A particularly simple example is provided by the Dirac field Ψ for which the
equation of motion is just the Dirac equation

(iγ μ ∂μ − m) Ψ (x) = 0 , (4.40)

where the γμ are the usual (4 × 4) matrices of Dirac theory which obey the
algebra
{γ μ , γ ν } ≡ γ μ γ ν − γ ν γ μ = 2g μν . (4.41)
The field Ψ itself is a 4 × 1 column matrix of four independent fields, a
so-called spinor. For the free field equation (4.40) it can be written in terms
of plane waves
Ψ (x) = w e−ikx (4.42)
where w is a 4 × 1-dimensional c-number spinor. This spinor is often de-
noted by u or v for positive and negative energy-eigenvalues of the Dirac
Hamiltonian in (4.40), respectively
 
u(k, s) u(k, s)
(α · k + βm) = ωk (4.43)
v(−k, s) −v(−k, s)

with ωk = + k2 + m2 , and to helicity
 
k u(k, s) u(k, s)
Σ· = 2s (4.44)
|k| v(−k, s) v(−k, s)
46 4 Relativistic Fields

with s = ±1/2. Here the 4 × 4 spin matrix Σ is built from the normal (2×2)
Pauli spin matrices via  
σ 0
Σ= (4.45)
0 σ
and the matrices α and β are related to the γ matrices by α = γ 0 γ and
β = γ0.
The spinors u and v are normalized as follows

u† (k, s)u(k, r) = δrs ωmk ū(k, s)u(k, r) = δrs

v † (k, s)v(k, r) = δrs ωmk v̄(k, s)v(k, r) = −δrs (4.46)

u† (k, s)v(−k, r) = 0 ū(k, s)v(k, r) = 0 .

The corresponding Lagrangian is given by

L = Ψ̄ (iγ μ ∂μ − m) Ψ , (4.47)

where the spinor Ψ̄ is defined by Ψ̄ = Ψ † γ0 . This can be seen by identifying


the fields Φα in (4.8) with the four components of the Dirac spinor Ψ̄ . Since
L does not depend on ∂μ Ψ̄ the equation of motion is simply given by

∂L
= (iγ μ ∂μ − m) Ψ = 0 . (4.48)
∂ Ψ̄

4.2 Symmetries and Conservation Laws

As in classical mechanics there is also in field theory a conservation law as-


sociated with each continuous symmetry of L. The theorem which describes
the connection between the invariance of the Lagrangian under a continu-
ous symmetry transformation and the related conserved current is known as
Noether’s theorem. In the following, this will be illustrated for different types
of symmetries which then lead to the well-known conservation laws.
The common expression in the arguments to follow is the change of the
Lagrangian density under a change of the fields and their derivatives (see
(4.4)). According to (4.5) and the Lagrange equations of motion (4.8) this
change is given by  
∂L
δL = ∂μ δΦα . (4.49)
∂(∂μ Φα )
4.2 Symmetries and Conservation Laws 47

4.2.1 Geometrical Space–Time Symmetries

In this section we investigate the consequences of translations in four-dimen-


sional space–time, i.e. infinitesimal transformations of the form

xν → xν = xν + ν , (4.50)

where ν is a constant infinitesimal shift of the coordinate xν . Under such


transformations the change of L is given by
∂L
δL = ν = ν ∂ ν L , (4.51)
∂xν
since L is a scalar.
If now L is required to be form-invariant under translations, it does not
explicitly depend on xν . In this case, δL is also given by (4.49). The changes
of the fields Φα appearing there are for the space–time translation considered
here given by
∂Φα
δΦα = ν = ν ∂ ν Φα . (4.52)
∂xν
Inserting (4.52) into (4.49) yields
 
∂L ν
δL = ν ∂μ ∂ Φα . (4.53)
∂(∂μ Φα )
Equating (4.53) and (4.51) finally gives
 
∂L ν μν
∂μ ∂ Φα − L g =0 , (4.54)
∂(∂μ Φα )
since the ν are arbitrary. By defining the tensor T μν as
∂L
T μν ≡ ∂ ν Φα − L g μν (4.55)
∂(∂μ Φα )
(4.54) reads
∂μ T μν = 0 . (4.56)
μν
The tensor T , as defined in (4.55), has no specific symmetry properties.
It can, however, always be made symmetric in its Lorentz-indices because
(4.56) does not define the tensor T uniquely. We can always add a term of
the form ∂λ Dλμν , where Dλμν is a tensor antisymmetric in the indices λ and
μ, such that T becomes symmetric.3
3
In classical mechanics the form invariance of the Lagrangian under rotations
leads to the conservation of angular momentum. Analogously, in a relativistic
field theory the form invariance of L under four-dimensional space–time rotations
(Lorentz covariance) leads to the conservation of a quantity that is identified with
the angular momentum of the field. To obtain the same form for the angular
momentum as in classical mechanics it is essential that T μν is symmetric.
48 4 Relativistic Fields

Relation (4.56) has the form of a continuity equation. Spatial integration


over a finite volume yields
⎛ ⎞
  2
d ⎝ ∂T iμ 3
T 0μ (x)d3 x⎠ = − d x = − S (μ) · n dS . (4.57)
dt ∂xi
V V S

Here n is a unit vector vertical on the surface S pointing outwards and S (μ)
is a three-vector:
S (μ) = (T 1μ , T 2μ , T 3μ ) . (4.58)
The surface integral on the rhs of (4.57) is taken over the surface S of
volume V. For localized fields it can be made to vanish by extending the
volume towards infinity. It is then evident that the quantities

P μ = T 0μ d3 x (4.59)

are conserved. These are the components of the four-momentum of the field,
as can be verified for the zeroth component,
   
∂L
0
P = T d x= 00 3
∂ Φα − L d3 x
0
∂(∂0 Φα )

= (Πα Φ̇α − L) d3 x = H , (4.60)

according to (4.9) and (4.10). The spatial components of the field momentum
are  
∂L
P k = T 0k d3 x = ∂ k Φα d3 x . (4.61)
∂(∂0 Φα )
Comparing (4.57) with (4.61) and (4.60) and assuming T to be symmetric
we see that the normal components of the vectors S (μ) in (4.58) describe
the energy-momentum flow through the surface S of the volume V .4 These
properties allow us to identify T μν as the energy-momentum tensor of the
field. For the Lagrangian (4.18) of electrodynamics T μν is just the well-known
Maxwell’s stress tensor.
As already mentioned at the beginning of this chapter these conservation
laws are special cases of Noether’s theorem, which can be summarized for the
general case as follows:
Each continuous symmetry transformation that leaves the Lagrangian
invariant is associated with a conserved current. The spatial integral
over this current’s zeroth component yields a conserved charge.
4
More precisely, S k(μ) denotes the flux of the μth component of the field momen-
tum in the direction xk .
4.2 Symmetries and Conservation Laws 49

4.2.2 Internal Symmetries

Relativistic field theories may contain conservation laws that are not con-
sequences of space-time symmetries of the Lagrangian, but instead are con-
nected with symmetries in the internal degrees of freedom such as, e.g., isospin
or charge.
We therefore now allow for a mixture of the different fields under the
transformation
Φα (x) → Φα (x) = e−iεqαβ Φβ , (4.62)
where  is an infinitesimal parameter and the qαβ are fixed c-numbers. We
then have
δΦα (x) = Φα (x) − Φα (x) = −iεqαβ Φβ (x) . (4.63)
The change of the Lagrangian is given by (4.49)
 
∂L
δL = ∂μ δΦα . (4.64)
∂(∂μ Φα )

If L is invariant under this variation δΦα , then we have


 
∂L
δL = ∂μ δΦα = 0 . (4.65)
∂(∂μ Φα )

Equation (4.65) is in the form of a continuity equation for the “current”


∂L 1
j μ (x) = δΦα . (4.66)
∂(∂μ Φα ) ε

Inserting the field variations δΦα (4.63) yields for the current
∂L
j μ (x) = −i qαβ Φβ . (4.67)
∂(∂μ Φα )

Equations (4.65) and (4.67) imply that the “charge”


 
∂L
Q = j (x) d x = −i
0 3
qαβ Φβ d3 x (4.68)
∂(∂0 Φα )
of the system is conserved. The physical nature of these “charges” and “cur-
rents” has to remain open. It depends on the specific form of the symmetry
transformation (4.62) and can be determined only by coupling the system to
external fields.

4.2.2.1 Example: Quantum Electrodynamics To illustrate this conser-


vation law, the theory of electromagnetic interactions is used as an exam-
ple. However, in contrast to the considerations in Sect. 4.1.1 we now con-
sider a coupled system of a fermion field Ψ (x) and the electromagnetic field
Aμ (x) to determine the physical meaning of the conserved current. Together
50 4 Relativistic Fields

with a quantization procedure this theory is called Quantum Electrodynamics


(QED). The Lagrangian is given by
1
L = − Fμν Fμν + Ψ̄ [iγ μ (∂μ + ieAμ ) − m] Ψ . (4.69)
4
L contains a part that describes the free electromagnetic field (first term). The
second term describes the fermion Lagrangian; it is obtained from the free
particle Lagrangian of (4.47) by replacing the derivative ∂μ by the covariant
derivative
Dμ = ∂μ + ieAμ (4.70)
(minimal coupling). Here e is the electron’s charge (e = −|e|).
The Lagrangian (4.69) is obviously invariant under a variation of the
fermion fields of the form

Ψ → Ψ  = e−ie Ψ . (4.71)

Comparison with (4.62) gives qαβ = e δαβ so that the conserved “current”
given by (4.67) is:
jμ (x) = e Ψ̄ γμ Ψ . (4.72)
Note that this conserved current is exactly the quantity that couples to the
electromagnetic field in (4.69). This property allows one to identify the cur-
rent (4.72) as the electromagnetic current of the electron fields.

4.2.2.2 Example: Scalar Electrodynamics The Lagrangian for the case


of a complex scalar field with charge q interacting with an electromagnetic
field is given by
1 ∗
L = − Fμν Fμν + (Dμ φ) (Dμ φ) − m2 φ∗ φ . (4.73)
4
This Lagrangian is simply the sum of the free electromagnetic Lagrangian
(4.18) and the Lagrangian for a complex scalar field (4.38), where in the latter
again the derivative ∂μ has been replaced – through minimal substitution –
by the covariant derivative Dμ = ∂μ + iqAμ (4.70).
The Lagrangian (4.73) is obviously invariant under the phase transforma-
tions

φ(x) −→ e−iεq φ(x)


φ∗ (x) −→ e+iεq φ∗ (x) (4.74)

The conserved current connected with this invariance can be obtained from
the definition (4.67)
 
μ ∂L ∂L ∗
j = −i qφ + (−q)φ
∂ (∂μ φ) ∂ (∂μ φ∗ )
 ∗ μ ∗ 
= iq φ D φ − φ (Dμ φ) . (4.75)
4.2 Symmetries and Conservation Laws 51

The conserved charge is then given by


    ∗ 
Q = d x j (x) = iq d3 x φ∗ D0 φ − φ D0 φ
3 0
. (4.76)

It is remarkable that now the electromagnetic field Aμ appears in the con-


served current (through the covariant derivative Dμ = ∂ μ + iqAμ ). Again
the conserved current provides the coupling to the electromagnetic field. If
the scalar field is real then the Lagrangian (4.73) can be invariant under the
transformation (4.74) only for q = 0. Equation (4.76) shows that in this case
the conserved charge Q = 0.
5 Path Integrals for Scalar Fields

In this chapter we apply the methods developed in Chap. 3 to the case of


scalar fields. There we showed that the vacuum expectation values of time-
ordered products of x̂ operators could be obtained as functional derivatives of
a generating functional. All these results can be taken over into field theory
remembering that fields play the role of the coordinates of the theory and the
spatial locations x correspond to the indices of the classical coordinates. This
implies that we can obtain the vacuum expectation values of time-ordered
field operators by performing the derivatives on an appropriate functional.
To discuss this functional is the main purpose of this chapter.
It is easy to see that these vacuum expectation values of time-ordered
products of field operators play an important role in quantum field theory.
Each field operator creates or annihilates particles and a time-ordered prod-
uct of field operators can thus describe the probability amplitude for a phys-
ical process in which particles are created and annihilated. The quantitative
information about such a process is contained in the S matrix which can be
obtained from the vacuum expectation values of time-ordered field operators
by means of the so-called reduction theorem which we will derive in Chap.
7. The remainder of this book will therefore be concerned with calculating
these expectation values and with developing perturbative methods for their
determination when an exact calculation is not possible.

5.1 Generating Functional for Fields


We assume that the system is described by a Lagrangian of the form
1 μ 
L (φ, ∂μ φ) = ∂ φ ∂μ φ − m2 φ2 − V (φ) . (5.1)
2
In order to obtain the functional W [J] for fields we note that the fields play
the role of the coordinates of the theory and that sums over the different
coordinates have to be replaced by integrals over the space-time coordinates.
In order to define more stringently what is actually meant by a path integral
for fields we write it down here in detail for a free field Lagrangian (V = 0).
In order to do so we bring it into a form as close as possible to the classical
definition (1.38).

U. Mosel, Path Integrals in Field Theory


© Springer-Verlag Berlin Heidelberg 2004
54 5 Path Integrals for Scalar Fields

We first Fourier-expand the field



d3k
φ(x, t) = qk (t)eik·x . (5.2)
(2π)3

For a real field φ we have



q−k = qk . (5.3)
1
Inserting this expansion into the free Lagrangian gives
 
1  μ 
L = Ld x = 3
∂ φ ∂μ φ − m2 φ2 d3x (5.4)
2
  
1 d3k d3k  /   0 
= qk qk k · k − m2 + q̇k q̇k ei(k+k )·x d3x .
2 (2π)3 (2π)3

Integration over d3x gives (2π)3 δ 3 (k + k ) so that we have



1 d3k  
L= 3
q̇k q̇−k − (k2 + m2 )qk q−k
2 (2π)

1 d3k  ∗ 
= 3
q̇k q̇k − ωk2 qk qk∗ (5.5)
2 (2π)

with ωk = k2 + m2 and using (5.3). Writing qk = Xk + iYk , and grouping
then the coordinates Xk and Yk into a new vector xk gives

1 d3k  2 
L= 3
ẋk − ωk2 x2k . (5.6)
2 (2π)

We now introduce a source density J which we also expand



d3k
J(x, t) = Jk (t)eik·x . (5.7)
(2π)3

With this expansion we obtain for the source term


  
3 d3k d3k
d x J(x)φ(x) = Jk (t)q −k (t) = jk (t)xk (t) (5.8)
(2π)3 (2π)3

where we have grouped Jk and −iJk into a new vector jk .


The Lagrangian (5.6) has the structure of a Lagrangian with quadratic
momentum dependence and constant coefficient, discussed in Sect. 1.3.1; also
the source term reads formally just the same as for the nonrelativistic sys-
tems treated in Chap. 3. We can thus again integrate the momentum depen-
dence out and obtain for the groundstate-groundstate transition amplitude
(cf. (3.23))
1
For ease of notation we drop the vector notation in the indices.
5.1 Generating Functional for Fields 55

W 0 [J] = lim 0tf |0ti J


tf →+∞
ti →−∞

 N n+1   -
n
n
N  iη (Ll +Jk xkl )
1 2
= lim dxkj e l=0 , (5.9)
η→0 2πh̄iη
k=1 j=1

with 
1 d3k  2 
Ll = L (xl , ẋl ) = ẋkl − ωk2 x2kl , (5.10)
2 (2π)3
where the first index of xkl denotes the coordinate and the second the time
interval. If we now identify the integration measure as
 N n+1  n
1 2
Dφ = lim dxkj , (5.11)
η→0 2πh̄iη
j=1 k

we can also write the generating functional for the free scalar theory as
 (    )
1/ μ
 2  20
W [J] = Dφ exp i ∂μ φ∂ φ − m − iε φ + Jφ d x 4
2
 (   )
ε 2 4
= Dφ exp i L (φ, ∂μ φ) + Jφ + i φ d x . (5.12)
2
Here the volume element d4x is given by the Lorentz-invariant expression
d4x = d3x dt . (5.13)
2
The term iεφ /2 with positive ε has been introduced in an ad hoc manner
to ensure the convergence of W when taking the fields to infinity with the
understanding that ultimately ε has to be taken to 0 (cf. the discussion at
the end of Sect. 3.1).
The second line of (5.12) also gives the generating functional for an inter-
acting theory with the interaction V included in L. This can be easily proven
by including an additional interaction in the Lagrangian (5.4) and Fourier-
expanding it. The generating functional for a scalar field theory is thus given
by   4 1
W [J] = Dφ e−i d x[ 2 φ(2+m −iε)φ+V (φ)−Jφ]
2
(5.14)

where now the Lagrangian in the form (4.35) has been used.
In analogy to (3.24) we define a normalized functional
W [J]
Z[J] = . (5.15)
W [0]
Again the infinite normalization factor inherent in W [J] cancels out in this
definition. Since W [J] involves an exponential it will often be convenient in
the following discussions to introduce its logarithm iS[J] which is itself a
functional of J
W [J] ≡ eiS[J] =⇒ S[J] = −i ln W [J] = −i ln Z[J] − i ln W [0] . (5.16)
56 5 Path Integrals for Scalar Fields

5.1.1 Euclidean Representation

In the preceding section we have ensured convergence of the generating func-


tional by introducing the ε-dependent term in the energies. In this subsection
we go back to the alternative method that relies on the Wick rotation that
we introduced in Sect. 3.21 and discuss the Euclidean representation of the
generating functional. This discussion also illustrates in some more detail the
remarks on integrating oscillatory functions made at the end of Sect. 1.3.1.
The real Euclidean space is obtained from Minkowski space by rotating
the real axis in the x0 plane by δ = −π/2 into the negative imaginary axis
(Wick rotation). We denote a space-time point in Euclidean space by xE ; it
is related to the usual space-time point x in Minkowski space by

xE = (x, x4 ) with x4 = ix0 = it . (5.17)

Under the Wick rotation t → −it and x4 thus becomes real. With this defini-
tion we can extend the usual Minkowski-space definitions of volume element
and space-time distance to Euclidean space

d4 xE ≡ d3 x dx4 = d3 x idt = id4 x


3
dx2E = dx2j + x24 = −dx2 . (5.18)
j=1

The kinetic term is given by

∂μ φ ∂ μ φ = ∂0 φ ∂ 0 φ + ∂i φ ∂ i φ = − (∂4 φ) − (∇φ) ≡ − (∂E φ)


2 2 2
(5.19)

and the d’Alembert operator is given by

∂2 ∂2 4
∂2 2
2= 2
− ∇ 2
= − 2 − ∇ 2
= − 2
= − (∂E ) ≡ −2E . (5.20)
∂t ∂x4 a=1
∂x a

With these transformations the generating functional for a free scalar field
in Minkowski space (5.12) becomes in its Euclidean representation
  4
WE [J] = Dφ e− d xE { 2 [(∂E φ) +m φ ]−Jφ} ,
1 2 2 2
0
(5.21)

2 2 2
with (∂E φ) = − (∂φ) . Because x4 is now real, (∂E φ) is always positive
and the exponent is negative definite; the integral thus converges and is well-
defined even without adding in the ε-dependent term. Since the exponent
is furthermore quadratic in the fields, WE0 [J] can be evaluated by using the
techniques for Gaussian integrals that are explained in App. B. Physical
results are then obtained by rotating backwards after all integrations have
been performed.
5.1 Generating Functional for Fields 57

Remembering that in field theory the fields play the role of the coordinates
of a Lagrangian theory we can now directly generalize some of the results of
Chap. 3 to field theory. In particular, we have that WE0 [J] of (5.21) is the
transition amplitude from the vacuum state of the free theory at t → −∞ to
that at t → +∞ under the influence of the external source J (cf. (3.27), so
that the normalized transition amplitude ZE0 is given by

WE0 [J] 0 + ∞|0 − ∞J


ZE0 [J] = = , (5.22)
0
WE [0] 0 + ∞|0 − ∞0

where |0 is the vacuum state of the free theory. Equation (5.21) shows that
the normalized transition amplitude can be understood as an integration of
the source action exp (+ d4xE Jφ) with the weights
 4
(φ) = e− 2 d xE [(∂E φ) +m φ ]
1 2 2 2
0
wE (5.23)

over all fields φ  4



Dφ wE
0
(φ)e d xE Jφ
ZE0 [J] =  . (5.24)
Dφ wE0 (φ)

Equation (3.47) then shows that for a function of the fields O(φ)
 
Dφ wE0
(φ) j Oj (φ(xj )) 
 = 0|T [ Oj (φ̂(xj ))]|0 . (5.25)
Dφ wE
0 (φ)
j

The field operators on the rhs here are those of free fields.
0
In the interacting case all these relations still hold if wE is replaced by a
weight function for the interacting theory and the vacuum state is now that
of the interacting theory which we will denote by |0̃. We obtain the weight
function of the interacting theory by writing in analogy to (5.23)
 4
wE (φ) = e− d xE { 2 [(∂E φ) +m φ ]+V (φ)}
1 2 2 2
(5.26)
  
− V (φ) d4xE
= e− V (φ) d xE e− 2 [(∂E φ) +m φ ] d xE ≡ e wE (φ)
4 1 2 2 2 4 0
.

With (5.25) we then get


 
 Dφ wE (φ) j Oj (φ)
0̃|T [ Oj (φ̂)]|0̃ = 
j
Dφ wE (φ)
  
(φ)e− V (φ) d xE j Oj (φ)
4
Dφ wE 0
=  
Dφ wE 0 (φ)e− V (φ) d4xE

 
ˆ
0|T [ j Oj (φˆ0 )e− V (φ0 ) d xE ]|0
4

=  . (5.27)
ˆ
0|T [e− V (φ0 ) d xE ]|0
4
58 5 Path Integrals for Scalar Fields

Here |0 on the rhs is the vacuum state of the non-interacting free theory
(V = 0) and all the field operators on the rhs are free field operators φ̂0
at all times if they were free at t → −∞. This can be seen from the path
0
integral which contains the free weight wE connected with free propagation.
A perturbative treatment can now be obtained by expanding the exponential
in powers of the interaction V .
In the following chapters it is often useful to think in terms of this proba-
bilistic interpretation even when we work in Minkowski metric. This is possi-
ble by relating Euclidean and Minkowski space through analytic continuation.
6 Evaluation of Path Integrals

Only a limited class of path integrals can be evaluated analytically [3] so that
one is often forced to use either numerical or perturbative methods. In this
section we first calculate the generating functional for free scalar fields, both
by a direct reduction method and, to gain familiarity with this technique,
with the methods of Gaussian integration developed in App. B.2.1. We then
show how more general types of path integrals can approximately be reduced
to the Gaussian form. In this reduction we find a systematic method for a
semiclassical expansion in terms of the Planck constant h̄. Finally, we briefly
discuss methods for the numerical evaluation of path integrals.

6.1 Free Scalar Fields


The generating functional for a free scalar field theory plays a special role
in the theory of path integrals. This is so because of two reasons: first, a
free-field theory is the simplest possible field theory and, second, as we have
just seen in the last section the effects of the interaction term V (φ) can be
described with the help of perturbation theory in which the functional of the
full, interacting theory is expanded around that of the free theory.

6.1.1 Generating Functional

The representation of the generating functional of a free field theory (V = 0 in


(5.14)) as a path integral is still quite cumbersome for practical applications.
For these it would be very desirable if we could factorize out the functional
dependence on J in form of a normal integral; the remaining path integral
would then disappear in the normalization.
In the following we will therefore separate the generating functional (5.9)
into 2 factors, one depending on J and the other one being an integral over φ,
i.e. just a number. For this purpose we start with the generating functional
  4 1
W [J] = Dφ e−i d x[ 2 φ(2+m −iε)φ−Jφ]
2
(6.1)

where the Lagrangian in the form (4.35) has been used. The field φ here is
an integration variable; it does not fulfill a Klein–Gordon equation! We now

U. Mosel, Path Integrals in Field Theory


© Springer-Verlag Berlin Heidelberg 2004
60 6 Evaluation of Path Integrals

introduce a field φ0 that does just that


/ 0
2 + (m2 − iε) φ0 (x) = J(x) . (6.2)

J thus plays the role of a source to φ0 . We now take this field φ0 as a reference
field and expand φ around it, setting φ = φ0 + φ . We thus obtain for the
integrand in the exponent in (6.1)
1  
(φ0 + φ ) 2 (φ0 + φ ) + (φ0 + φ ) m2 − iε (φ0 + φ ) − J (φ0 + φ )
2
1 /  0
= φ 2 + m2 − iε φ
2
1 /  0 1 /  0
+ φ0 2 + m2 − iε φ0 + φ0 2 + m2 − iε φ
2 2
1 /  2 0
+ φ 2 + m − iε φ0 − Jφ0 − Jφ . (6.3)
2
The third and the fourth terms on the rhs give the same contribution when
integrated over. The second term gives, according to (6.2), 12 Jφ0 . Collecting
all terms we therefore have for the action in (6.1)
  
1 /  0
S[φ, J] = − d4x φ 2 + m2 − iε φ − Jφ
2
 
1 /  0 1
= − d4x φ 2 + m2 − iε φ + Jφ0
2 2
/  2 0 3

+ φ 2 + m − iε φ0 − Jφ0 − Jφ . (6.4)

In the last line of this equation we can again apply (6.2) to obtain

1 4 /  0 3
S[φ, J] = − d4x φ 2 + m2 − iε φ − Jφ0 (6.5)
2
We are now very close to our aim to factorize out the J dependence. To
do so we solve (6.2) by writing

φ0 (x) = − DF (x − y)J(y) d4y , (6.6)

where DF , the so-called Feynman propagator, fulfills the equation


/  0
2 + m2 − iε DF (x) = −δ 4 (x) (6.7)

in complete analogy to the nonrelativistic propagator K in Chap. 1. Using



4 d4k −ikx
δ (x) = e (6.8)
(2π)4

we obtain by inverting (6.7)


6.1 Free Scalar Fields 61

d4k e−ik(x−y)
DF (x − y) = . (6.9)
(2π) k − m2 + iε
4 2

Substituting (6.6) into the action (6.5) gives


   
1 
/  2 0 
S[φ, J] = − d x φ 2 + m − iε φ + J(x) DF (x − y)J(y) d y
4 4
2

1 4 /  0 3
= − d4x φ 2 + m2 − iε φ
2

1
− J(x)DF (x − y)J(y) d4x d4y . (6.10)
2
The exponential of the last term no longer depends on φ and can, therefore,
be pulled out of the path integral (6.1). The path integral involving the ex-
ponential of the first term appears also in the denominator of the normalized
generating function (5.15) and thus drops out.
We thus obtain now for the normalized generating functional

Z0 [J] =
W [J]
=e − i
2 J(x)DF (x − y)J(y) d4x d4y . (6.11)
W [0]

This is the groundstate-groundstate (vacuum-vacuum) transition amplitude


for a free scalar field theory. Note that it no longer involves a path integral.

6.1.2 Feynman Propagator

The imaginary part of the mass, originally introduced to achieve convergence


for the path integrals, appears here now in the denominator of the propagator
DF 
d4k e−ikx
DF (x) = . (6.12)
(2π)4 k 2 − m2 + iε
It determines the position of the poles in DF , which are, in the k0 -integration,
at
k02 = k2 + m2 − iε (6.13)
or
k0 = ±ωk ∓ iδ (6.14)

with ωk = + k2 + m2 . The poles are therefore located as indicated in Fig.
6.1.
The location of the poles, originally introduced only in an ad-hoc way to
achieve convergence of the path integrals, determines now the properties of
the propagator of the free Klein–Gordon equation, the Feynman propagator.
This can be seen by rewriting the Feynman propagator DF in the following
form
62 6 Evaluation of Path Integrals

Im k0

–ω k+iδ

Re k0

+ω k–iδ

Fig. 6.1. Location of the poles in the Feynman propagator


d4k e−ikx
DF (x) = (6.15)
(2π)4 k 2 − m2 + iε
 3
d k dk0 e−ikx
=
(2π)4 k02 − k2 − m2 + iε
 3  
d k dk0 −ikx 1 1 1
= e − .
(2π)4 2ωk k0 − ωk + iδ k0 + ωk − iδ

We now first perform the integration over k0 . Since the exponential con-
tains a factor e−ik0 t , the path can be completed in the upper half plane for
t < 0 and in the lower half for t > 0. Cauchy’s theorem then gives 2πi × the
sum of the residues at the enclosed poles

d3k eik·x  
DF (x) = i 3
−Θ(−t)e+iωk t − Θ(t)e−iωk t (6.16)
(2π) 2ωk

(in the second term here an extra “−” sign appears because of the negative
direction of the contour integral).
It can now be shown that DF propagates free fields with negative fre-
quencies backwards in time and those with positive frequencies forwards. To
demonstrate this we write

d3k eik·x  
DF (x) = −i 3
Θ(−x0 )e+iωk x0 + Θ(x0 )e−iωk x0 (6.17)
(2π) 2ωk
 
d3k 1 ikx d3k 1 −ikx
= −iΘ(−x0 ) 3
e − iΘ(x 0 ) e .
(2π) 2ωk (2π)3 2ωk

Here we have changed k → −k in the first integral, which does not change
its value under this substitution.
6.1 Free Scalar Fields 63

The integrands are products of solutions of the Klein–Gordon equation


1
e∓ikx
(±)
fk (x) = √ (6.18)
2ωk
which fulfill the normalization and orthogonality conditions appropriate for
a scalar field (4.76)
  
i d3x fk (x)f˙k (x) − f˙k (x)fk (x) = ±(2π)3 δ 3 (k − k) . (6.19)
(±)∗ (±) (±)∗ (±)

We can thus write



d3k 1 1
iDF (x) = Θ(−x0 ) √ e−ik0 √ eikx (6.20)
(2π)3 2ωk 2ωk

d3k 1 1
+ Θ(x0 ) √ eik0 √ e−ikx
(2π)3 2ωk 2ωk
 
d3k (−)∗ (−) d3k (+)∗ (+)
= Θ(−x0 ) f (0)f (x) + Θ(x ) f (0)fk (x) .
(2π)3 k k
(2π)3 k
0

Comparison with (1.13) shows that negative-frequency solutions are prop-


agated backwards in time, and positive-frequency solutions forward. This
particular behavior is a consequence of the location of the poles relative to
the integration path. This location is fixed by the sign of the ε term which
in turn was needed to achieve convergence for the generating functional.

6.1.2.1 Euclidean Representation The Feynman propagator can also be


given in a Euclidean representation. We can define a corresponding Euclidean
momentum space by requiring that k 0 x0 = k4E xE
4 ; this condition ensures that
a plane wave propagating forward in time does so both in Euclidean and in
Minkowski space. We thus get

kE = (k, k4 ) with k4 = −ik0 . (6.21)

The eigenvalues of 2E (5.20) become k 2 = k02 − k2 = −(k42 + k2 ) = −kE


2
< 0.
This gives for the momentum space volume element and the energy-mo-
mentum distance in Euclidean space

d4 kE ≡ d3 k dk4 = −d3 k idk0 = −id4 k


3
dkE2
= dkj2 + dk42 = −dk2 . (6.22)
j=1

Combining both of these definitions yields for the typical exponent of a plane
wave

kx = kμ xμ = k0 x0 − k · x = ik4 (−i)x4 − k · x
= k4 x4 − k · x = kE xE . (6.23)
64 6 Evaluation of Path Integrals

Note that this is not equal to the Euclidean scalar product. However, in
Fourier transforms, where this expression often appears, we always have an
integration over d3 k and can thus change k → −k; thus in these Fourier
integrals – and only there – we can replace kx by kE xE .
The Feynman propagator can now be rewritten. For that purpose we
choose a different path for the integration over k0 after we have Wick-rotated
the time axis. Instead of integrating along the real energy (k0 ) axis we inte-
grate along the imaginary energy axis. We then close the integration path in
the right half of the k0 plane for t > 0 and in the left half for t < 0. Since in
this way the same poles are included as on the original path the value of the
integral does not change.
The Feynman propagator in its Euclidean representation then reads
 4
d kE e−ikE xE
DF (x) = − i 2 + m2
(2π)4 kE
 4
d kE e−ikE xE
= −i . (6.24)
(2π) k + k42 + m2
4 2

Because k4 is real, the integral no longer contains any poles on its integration
path and is therefore well defined.

6.1.3 Gaussian Integration

In Sect. 1.3 we have already used a Gaussian integral relation to integrate


out the p-dependence of the path integral. In many cases the generating
functions appearing in field theory are of a form that contains the fields and
their derivatives only in quadratic form so that again a Gaussian method can
be used. In order to gain familiarity with this technique we derive in this
section again the generating functional for the free scalar field theory.
We thus apply the Gaussian integration formulas of Sect. B.2.1 to the
generating functional (6.1)
  
W [J] = Dφ e− 2 φ(2+m −iε)φ d x ei Jφ d x .
i 2 4 4
(6.25)

In order to make the matrix structure of the exponent more visible we use
two-fold partial integration and write
  
Dφ e− 2 φ(x)δ (x−y)(2y +m −iε)φ(y) d x d y+i Jφ d x
i 4 2 4 4 4
W [J] =
  
Dφ e− 2 φ(x)[(2y +m −iε)δ (x−y)]φ(y) d x d y+i Jφ d x (6.26)
i 2 4 4 4 4
=
  1
→ Dφ e− { 2 φ(xE )[(−2y +m )δ(xE −yE )]φ(yE ) d yE −Jφ} d xE .
E 2 4 4
6.1 Free Scalar Fields 65

In the last step we have gone over to the Euclidean representation of the
generating functional (cf. (5.21)), using 2E = −2 and δ 4 (xE −yE ) = −iδ 4 (x−
y) (cf. Sect. 5.1.1).
The integrals appearing here are now of Gaussian type and can thus be
integrated by using the expressions developed in the last section. We first
identify the matrix in the Gaussian integration formula (B.18) which we
write in the form
 n
− 12 xT D−1 x+JT x n (2π) 2 T
e 2 B DB .
1
e d x=  (6.27)
det(D ) −1

as
−1
y + m )δ (xE − yE ) .
(x, y) = (−2E 2 4
DE (6.28)
2
The inverse of DE is real with positive eigenvalues (kE + m2 > 0) as required
by the derivation in Appendix B.2.1. It is also symmetric as can be seen by
writing the d’Alembert operator in a discretized form, e.g.
1
d2 1 1
2
φ(x)11 = lim 2 [φ(xi + h) − 2φ(xi ) + φ(xi − h)]
dx xi h→0 h
−1
= Dij φ(xj ) (6.29)

with φ(xi + h) = φ(xi+1 ) etc. and


−1
Dij = δi,j−1 − 2δi,j + δi,j+1 .

Thus the necessary conditions for the application of (6.27) are fulfilled. Ap-
plying now (B.18) gives
/ 0− 1
W E [J] = det(−2E + m2 )δ(xE − yE ) 2
 −1
× e 2 J(xE ) DE (xE ,yE ) J(yE ) d xE d yE .
1 4 4
(6.30)

The Wick rotation back to real times is easily performed by the transfor-
mation (5.20) 2E → −2 − iε; it reintroduces the term +iε to guarantee the
proper treatment of the poles. In this case D−1 becomes
−1
DE (x, y) → D−1 (x, y) = (2x + m2 − iε)iδ(x − y) . (6.31)

and the volume element

d4xE d4yE → −d4 xd4 y (6.32)

so that we finally obtain1


1
Note that formally we could have obtained this also by using (B.18) with
D−1 (x, y) = i(2y + m2 − iε)δ 4 (x − y).
66 6 Evaluation of Path Integrals
 /   0−1
e− 2 J(x) i 2y + m2 + iε δ 4 (x − y)
1 1
J(y) d4x d4y .
W [J] = 
det(D−1 )
(6.33)
The determinant of a matrix A is in general given by the product of its
eigenvalues ai . Therefore we have
 
 
ln(det A) = ln ai = ln ai = tr ln A . (6.34)
i i

In the present case, though, D−1 contains an operator and the notation
det D−1 deserves some explanation. The matrix is given by

−1
    1
D (x, y) = 2 + m iδ(x − y) = 2 + m i
2 2
d4k eik(x−y)
(2π)4

d4k  2 
=i −k + m2 eik(x−y)
(2π)4
 
1    1
= i d4k d4k   eik x −k 2 + m2 δ 4 (k − k  )  e−iky .
(2π)4 (2π)4
Thus the momentum representation of the operator A is given by
 
D−1 (k, k ) = −k 2 + m2 δ 4 (k − k  ) . (6.35)

We now evaluate the trace of the logarithm of this matrix where – in


accordance with (6.34) – the logarithm of a matrix is explained by taking
the logarithm of each of the diagonal elements, i.e. the eigenvalues, after the
matrix has been diagonalized. This yields in the present case
 (  
d4k 
tr ln D−1 (x, y) = d4x d4y δ 4 (x − y) 4
d4k  ei(k x−ky)
(2π)
)
× ln(−k 2 + m2 )δ 4 (k − k  )

d4k
= d4x ln(−k 2 + m2 ) (6.36)
(2π)4
We now determine an explicit form for D itself. This can be found by
starting from the definition of the inverse of D

δ 4 (x − y) = D−1 (x, z)D(z, y) d4z

/   0
= i 2z + m2 − iε δ 4 (x − z) D(z, y) d4z

 
= i δ 4 (x − z) 2z + m2 − iε D(z, y) d4z
 
= i 2x + m2 − iε D(x, y) . (6.37)
6.2 Interacting Scalar Fields 67

This is just the defining equation for the Feynman propagator DF (6.7), so
that we have
D(x, y) = iDF (x − y) . (6.38)
For the normalized generating functional we thus obtain

= e− 2 J(x)DF (x − y)J(y) d x d y .
W [J] i 4 4
Z0 [J] = (6.39)
W [0]

Equation (6.39) is the result (6.11) derived earlier in a different way. The
propagator that appears here is just given by the inverse of the Klein–Gordon
Operator
 −1
DF (x − y) = − 2 + m2 δ(x − y) , (6.40)
i.e. of that operator that appears between the two fields in the Lagrangian
for a free Klein–Gordon field
1
L = − φ(2 + m2 )φ . (6.41)
2

6.2 Interacting Scalar Fields


The path integral, and thus also the generating functional, can be obtained
in closed, analytical form only for a few very special cases [3]. In all other
cases one has to resort to either approximation schemes or numerical meth-
ods. For both of them we give examples in this section. The perturbative
method, which relies on the smallness of any interaction relative to the free
field Lagrangian will be treated in separate chapters.

6.2.1 Stationary Phase Approximation

If the path integral in question is not that over a Gaussian, it can be approx-
imately brought into a Gaussian form by using the so-called stationary phase
or saddle point method. In this method one first looks for the stationary
point of the exponent in the path integral. As explained earlier this will give
a major contribution to the path integral. The remaining contributions are
approximated by expanding the exponent around the stationary point.
We illustrate this method here for the case of a scalar field with self-
interactions. The Lagrangian is given by
1  
L = − φ 2 + m2 φ − V (φ) (6.42)
2
and the action 
S[φ, J] = d4x (L + Jφ) (6.43)
68 6 Evaluation of Path Integrals

is a functional of the field φ and the source J. We next determine the sta-
tionary point by looking for the zero of the functional derivative
1
δS[φ, J] 11  
= − 2 + m2 φ0 (x) − V  (φ0 (x)) + J(x) = 0 ;
!
1 (6.44)
δφ(x) φ0

this is the classical equation of motion corresponding to the action S[φ, J].
The stationary field is just the classical field; the corresponding classical
action is
 ( )
1  
S[φ0 , J] = − d x4
φ0 2 + m φ0 + V (φ0 ) − Jφ0 .
2
(6.45)
2

We now expand S[φ, J] around this stationary field (cf. (B.35),(B.37))

S[φ, J] = S[φ0 , J] (6.46)


 1
1 2
δ S 1
+ d4x1 d4x2 1 [φ(x1 ) − φ0 (x1 )] [φ(x2 ) − φ0 (x2 )] + · · · .
2 δφ(x1 )δφ(x2 ) 1φ0

The second functional derivative appearing here can be obtained by varying


the first derivative (6.44). We thus get from the definition (B.26)
1 1
δ2 S 1 δ / 0 1
1 = − (2 + m )φ + V (φ) − J 1 11
2 
(6.47)
δφ(x2 )δφ(x1 ) 1φ0 δφ(x2 ) φ0

Using now (B.32) we get


1
δ2 S 1 / 0
1 = − 2 + m2 + V  (φ0 ) δ 4 (x2 − x1 ) , (6.48)
1
δφ(x2 )δφ(x1 ) φ0 1

which is an operator. The index 1 here means that the corresponding expres-
sions are to be taken at the point x1 .
The action (6.45) is calculated at the fixed classical field φ0 . It can, there-
fore, be taken out of the path integral so that we finally obtain
  4
W [J] = Dφ e h̄ d x (L+Jφ)
i
(6.49)
  
i
= e h̄ S[φ0 ,h̄J] Dφ exp −
i
d4x1 d4x2
2h̄

4 4/ 0 3
× [φ(x1 ) − φ0 (x1 )] 2 + m2 + V  (φ0 ) 1 δ 4 (x2 − x1 ) [φ(x2 ) − φ0 (x2 )]}

+ ... . (6.50)

In order to facilitate the following discussion we have put the unit of action,
h̄, explicitly into this expression by setting i → i/h̄ and J → h̄J.
6.2 Interacting Scalar Fields 69

The path integral remaining here is now in a Gaussian form. It can be


evaluated after a Wick rotation, just as in the developments leading to (6.33).
After a “coordinate
√ transformation” φ → φ = φ − φ0 and after scaling the
fields by φ → h̄φ we get for the generating functional
4 /   03− 21
W [J] = e h̄ S[φ0 ,h̄J] det i 2 + m2 + V  (φ0 ) δ 4 (x2 − x1 )
i
. (6.51)

We now perform a normalization with respect to the free case (6.33), i.e. to
4 /   03− 12
W0 [0] = det i 2 + m2 δ 4 (x2 − x1 )
− 12
= {A(x1 , x2 )} (6.52)

with A from (6.31); the index 0 on W denotes V = 0. This gives for the
normalized generating functional
W [J]
= e h̄ S[φ0 ,h̄J]
i
W̃ [J] = (6.53)
W0 [0]
 ( )− 12
−1 
× det d z A (x2 , z) (A(z, x1 ) + iV (φ0 (x1 ))) δ (z − x1 )
4 4
.

With A−1 = iDF (6.38) we get


4 / 03− 12
W̃ [J] = e h̄ S[φ0 ,h̄J] det δ 4 (x1 − x2 ) − DF (x2 − x1 )V  (φ0 (x1 ))
i
(6.54)

Our aim is now to write the inverse root of the determinant as a correction
term to the classical action. For this purpose we use (6.34)
1
− 12 − tr ln[. . .]
{det[. . .]} =e 2 . (6.55)

The matrix [. . .] is given by


/ 0
x2 |1 − DF V  (φ0 )|x1  = δ 4 (x1 − x2 ) − DF (x2 − x1 )V  (φ0 (x1 )) . (6.56)

The trace of its logarithm is then given by



tr ln[. . .] = d4x ln [1 − DF (0)V  (φ0 (x))] . (6.57)

We can now write


W̃ [J] = e h̄ S[φ0 ,J]
i
(6.58)
with
 ( ) 
1  
S[φ0 , J] = − d x φ0 2 + m φ0 + V (φ0 ) + h̄ d4x Jφ0
4 2
2

i
+ h̄ d4x ln [1 − DF (0)V  (φ0 (x))] + O(h̄2 ) . (6.59)
2
70 6 Evaluation of Path Integrals

The first line is just the classical action S[φ0 , J]. The two terms can be
summed with a resulting action
 ( )
1  
S[φ0 , J] = − d x 4
φ0 2 + m φ0 + Veff (φ0 ) + h̄Jφ0
2
(6.60)
2

with the effective potential


i
Veff (φ(x)) = V (φ(x)) − h̄ ln [1 − DF (0)V  (φ0 (x))] . (6.61)
2
Expression (6.59) shows that the saddle point approximation amounts to
an expansion of the action in powers of h̄. This is in accordance with the
discussion in Sect. 1.4 that quantum mechanics describes the fluctuations
of the action around the classical path. The potential Veff incorporates the
effects of these fluctuations into a classical potential.
Equation (6.59) suggests a perturbative
/ treatment through
0 an expansion
of the logarithm (ln(1 − x) = − x + x2 /2 + x3 /3 + . . . ) in terms of DF V  ,
i.e. the strength of the potential. The trace corresponds to an integration over
x such that the initial and final space-time points in the individual terms in
the expansion are identical, i.e. to a closed loop integration. With higher
orders in the expansion of the logarithm more and more vertices appear,
but they are always located on this one closed loop. This loop expansion
approximates the quantum mechanical behavior, whereas the perturbation
treatment takes the effects of interactions into account.
If we take φ0 as a constant field, i.e. if we neglect its space-time dependence
through the d’Alembert operator in (6.44), then the operator expression 1 −
DF V  (φ0 ) becomes local in momentum space. In this case we can evaluate
the trace of its logarithm by integrating in (6.60) over the eigenvalues

x| ln [1 − DF V  (φ0 )] |x


  
d4k 1 
= ln 1 − 2 V (φ0 ) . (6.62)
(2π)4 k − m2 + iε

This equation writes the original matrix in x-space as a result of a unitary


transformation of a diagonal matrix in k-space.

6.2.2 Numerical Evaluation of Path Integrals

An alternative method for the evaluation of path integrals is that of direct


numerical computation; with rapidly increasing computer power this method
becomes more and more important nowadays.

6.2.2.1 Imaginary Time Method The generating functional is in general


given by
W [J] = 0|e−i(Ĥ+J)(tf −ti ) |0 (6.63)
6.2 Interacting Scalar Fields 71

for ti → −∞ and tf → +∞, as we have seen in Chap. 3. After a Wick


rotation this becomes

W [J] = lim 0|e−β(Ĥ+J) |0 , (6.64)


β→∞

where β denotes the real Euclidean time.


It is immediately obvious that (6.64) also equals the groundstate expec-
tation value of the statistical operator of quantum statistics if we identify β
with the inverse temperature, i.e. β = 1/T .
Inserting the explicit definition (5.9) yields

 N n+1   β
-
n
n
N  − n+1 (Hl +Jk xkl )
n+1 2
W [J] = lim dxkj e l=0 (6.65)
n→0 2πβ
k=1 j=1

with
1  2 
Hl = ẋkl + ωk2 x2kl + V (x1l , x2l , x3l , . . . , xN l ) . (6.66)
2
k

The first index on the coordinates xkl here labels the momentum in the
Fourier-transform of the fields whereas the second one denotes the time-step.
The expectation value of any operator O(φ) can now be written as (cf.
(5.27)) 
Dφ w(φ)O(φ)
O =  . (6.67)
Dφ w(φ)
with
β
-
w(x) = e− n+1 (ẋ2kl +ωk xkl )+V (x1l ,x2l ,x3l ,...,xN l )+Jk xkl ]
2 2
[1
kl 2 . (6.68)

The bold-faced coordinate x denotes the set of all coordinates xkl . The weight
function w(x) is of Gaussian shape and can, therefore, analytically be nor-
malized into a probability

w(x)
P (x) =  , (6.69)
dx w(x)

so that O can be written as



O = lim dx P (x)O(x) . (6.70)
n→0

The multiple integrals appearing here can be evaluated by a Monte Carlo


technique which samples the integrand at a large number of points, where
each ‘point’ really corresponds to a full path x(t). Given a certain point
x = x1 , . . . , xn one randomly chooses a new point x = x1 , . . . , xn , often by
just changing one single coordinate. One then evaluates
72 6 Evaluation of Path Integrals

P (x )
r= . (6.71)
P (x)
If r is larger than 1, the new point is accepted. If r < 1, on the other hand,
then a random number ρ between 0 and 1 is picked. If ρ < r, then the new
point is also accepted, otherwise it is rejected. This method is repeated until
a large enough number of points is sampled. In this way the most important
regions in x space are sampled, thus generating finally M accepted points
xm . The integral is then approximated by
M
1 
O = O(xm ) . (6.72)
M m=1

The sampling algorithm just described is known after its inventor as the
Metropolis algorithm; it plays an important role in numerical evaluations of
statistical physics expressions.
In the present case of evaluating the generating functional for gs-to-gs
transitions one has to choose a fixed value of β, i.e. the Euclidean time.
The calculations then have to be performed for several values of β with a
subsequent extrapolation to β → ∞ (or T → 0). All the n-point functions can
be expressed in a similar way. Thus, in this way all the correlation functions of
the theory can be computed. Methods of this kind are widely used nowadays
in so-called Lattice Gauge Theories [6, 7] where they are used to exploit the
physics contents of non-Abelian gauge field theories (see Chap. 13).

6.2.3 Real Time Formalism

The major difficulty in evaluating a path integral numerically in real time


stems from the oscillatory character of the integrand. The discretized form
of W (5.9) can – in an obvious abbreviation – be written as the limit of a
multidimensional integral

W = dx eiS(x) , (6.73)

where x is a n-dimensional vector. Importance sampling such as the one


just discussed in the last section cannot directly be used because there is
no positive probability weight function in the integrand. This is the reason
why there are so far no explicitly time-dependent solutions of field theories.
In other fields of physics, such as e.g. investigations of dynamics of chemical
reactions, some progress has been made which is briefly described in this
section, closely following [8].
A weight function can be inserted into (6.73) by a mathematical trick.
Using (B.18) in the form
  T
dx0 det[A/2π]e− 2 (x−x0 ) A(x−x0 ) = 1
1
(6.74)
6.2 Interacting Scalar Fields 73

we can write
  T
det[A/2π]e− 2 (x−x0 ) A(x−x0 ) iS(x)
1
W = dx dx0 e . (6.75)

The Gaussian factor under the integral ensures that values of x close to x0
will contribute the most to the integral. The function S(x) can, therefore, be
expanded around x0
1
S(x) = S(x0 ) + S1 (x)(x − x0 ) + (x − x0 )T S2 (x0 )(x − x0 ) + . . . (6.76)
2
with 1 1
dS 11 d2 S 11
S1 (x0 ) = and S2 (x0 ) = . (6.77)
dx 1x0 dx2 1x0
After inserting this expansion we can perform the x-integration and obtain
from (B.18)
   12
det[A] T −1
e− 2 S1 (x0 )[A−iS2 (x0 )] S1 (x0 )
1
W = dx0 eiS(x0 ) .
det[A − iS2 (x0 )]
(6.78)
If we now use that (6.74) is still approximately valid even if A is a function
of x0 and choose
A = A(x0 ) = iS2 (x0 ) + c−1 1 (6.79)
with c > 0 we obtain
 c
iS(x0 )
− ST1 (x0 )S1 (x0 )
1
W = dx0 e (det[1 + icS2 (x0 )]) e 2
2
. (6.80)

The function
c
− ST1 (x0 )S1 (x0 )
P (x0 ) = e 2 (6.81)
thus provides a probability distribution for sampling the remaining integrand
and the expression can be evaluated with the Monte Carlo method discussed
in the last section. We obtain, therefore,
M
1  iS(xi ) 1
W = e (det [1 + icS2 (xi )]) 2 , (6.82)
M i=1

where the M points xi are taken randomly from the distribution P (x).
For large values of c the points with S1 ≈ 0 will contribute the most to
W . This is just the stationarity condition discussed in Sect. 6.2.1 which corre-
sponds to the classical solution. Since only a few points close to this configura-
tion will contribute significantly, a good sampling with good statistics can be
obtained. On the other hand, for small c the probability distribution becomes
74 6 Evaluation of Path Integrals

broad and the statistics correspondingly worse; however, quantum mechani-


cal effects beyond the one-loop approximation now start to contribute. Thus,
by choosing the proper value of c quantum mechanics can be switched on,
but it is obvious that then the difficulties due the rapid oscillations of the
integrand reappear. The dynamical treatment of many-particle systems thus
presents formidable problems, even to today.
7 Transition Rates and Green’s Functions

The ultimate aim of all our field-theoretical developments in this book is


to calculate reaction and transition rates for processes involving elementary
particles. In this chapter, we therefore, now derive a connection between
these transition rates and expectation values of field operators, the so-called
reduction theorem.

7.1 Scattering Matrix

The typical initial state of a scattering experiment is that of widely separated


on-shell particles at t → −∞. On-shell here means that these particles fulfill
the free energy-momentum dispersion relation. The groundstate of the system
is the state of lowest-energy, i.e. the state with no particles present. At large
times t → +∞ the final state is again that of free, non-interacting on-shell
particles. The vacuum state of the theory is unique and is therefore the same
as the initial vacuum state.
The transition rate of any quantum process, be it a scattering process
m + n → m + n or a decay m → n + n , is determined by the so-called
S-matrix. We define the S-matrix as the probability amplitude for a process
that leads from an ingoing state |α, in  to an outgoing state |β, out . The
particles are assumed to move freely in these asymptotic states outside the
range of the interaction; both of these states can therefore be characterized by
giving the momenta of all participating particles and possibly other quantum
numbers as well, all of which are denoted by α and β. The S-matrix is thus
given by
Sβα = β, out|α, in . (7.1)
We can then also introduce an operator Ŝ that transforms in-states (bras)
into out-states (bras)
β, out| = β, in|Ŝ (7.2)
so that we have
Sβα = β, in|Ŝ|α, in . (7.3)
Ŝ is a unitary operator. This can be seen by taking the hermitian conjugate
of (7.2) and writing

U. Mosel, Path Integrals in Field Theory


© Springer-Verlag Berlin Heidelberg 2004
76 7 Transition Rates and Green’s Functions

β, in|Ŝ Ŝ † |α, in = β, out|α, out = δα,β . (7.4)

Thus we have
Ŝ Ŝ † = 1. (7.5)
In Sect. 2.3 we have expressed the matrix element Sβα in terms of wave-
functions. There we showed that S (see (2.34)) could be written as

Sβα = Ψβ∗ (x, t → +∞)Ψα(+) (x, t → +∞) d3x . (7.6)

Here Ψ (+) was a wavefunction that fulfilled an “in” boundary condition, i.e.
it evolved forward in time, starting from an incoming free plane wave at
t → −∞. Ψβ , on the other hand, was a free plane wave for t → +∞.
We now generalize these considerations to fields and introduce the free
asymptotic fields φin and φout

φin (x, t) = lim φ(x, t)


t→−∞
φout (x, t) = lim φ(x, t) . (7.7)
t→+∞

The field operators transform in the standard way under the unitary trans-
formation S
φ̂out = Ŝ † φ̂in Ŝ . (7.8)
Using now the time-development operator U on φin gives for the S- matrix

Sβα = φout |U (+∞, −∞)|φin  (7.9)

with the time-development operator (1.15). With

U (+∞, −∞) = U (+∞, t0 )U (t0 , −∞)

we can write the S-matrix also as


(−) (+)
Sβα = φout |U (+∞, t0 )U (t0 , −∞)|φin  = φout (t0 )|φin (t0 ) . (7.10)

This form corresponds to (7.6).


As is obvious from its definition the S-matrix determines all the transition
rates possible within the field theory. For example, a cross-section for a 1+1 →
1 + 1 collision is simply given by the absolute square of S, multiplied with
the available phase-space of the outgoing particles and normalized to the
incoming current [10]. The Reduction Theorem, to be derived in the following
section, provides a link between the S-matrix and expectation values of time-
ordered products of field operators that can be calculated as derivatives of
generating functionals.
7.2 Reduction Theorem 77

7.2 Reduction Theorem


Since the asymptotic states appearing in the S matrix are those of free, on-
shell particles we can describe them as non-interacting quantum excitations
of the vacuum of the theory with free dispersion relations. We, therefore, first
review the basic properties of free field creation and annihilation operators
in the next subsection.

7.2.1 Canonical Field Quantization

The field operators are obtained by quantizing the free asymptotic fields φ̂in
and φ̂out . This is done in the usual way by imposing commutator relations for
the fields and their momenta. We impose the canonical commutator relations
of quantum mechanics for coordinates and corresponding canonical momenta.
Remembering that in field theory the fields play the role of the classical
coordinates we thus impose

Π̂(x, t), φ̂(x , t) = − iδ 3 (x − x )

Π̂(x, t), Π̂(x , t) = 0

φ̂(x, t), φ̂(x , t) = 0 . (7.11)

We now employ the normal mode expansion of the fields (see the discus-
sion at the start of Sect. 5.1) and momenta

d3k 1  
φ̂(x, t) = 3
√ a(k; t) eik·x + a† (k; t) e−ik·x
(2π) 2ωk

d k ωk 
3 
Π̂(x, t) = −i 3
√ a(k; t) eik·x − a† (k; t) e−ik·x (7.12)
(2π) 2ωk

with ωk = k2 + m2 . Since the fields φ̂ and the momenta Π̂ are now opera-
tors, the ’expansion coefficients’ a and a† are operators as well.
The inverse Fourier transformation is given by
  
1
a(k; t) = √ e−ik·x ωk φ̂(x, t) + iΠ̂(x, t) d3x
2ωk
  
† 1
a (k; t) = √ eik·x ωk φ̂(x, t) − iΠ̂(x, t) d3x (7.13)
2ωk

Using the commutator relations (7.11) we obtain also the commutator


relations for the operators ak and a†k
/ 0
a(k; t), a† (k ; t) = (2π)3 δ 3 (k − k )
/ 0
[a(k; t), a(k ; t)] = a† (k; t), a† (k ; t) = 0 . (7.14)
78 7 Transition Rates and Green’s Functions

The asymptotic in and out fields are free fields so that the time-dependence
of their operators a and a† is harmonic

ain,out (k; t) = a(k; 0)e−iωk t


a†in,out (k; t) = a† (k; 0)e+iωk t (7.15)

as can be obtained from the Heisenberg equation of motion. The asymptotic


in and out fields φ̂in,out are then given by

d3k 1  −ikx †

φ̂in,out (x, t) = √ ain,out (k; 0) e + ain,out (k; 0) e +ikx
(2π)3 2ωk
 3  
dk † ∗
= ain,out (k; 0) f k (x) + ain,out (k; 0) f k (x) , (7.16)
(2π)3

with the scalar product kx = ωk t − k · x and


1
fk (x) = √ e−ikx . (7.17)
2ωk
˙
Remembering that for free fields Π̂ = φ̂, the free field annihilation and
creation operators for the in and out states are given in terms of the fields
and momenta as1
  
a†in,out (k; 0) = −i d3x fk (x) ∂t φ̂in,out (x, t) + iωk φ̂in,out (x, t)
  
ain,out (k; 0) = i d3x fk∗ (x) ∂t φ̂in,out (x, t) − iωk φ̂in,out (x, t) . (7.18)

The operators (7.18) are the same as the ones used in the well known
algebraic treatment of the harmonic oscillators for the normal field modes.
The a† are the creation and the a the annihilation operators for free field
quanta and the vacuum (groundstate) of the free field theory is given by
a|0 = 0.

7.2.2 Derivation of the Reduction Theorem

In a realistic physics situation the scattering or decay processes that we aim to


describe involve interactions between the particles. Simulating the situation
in a scattering experiment we, therefore, now assume that the interactions
between the particles are adiabatically being switched on and off; adiabatically
here means without energy transfer. The asymptotic “in” and “out” states are
thus free states which can be described by the free field operators (7.18). The
operators a† (t) and a(t) in (7.13) have a harmonic time-dependence (7.15)
only for times t → ±∞. The free field operators (7.18) acting on the vacuum
1
We now drop the vector notation of the momenta k in the operators a and a† .
7.2 Reduction Theorem 79

of the full, interacting theory create and annihilate field quanta only at times
t → ±∞ while they do so at all times when acting on the vacuum state of
the noninteracting theory. They can thus be used to describe the asymptotic
states.
We can thus write for the S matrix element (7.1)

Sβα = β, out|α, in = β, out|a†in (k)|α − k, in , (7.19)


where we have assumed that the in-state |α, in contained a free particle with
three-momentum k ; |α −k, in is then that in state in which just this particle
is missing. We can further write this as
β, out|α, in = β, out|a†out (k)|α − k, in (7.20)
+ β, out|a†in (k) − a†out (k)|α − k, in .

In the first term β, out|a†out (k) = β − k, out| (= 0, if β, out| does not
contain a particle with momentum k).
We now rewrite the in and out creation operator into a more compact
form 

a†in,out (k) = −i d3x fk (x) ∂ t φ̂in,out (x, t) (7.21)

with
↔ ∂ φ̂ ∂f

f (t) ∂ t φ̂(x, t) = f φ̂ . (7.22)
∂t ∂t
With this expression we can get the S-matrix element into the form
β, out |α, in = β − k, out|α − k, in (7.23)

↔ 
− iβ, out| d3x fk (x) ∂ t φ̂in (x) − φ̂out (x) |α − k, in .

For only 2 particles in the in and out states the first term on the rhs represents
a single particle transition matrixelement. In this case it can obviously only
contribute if both particles do not change their energy and momentum, i.e. if
|out and |in are identical. It is then just the forward scattering amplitude.
The rhs of (7.23) is time-independent. We can see this by calculating the
time derivative of its integrand
 ↔   
∂t f (x) ∂ t φ̂in (x) = f ∂t2 φ̂in − ∂t2 f φ̂in . (7.24)

Here f (x) = e−ikx and φ̂in both solve the free Klein–Gordon equation with
the same mass. We therefore have
 ↔ 
∂t f (x) ∂ t φ̂in (x) = f ∇2 φ̂in − (∇2 f )φ̂in . (7.25)

The integral over this expression vanishes after twofold partial integration of
one of the terms on the rhs. The same holds, of course, for the term involving
φ̂out in (7.23).
80 7 Transition Rates and Green’s Functions

The rhs of (7.23) is thus indeed time-independent. We can, therefore, take


it at any time and in particular also at t → ±∞. Then we can replace the in
and out fields at these times by the limits (7.7) of the field φ̂(x). This gives
for the S-matrix element

β, out|α, in = β − k, out|α − k, in




+ lim iβ, out| d3x fk (x) ∂ t φ̂(x)|α − k, in
t→+∞


− lim iβ, out| d3x fk (x) ∂ t φ̂(x)|α − k, in . (7.26)
t→−∞

We now write this expression in a covariant form by using


   

lim − lim d3x f (x) ∂ t φ̂
t→+∞ t→−∞


+∞   
∂  ↔   
= dt d3x f (x) ∂ t φ̂ = d4x f ∂t2 φ̂ − ∂t2 f φ̂
∂t
−∞

/  0
= d4x f ∂t2 − (∇2 − m2 )f φ̂ . (7.27)

Note that here φ̂ is an interacting field, since we integrate now over all times.
Thus, in contrast to (7.24) φ̂ does not solve the free Klein–Gordon equation
and, consequently, this integral does not vanish. Twofold partial integration
in the second term in the last line of (7.27) allows us now to roll the Laplace
operator from f over to φ̂. This gives
   
d x f ∂t φ̂ − (∂t f )φ̂ = d4x f (x)(2 + m2 ) φ̂(x) .
4 2 2
(7.28)

We thus have

β, out|α, in = β − k, out|α − k, in (7.29)



+ i d4x fk (x)(2 + m2 )β, out|φ̂(x)|α − k, in .

In this expression we have removed one particle from the in state.


We now continue by removing one particle with the momentum k  from
the out state by going through exactly the same steps as before. Disregarding
the first term in (7.29), that contributes only to forward scattering, we get

β, out|φ̂(x)|α − k, in = β − k  , out|aout (k  )φ̂(x)|α − k, in


= β − k  , out|φ̂(x)ain (k  )|α − k, in (7.30)
  
+ β − k , out|aout (k )φ̂(x) − φ̂(x)ain (k )|α − k, in .
7.2 Reduction Theorem 81

We next replace the annihilation operators by the corresponding field opera-


tors as in (7.21) (here we have to take the hermitian conjugate operator) and
obtain

β, out|φ̂(x)|α − k, in = β − k  , out|φ̂(x)|α − k − k  , in (7.31)


  
↔ 
+i d3x β − k  , out| fk∗ (x ) ∂ t φ̂out (x )φ̂(x) − φ̂(x)φ̂in (x ) |α − k, in .

Taking now the limits t → ±∞ gives, as above, for this expression

β − k  , out|φ̂(x)|α − k − k  , in

∂  ↔  
+ iβ − k  , out| d4x  fk∗ (x ) ∂ t T φ̂(x )φ̂(x) |α − k, in
∂t
  
= . . . + iβ − k  , out| d4x fk∗ (x )∂t2 T φ̂(x )φ̂(x)
   
− ∂t2 fk∗ (x ) T φ̂(x )φ̂(x) |α − k, in . (7.32)

We now use again (7.28) and obtain


 
= · · · + i β − k  , out| d4x fk∗ (x )(2 + m2 )T φ̂(x )φ̂(x) |α − k, in .

Combining this result with (7.29) finally gives (neglecting the forward scat-
tering amplitude)

β, out|α, in = i 2
d4x d4x fk∗ (x )fk (x) (7.33)

× (2 + m2 )(2 + m2 )β − k  , out|T φ̂(x )φ̂(x) |α − k, in .

This reduction can obviously be continued on both sides until all parti-
cles have been removed from the states and we have (with n particles with
momenta k  in the out state and m particles with momenta k in the in state)

Sβα = β n k  , out|α m k, in (7.34)


 m  n
= im+n d4xi d4xj fk∗ (xj )fki (xi )
j
i=1 j=1

× (2j + m2 )(2i + m2 )0|T φ̂(x1 )φ̂(x2 ) . . . φ̂(xn )φ̂(x1 )φ̂(x2 ) . . . φ̂(xm ) |0 .

This is the so-called Reduction Theorem that enables us to express the S-


matrix in terms of the (n + m)-point Green’s function, sometimes also called
the correlation function
82 7 Transition Rates and Green’s Functions

G(x1 , x2 , . . . , xn , x1 , x2 , . . . , xm ) (7.35)



= 0|T φ̂(x1 )φ̂(x2 ) . . . φ̂(xn )φ̂(x1 )φ̂(x2 ) . . . φ̂(xm ) |0 .

Note that in the reduction theorem (7.34) the information about the in-
teraction of the particles is contained in the (interacting) field operators φ̂.
Also the vacuum appearing here is that of the full, interacting theory. The
n + m-point Green’s function appearing there is, therefore, also that of the
interacting theory!
The physical process described by the reduction theorem is that of m on-
shell particles in the initial state at the asymptotic space-time coordinates
x1 , . . . , xm and n on-shell particles at the space-time coordinates x1 , . . . , xn
with an interaction region in between these sets of coordinates. Because the
Klein–Gordon operators 2 + m2 = −DF−1 (6.40) the operators 2 + m2 , when
acting on G, just remove the propagators from the interaction region out to
the asymptotic points, creating so-called vertex functions

Γ (x1 , . . . , xn , x1 , . . . , xm ) (7.36)


m  n
= im+n (2j + m2 )(2i + m2 )G(x1 , x2 , . . . , xn , x1 , x2 , . . . , xm )
i=1 j=1

The appearance of the functions fk (x) in (7.34) then amounts to replacing


the propagators for the external lines by the asymptotic, free fields.
The reduction theorem (7.34) then gives the transition rate Sβα as the
Fourier transform of this vertex function. The Fourier transform in (7.34) con-
tains factors of the form exp(ik  x ) for the outgoing particles and exp(−ikx)
for the incoming ones. This could be symmetrized by changing all outgoing
momenta k  → −k  . This gives

Sβα = β n − k  , out|α m k, in


 m   n
1 1  
= d4xi d4xj 5 √ e−i(kj xj +ki xi )
i=1 j=1 2ωkj 2ωki
× Γ (x1 , x2 , . . . , xn , x1 , x2 , . . . , xm ) . (7.37)

Here all particles, ingoing and outgoing ones, are treated symmetrically so
that from now on no such distinction has to be made in the vertex function
Γ or the Green’s function G.
The connection between the S-matrix and the Green’s function as ex-
pressed by the reduction theorem has been derived here only for scalar fields,
but it is valid in general. The only formal difference is that the Klein–Gordon
operator 2 + m2 has to be replaced by the corresponding free-field operator
and the free field wave functions by their corresponding counterparts.
It is also important to note that the method of canonical quantization used
here to derive the reduction theorem has been used only for the asymptotic
7.2 Reduction Theorem 83

states. Thus, even for fields where this method runs into difficulties when
interactions are present, like, e.g., the gauge fields to be treated in Chap. 12,
the reduction theorem holds in the form given above.
In the remainder of this book we will be concerned with calculating the
correlation functions by using path integral methods. Once these correlation
functions are known the reduction theorem allows us to calculate any reaction
rate or decay probability.
8 Green’s Functions

In Chap. 7 we have found that all the S-matrix elements can be calculated
once the correlation, or Green’s, functions are known. In this chapter we
discuss how these functions can be obtained as functional derivatives of the
generating functionals of the theory.

8.1 n-point Green’s Functions


In Chap. 7 we have seen that the correlation function, i.e. the vacuum expec-
tation value of time-ordered field operators,

G(x1 , x2 , . . . , xn ) = 0|T φ̂(x1 )φ̂(x2 ) . . . φ̂(xn ) |0 . (8.1)

determines the transition rate for all physical processes. Remembering that
in field theory the field operators play the role of the coordinates in classical
quantum theory we can now directly use the results obtained in Sects. 3.2
and 5.1.1 and write using (3.47)

G(x1 , x2 , . . . , xn ) = 0|T φ̂(x1 )φ̂(x2 ) . . . φ̂(xn ) |0

Dφ φ(x1 )φ(x2 ) . . . φ(xn )eiS[φ]
=  (8.2)
Dφ eiS[φ]

with

+∞

S[φ] = L(φ, ∂μ φ)d4x . (8.3)


−∞

The vacuum here is that of the full, interacting Hamiltonian.


The latter expression can also be obtained as a functional derivative of the
generating functional of the theory (cf. (3.47)) so that we can also equivalently
define the n-point Green’s function by
 n 1
1 δ n Z[J] 1
G(x1 , x2 , . . . , xn ) = 1 . (8.4)
i δJ(x1 )δJ(x2 ) . . . δJ(xn ) 1J=0

U. Mosel, Path Integrals in Field Theory


© Springer-Verlag Berlin Heidelberg 2004
86 8 Green’s Functions

We note that G(x1 , . . . , xn ) is a symmetric function of its arguments.


Therefore, according to (B.35–B.37) the following relation holds
 1 
Z[J] = dx1 . . . dxn in G(x1 , x2 , . . . , xn )J(x1 )J(x2 ) . . . J(xn ) . (8.5)
n
n!

Connected Green’s Functions. Guided by (3.22) we define a functional S[J]


by the relation
Z[J] = eiS[J] (8.6)
and introduce the so-called connected Green’s functions Gc in terms of S[J]
defined by the relation
 n−1 1
δn S 1
1 1
Gc (x1 , . . . , xn ) = 1 . (8.7)
i δJ(x1 ) . . . δJ(xn ) 1
J=0

The name of this correlation function and its physics content will become
clear later in this section.

8.1.1 Momentum Representation

Very often it is advantageous to work in momentum space because the exter-


nal lines of Feynman graphs represent free particles with good momentum.
In general the transformation of the Green’s function into the momentum-
representation is given by

e−i(p1 x1 +p2 x2 +...+pn xn ) G(x1 , x2 , . . . , xn ) d4x1 d4x2 . . . d4xn

= (2π)4 δ 4 (p1 + p2 + . . . + pn ) G(p1 , p2 , . . . , pn ) (8.8)

The δ-function here reflects the momentum conservation due to translational


invariance. This can be seen by performing pairwise transformations of two
space-time points to their c.m. point and their relative coordinate. If we
then assume that G depends only on the latter, the integral over the c.m.
coordinate can be performed and yields the δ-function. As in our discussion
around (7.37) we take all the momenta as pointing into the vertex (see Fig.
8.1).

8.1.2 Operator Representations

Operator representation of the generating functional. For completeness, we


now derive an alternative expression for the generating functional Z[J]. We
start by defining the operator functional

Ẑ[J] = T ei J(x)φ̂(x) d x
4
(8.9)
8.1 n-point Green’s Functions 87

p1 p4
p5
p2

p6
p3 p7

Fig. 8.1. Momentum representation of the n-point function. Note that all momenta
are pointing into the shaded interaction region

where φ̂ is an operator! If we form the functional derivatives of Ẑ[J] we get,


in analogy to (3.43),
 n 
1 δ n Ẑ
= T φ̂(x1 ) . . . φ̂(xn )Ẑ[J] , (8.10)
i δJ(x1 ) . . . δJ(xn )

so that, because of Ẑ[0] = 1 and (8.2) together with (8.4),


  1
δ n 0|Ẑ[J]|0 11 
n
1
1 = 0|T φ̂(x1 ) . . . φ̂(xn ) |0
i δJ(x1 ) . . . δJ(xn ) 1
J=0
 n 1
1 δn Z 1
= 1 . (8.11)
i δJ(x1 ) . . . δJ(xn ) 1J=0

Thus all the functional derivatives of 0|Ẑ[J]|0 agree with those of Z[J] at
J = 0. According to (8.4) and (8.5) the two expressions therefore have to be
equal
Z[J] = 0|Ẑ[J]|0 . (8.12)

Functional form of the scattering operator. After having seen that the Green’s
functions can be obtained as functional derivatives of a generating functional
in this section we show that the scattering operator Ŝ (cf. (7.3)) can also be
written in a functional form as

δ
i φ̂in (x)(2 + m2 ) d4 x
Ŝ = : e δJ(x) : Z[J]|J=0 (8.13)
∞ ( 
ik
= d4x1 . . . d4xk : φ̂in (x1 ) . . . φ̂in (xk ) :
k!
k=0
)
    δk
× 21 + m . . . 2 k + m
2 2
Z[J] |J=0 .
δJ(x1 ) . . . δJ(xk )
Here the : : symbol denotes the so-called normal ordered product of field
operators. This normal-ordered product is defined in such a way that all
88 8 Green’s Functions

operators in it are reordered so that all the annihilation operators are moved
to the right. This reordering takes place without a sign change for boson
fields and with a sign-change for each pairwise exchange for fermion fields.
δ
The operator (2 + m2 ) δJ(x) in (8.13) acts only on Z[J].
The matrix elements of the operator (8.13) indeed agree with (7.34). This
can be seen by considering again a matrix element with n particles in the out
states and m particles in the in state. If we consider only processes in which
all particles change their state then in the expansion of the exponential in
(8.13) only that term can contribute that contains exactly m + n powers of
the fields. We thus have
 
m  m+n

m+n
n, out|Ŝ|m, in = i 4
d xi d4xj (8.14)
i=1 j=m+1
1
× n| :φ̂in (x1 ) . . . φ̂in (xm+n ): |m
(m + n)!
× (21 + m2 )(22 + m2 ) . . . (2m+n + m2 ) im+n G(x1 , x2 , . . . , xm+n ) .

Here the in field operator φ̂in (x) can simply be replaced by operators of the
free field (7.7). The normal product reorders the expansion (8.14) such that
all the annihilation operators are on the right. Since each field contains two
independent sums over positive and negative energy eigenstates (cf. (7.16))
we have in total 2m+n operator products; of these only the term with n
creation operators and m annihilation operators can contribute. This gives
with the expansion (7.16)1

n|:φin (x1 ) . . . φin (xm+n ):|m


n
 n+m

(m + n)!
= n| fk∗ (xk )a†k fkl (xl )al |m
m! n! k
k=1 l=n+1
n
 n+m

(m + n)!
= fk∗ (xk ) fkl (xl ) . (8.15)
m! n! k
k=1 l=n+1

The degeneracy factor in front of the matrix element follows from the bi-
nomial expansion of the individual terms in the normal mode expansion
(fk ak + fk∗ a†k )m+n . After inserting (8.15) into (8.14) the integration there
over the xi and xj just gives an extra degeneracy factor m!n! which cancels
the one in (8.15). This finally yields the same expression as in (7.34).

1
Strictly speaking, this equation is correct only under the integrals of (8.14) since
an exchange symmetry has been used to generate the factorials.
8.2 Free Scalar Fields 89

8.2 Free Scalar Fields


We consider first the case of free fields. In this case the generating functional
can be given analytically (6.11)

Z0 [J] = e− 2 J(x)DF (x−y)J(y) d x d y .
i 4 4
(8.16)
The first functional derivative vanishes at J = 0 because the integral (8.16)
is Gaussian. For the second functional derivative we obtain

δ 2 Z0 [J]
= −iDF (x1 − x2 ) e− 2 J(x)DF (x−y)J(y) d x d y
i 4 4

δJ(x1 )J(x2 )

+ (−i) 2
d4x d4y DF (x1 − x)DF (x2 − y)J(x)J(y) (8.17)

× e− 2 J(x)DF (x−y)J(y) d x d y ,
i 4 4

so that we have
1
δ 2 Z0 [J] 11
G0 (x1 , x2 ) = − = iDF (x1 − x2 ) . (8.18)
δJ(x1 )δJ(x2 ) 1J=0
The two-point function for the free scalar field theory is thus just the
Feynman propagator. It is therefore also a solution of (cf. (6.7))
/  0
2 + m2 − iε G(x1 , x2 ) = −iδ 4 (x) (8.19)

8.2.1 Wick’s Theorem

The higher order derivatives can be most easily obtained by expanding (8.16)
∞   n
1 i
Z0 [J] = − J(x)DF (x − y)J(y) dx dy (8.20)
n=0
n! 2
∞  n 
1 i
= 1+ − dx1 . . . dx2n D12 D34 . . . D2n−1 2n J1 J2 . . . J2n
n=1
n! 2

with the shorthand notation Dij = DF (xi − xj ), Jk = J(xk ). Noting that Z


always contains even powers of J, it is immediately evident that all n-point
functions with odd n vanish because an odd functional derivative of an even
function always vanishes at J = 0.
Taking now the 2k-th functional derivative of Z0 and using (8.4) and
(B.36) we obtain
 2k
1 δ 2k Z0
G0 (x1 , x2 , . . . , x2k ) = |
i δJ1 . . . δJ2k J=0
 k 
i 1
= Dp1 p2 . . . Dp2k−1 p2k (8.21)
2 k!
P
90 8 Green’s Functions

where the sum runs over all permutations (p1 , p2 , . . . , p2k ) of the numbers
(1, 2, . . . , 2k). The factor in front of the sum removes the double counting
because of the symmetry Dp2 p1 = Dp1 p2 (2k ) and because of the random
order of factors under the sum (k!).
Equation (8.21) states that the n-point function of a system of free bosons
can be written as a properly normalized and symmetrized product of two-
point functions. This is one version of the so-called Wick’s theorem.
As an explicit example we consider the case n = 4. We then have
1 
G0 (x1 , x2 , x3 , x4 ) = − Dp1 p2 Dp3 p4 . (8.22)
8
P ∈S4

Among the 24 terms in the sum, 12 are pairwise equal because the product
of the two propagators commutes. Furthermore, the propagator Dp1 p2 and
Dp2 p1 are pairwise equal. Thus, there are only 24 : 2 : 2 : 2 = 3 essentially
distinct terms in the sum; the factor 1/8 just takes care of all the others. We
thus have

G0 (x1 , x2 , x3 , x4 ) = − DF (x1 − x2 )DF (x3 − x4 ) (8.23)


− DF (x1 − x3 )DF (x2 − x4 )
− DF (x1 − x4 )DF (x2 − x3 ) .

This result allows an easy physical interpretation: the four-point function


represents an amplitude for connecting the four space-time points x1 , . . . , x4 .
This amplitude is given as a sum over the partial amplitudes for pairwise
connections of all points.
The first few n-point Green’s functions are therefore – according to (8.18),
(8.21) and (8.23) – given by

G0 (x1 ) = 0 (8.24)
G0 (x1 , x2 ) = 0|T [φ(x1 )φ(x2 )] |0 = iDF (x1 − x2 )
G0 (x1 , x2 , x3 ) = 0
G0 (x1 , x2 , x3 , x4 ) = 0|T [φ(x1 )φ(x2 )φ(x3 )φ(x4 )] |0
= − DF (x1 − x2 )DF (x3 − x4 )
− DF (x1 − x3 )DF (x2 − x4 )
− DF (x1 − x4 )DF (x2 − x3 )
etc. (8.25)

It is essential to realize that here |0 is the vacuum state of the free Hamil-
tonian because the n-point function was obtained as a functional derivative
of the non-interacting functional Z0 (8.20). All n-point functions with odd n
vanish.
Wick’s theorem thus allows us to calculate any n-point function of the
free theory in terms of combinations of the well-known Feynman propagator.
8.2 Free Scalar Fields 91

8.2.2 Feynman Rules

As in the classical case (cf. Sect. 2.3) we can again represent these results in
a graphical way. The Feynman rules, that establish the connection between
the algebraic and the graphical representation, are for the case of free fields
still rather trivial.
They are given by
1) Each Feynman propagator is represented by a line:
= iDF (x − y) .
x y
2) Each source is represented by a cross: = iJ(x) .
x
3) There is an integration over all space-time coordinates of the currents .
4) Each diagram has a factor that takes its symmetry under exchange of
external lines into account.
With rule 1) we get, for example, for the fourpoint function (8.23), i.e.
the two-particle Green’s function
1 2 1 2
1 2
G(x1 , x2 , x3 , x4 ) = + + (8.26)
3 4
3 4 3 4
Each line connecting the two points x and y denotes the free propagator and
gives a factor iDF (x − y).
We now set
Z0 [J] = eiS0 [J] (8.27)
with 
i
iS0 [J] = − d4x d4y J(x)DF (x − y)J(y) . (8.28)
2
By using rules 2), 3) and 4) we find immediately

iS0 = (8.29)
This example illustrates nicely the symmetry factor. The endpoints could be
exchanged without changing the result; this corresponds to an exchange of
the integration variables x and y in (8.28). The symmetry factor is, corre-
spondingly, 1/2.
Now we expand

Z0 [J] = eiS0 [J] = e


 
i
= 1+ − J(x)DF (x − y)J(y) d4x d4y
2
  2
1 i
+ − J(x)DF (x − y)J(y) d4x d4y + · · ·
2! 2
92 8 Green’s Functions

1 2 1 3
= 1 + iS0 [J] + (iS0 [J]) + (iS0 [J]) + . . .
2! 3!
1 1
= 1+ + + + ... . (8.30)
2! 3!

We now call all graphs that hang together connected graphs and all the
others unconnected graphs. In our simple case here S0 [J] is represented by
only one connected graph (8.29), whereas Z0 [J] – through the power expan-
sion (8.30) of the exponential function – generates unconnected graphs as
well.
The connected Green’s function can be obtained from its definition (8.7)
as
1 δ 2 S0
Gc (x1 , x2 ) = | = iDF (x1 − x2 ) . (8.31)
i δJ(x1 )δJ(x2 ) J=0
All higher functional derivatives of S0 [J] vanish. Gc is in this free case thus
just given by the Feynman propagator.

8.3 Interacting Scalar Fields


In this section we consider Lagrangians of the form

L = L0 − V (φ) , (8.32)

where L0 is the free scalar Lagrangian (5.1) and V represents a self-interaction


of the field. For such a Lagrangian the generating functional for the n-point
functions can no longer be given in closed form. In order to obtain the n-point
function one has to resort to perturbative methods.
The n-point function then follows from its definition (8.2)

Dφ φ(x1 )φ(x2 ) . . . φ(xn ) eiS[φ]
G(x1 , x2 , . . . , xn ) =  (8.33)
Dφ eiS[φ]

with the exponential in the generating functional of the form



iS[φ] i d4x (L0 −V +i 2ε φ2 )
e =e . (8.34)

The action exponential can be Taylor-expanded in the interaction strength



   N
1
eiS[φ] = −i d4x V eiS0 [φ] (8.35)
N!
N =0

with the free action


8.3 Interacting Scalar Fields 93
  
ε
S0 [φ] = d4x L0 + i φ2 . (8.36)
2
Inserting this into (8.33) gives the n-point function of the interacting theory
in terms of powers of the interaction and the free-field action

Dφ φ(x1 )φ(x2 ) . . . φ(xn ) eiS[φ]
G(x1 , x2 , . . . , xn ) =  (8.37)
Dφ e iS[φ]

 ∞
   N
1
Dφ φ(x1 )φ(x2 ) . . . φ(xn ) −i d x V (φ(x))
4
eiS0 [φ]
N!
N =0
=   ∞   N
1
Dφ −i d x V (φ(x))
4
eiS0 [φ]
N!
N =0

By using (8.2) and the developments in Sects. 3.2 and 5.1.1 we can rewrite
this equation also in terms of normalized vacuum expectation values. The last
line of (8.37) involves the free action S0 and thus free propagation. There-
fore, if the field operators are those of free in fields at asymptotic times they
remain so even during propagation. Also the vacuum appearing in the quan-
tum mechanical vacuum expectation value is then that of the non-interacting
theory so that we also get

G(x1 , x2 , . . . , xn ) = 0̃|T φ̂(x1 )φ̂(x2 ) . . . φ̂(xn ) |0̃ (8.38)
 ∞   N 
 1
0|T φ̂in (x1 ) . . . φ̂in (xn ) −i V̂ (φ̂in (x)) d x
4
|0
N!
N =0
=  ∞ N  .
 1  
0|T −i V̂ (φ̂in (x)) d x
4
|0
N!
N =0

Equation (8.38) is just the Wick-rotated equation (5.27). As in (5.27) |0̃


denotes here the vacuum state of the full, interacting Hamiltonian, whereas
|0 is that of the free Hamiltonian.
With (8.38) we have achieved a remarkable result: (8.38) expresses the
expectation value of the time-ordered product of field operators in the vacuum
state of the full, interacting theory by a perturbative expansion over free
field expectation values. The latter can be calculated as path integrals over
products of classical fields and powers of the interaction (8.37). This enables
us to calculate G, and ultimately also the scattering matrix S (Chap. 7),
perturbatively up to any desired order in the interaction.

8.3.1 Perturbative Expansion


Equation (8.37) allows us to calculate the perturbative expansion of the full
Green’s function up to any desired order in V . Alternatively, these higher or-
94 8 Green’s Functions

der terms can also be obtained as functional derivatives of the free generating
functional Z0 [J] which is known.
According to our general considerations the functional for the Lagrangian
(8.32) is given by
  4 ε 2
Z[J] = Z0 Dφ ei d x (L0 −V (φ)+Jφ+i 2 φ )
  4  4 ε 2
= Z0 Dφ e−i d x V (φ) ei d x (L0 +Jφ+i 2 φ ) . (8.39)

Here Z0 is just the inverse of the path integral for J = 0: Z0 = Z[0]−1 .


We now use the relation
 4  4
1 δ ε 2 ε 2
ei d x (L0 +Jφ+i 2 φ ) = φ(y) ei d x (L0 +Jφ+i 2 φ ) . (8.40)
i δJ(y)
This relation, read from right to left, will also be true for any Taylor-
expandable function V (φ), as can be seen by expanding V into a series in
powers of φ. We thus have also
 ( ) 
i d4x (L0 +Jφ+i 2ε φ2 ) 1 δ ε 2
ei d x (L0 +Jφ+i 2 φ )
4
V [φ(y)] e =V (8.41)
i δJ(y)
and consequently, after exponentiation, also
 4 
i d4x (L0 +Jφ+i 2ε φ2 )
e−i d y V [φ(y)] e
 δ   4
−i d4y V 1i δJ(y) ε 2
=e ei d x(L0 +Jφ+i 2 φ ) . (8.42)

This relation allows us to take the V -dependent factor in (8.39) out of the
path integral
  δ  
−i d4y V 1i δJ(y) i d4x (L0 +Jφ+i 2ε φ2 )
Z[J] = Z0 e Dφ e . (8.43)

The last path integral in (8.43) can be expressed in terms of the free two-
particle propagator introduced in the last section. We thus have
  δ  
−i d4z V 1i δJ(z)
e− 2 J(x)DF (x−y)J(y) d x d y
i 4 4
Z[J] = Z0 e
  δ 
−i d4x V 1i δJ(x)
= Z0 e eiS0 [J]
 4 1 δ 
−i d x V i δJ(x)
= Z0 e Z0 [J] (8.44)

with the free functional Z0 [J] defined in (8.16). The free normalization has
now been absorbed into Z0 . Expanding the exponential that contains the
interaction V then gives the perturbative expansion for Z[J]

    N
1 1 δ
Z[J] = Z0 −i d x V
4
Z0 [J] . (8.45)
N! i δJ(x)
N =0
8.3 Interacting Scalar Fields 95

The connection with (8.37) is then

G(x1 , x2 , . . . , xn )
 ∞
   N
1
Dφ φ(x1 )φ(x2 ) . . . φ(xn ) −i d4x V (φ(x)) eiS0 [φ]
N!
N =0
=  ∞
   N
1
Dφ −i d4x V (φ(x)) eiS0 [φ]
N!
N =0
 n
1 δ n Z[J]
= | (8.46)
i δJ(x1 ) . . . δJ(xn ) J=0
 n ∞
    N
1 1 δn 1 δ
= Z0 −i d4x V Z0 [J] |J=0
i N ! δJ(x1 ) . . . δJ(xn ) i δJ(x)
N =0

Since we will later on be mostly interested in connected graphs we do


not need Z[J] directly but instead its logarithm. We, therefore, now expand
the functional iS[J] = ln Z[J] in powers of the interaction V . We start by
inserting a factor 1 = exp (+iS0 ) exp (−iS0 ) between Z0 and the exponential
in (8.44) and taking the logarithm
  4 1 δ

ln Z[J] = ln Z0 + ln 1 · e−i d x V ( i δJ ) eiS0 [J]
  4 
= ln Z0 + iS0 + ln e−iS0 e−i d x V eiS0
   4 1 δ

= ln Z0 + iS0 [J] + ln 1 + e−iS0 [J] e−i d x V ( i δJ ) − 1 eiS0 [J]
= iS[J] . (8.47)

A perturbation theoretical treatment is now based on a Taylor expansion of


the logarithm. For that purpose we abbreviate
  4 
ε[J] = e−iS0 [J] e−i d x V − 1 eiS0 [J] (8.48)

and obtain

iS[J] = ln Z[J] = ln Z0 + iS0 [J] + ln(1 + ε[J]) (8.49)


 
1
= ln Z0 + iS0 [J] + ε[J] − ε2 [J] + O(ε3 ) .
2

Equation (8.49) represents an expansion of S in powers of the small (for


V → 0) quantity ε.
In order to obtain a perturbative expansion in terms of the potential V we
now rearrange the expansion (8.49). First we expand ε[J] of (8.48) in terms
of the strength of the interaction. This gives
96 8 Green’s Functions
   
1 δ
ε[J] = e−iS0 [J] −i d4x V (8.50)
i δJ
(   )2 +
1 1 δ
+ −i d x V
4
+ · · · e+iS0 [J] .
2! i δJ

We now insert this expression into (8.49) and obtain


 
1 2
iS[J] = ln Z0 + iS0 [J] + ε[J] − ε [J] + . . .
2
(   )
−iS0 [J] 1 δ
= ln Z0 + iS0 [J] + e −i d x V
4
eiS0 [J]
i δJ
(   )2
1 1 δ
+ e−iS0 [J] −i d4x V eiS0 [J]
2! i δJ
 (   ) 2
1 −iS0 [J] 1 δ
− e −i d x V 4
eiS0 [J]
+ O(V 3 )
2 i δJ
1 2
= ln Z0 + iS0 [J] + iS1 [J] + iS2 [J] − (iS1 [J]) + O(V 3 ) ,
2
(8.51)

with

i
iS0 [J] = − J(x)DF (x − y)J(y) d4x d4y
2
(   )
1 δ
iS1 [J] = e−iS0 [J] −i d4x V e+iS0 [J]
i δJ
(   )2
1 −iS0 [J] 1 δ
iS2 [J] = e −i d x V
4
e+iS0 [J] . (8.52)
2! i δJ

Equation (8.51) represents a perturbative expansion of S in powers of V .


Note that the term of second order in the interaction V receives contributions
both from the linear and the quadratic term in the original expansion (8.49)
in ε. The term ∼ S12 obviously just represents an iteration of the linear term.
The connected Green’s functions of the interacting theory can all be ob-
tained as functional derivatives of the functional S[J]. Using the perturbative
expansion for the latter, which involves only S0 [J], then implies that the in-
teracting Green’s function can all be represented in terms of the free Green’s
functions.
9 Perturbative φ4 Theory

In this chapter we apply the formalism developed in the preceding chapter


to the so-called φ4 theory whose Lagrangian is given by
g
L = L0 − V (φ) = L0 − φ4 . (9.1)
4!
Here g is a coupling constant. This φ4 theory is a prototype of a field theory
with self-interactions. It serves as a didactical example which exhibits all
phenomena of more complex field theories.

9.1 Perturbative Expansion of the Generating Function


We start with the generating functional for connected Green’s functions (8.51)
1 2
iS[J] = ln Z0 + iS0 [J] + iS1 [J] + iS2 [J] − (iS1 [J]) + O(V 3 ) . (9.2)
2
Inserting the interaction of φ4 theory (9.1) into (8.52) we obtain

i
iS0 [J] = − d4z d4y J(z)DF (z − y)J(y) ,
2

ig −iS0 [J] δ4
iS1 [J] = − e d4x 4 eiS0 [J] (9.3)
4! δJ (x)
and
 2 
1 −ig δ4 δ4
iS2 [J] = e−iS0 [J] d4x d4y 4 eiS0 [J] .
2! 4! δJ (x) δJ 4 (y)
For notational convenience in the following we now introduce the Si defined
by
ig
iS1 [J] = − S̃1 [J]
4!
 2
−ig
iS2 [J] = S̃2 [J] . (9.4)
4!
The generating function for the connected Green‘s functions in φ4 theory
then reads

U. Mosel, Path Integrals in Field Theory


© Springer-Verlag Berlin Heidelberg 2004
98 9 Perturbative φ4 Theory
 2  2
−ig −ig 1 −ig
iS[J] = ln Z0 + iS0 [J] + S̃1 [J] + S̃2 [J] − S̃1 [J] +... .
4! 4! 2 4!
(9.5)
Since each of the Si [J] contains functional derivatives of the known S0 [J] we
can now evaluate the functional derivatives of S[J] and obtain all the Green’s
functions.

9.1.1 Generating Functional up to O(g)

First, we calculate the term linear in g



−iS0 [J] δ4
S̃1 [J] = e d4x 4 e+iS0 [J] (9.6)
δJ (x)
 
δ4
e− 2 J(z)DF (z−y)J(y) d z d y .
i 4 4
−iS0 [J]
=e d4x 4
δJ (x)

The fourth functional derivative appearing here is most easily obtained by


going to a discrete representation. Noting that

∂ 4 − i Ji Dij Jj
e 2 = [−3Dkk Dkk + 6iDkk (DJ)k (DJ)k (9.7)
∂Jk4
+ (DJ)k (DJ)k (DJ)k (DJ)k ] e− 2 Ji Dij Jj
i

we obtain (k =x)
ˆ

S̃1 [J] = −3 DF (x − x)DF (x − x) d4x

+ 6i DF (y − x)DF (x − x)DF (x − z)J(y)J(z) d4x d4y d4z

+ [DF (x − y)DF (x − z)DF (x − v)DF (x − w)
0
× J(y)J(z)J(v)J(w) d4x d4y d4v d4w d4z . (9.8)

The first term has no sources and will, therefore, not contribute to any Green’s
function, the second term is quadratic in J and thus contributes to the two-
point function and the last term here has the structure of a point interaction
of four fields generated by independent sources at y, z, v, and w and thus
contributes only to the four-point function.

9.1.1.1 Feynman Rules We can again represent these results in a graphical


form by using the rules given in Sect. 8.2.2. These were
1) propagator: iDF (x − y) =
x y
2) source: iJ(x) =
x
9.1 Perturbative Expansion of the Generating Function 99

3) Integration over the space-time coordinates of the sources


4) Symmetry factor for each diagram
We supplement these now for the interacting theory by the additional rules
−ig
5) Each interaction is represented by a dot: =
4!

6) Integration d4x for each loop.
If we represent the interacting connected functional iS[J] by a double
line
iS[J] = (9.9)
we can draw the graphs for
−ig
iS[J] = ln Z0 + iS0 [J] + S̃1 [J] + O(g 2 ) (9.10)
4!
as
= ln Z0 + (9.11)
⎛ ⎞
w z
⎜ x ⎟
⎠ + O(g )
2
+⎝ x + +
y x z y v
The first graph on the right-hand side is again the zeroth order term (8.29),
whereas the graphs in the parentheses represent all the terms of first order
in the interaction (9.8). The first one of these, corresponding to the first
integral in (9.8), describes a process without any external lines. This is a
vacuum process that takes place regardless if physical particles are present or
not; it constitutes a background to all physical processes. The second graph
with the single loop, corresponding to the second integral in (9.8), describes
a mass change due to the self-interaction that we will discuss in the next
section. The third graph, finally, the last integral in (9.8) describes a true
interaction process.
The symmetry factors in these graphs are the factors in front of the inte-
grals in (9.8); they can be obtained as follows:
• The first graph carries the factor 1/2 as explained in Sect. 8.2.2.
• To construct the vacuum graph we pick one of the 4 legs of the interaction
vertex and then connect with any of the other three free legs; there are
always 2 pairwise equal loops. Thus in total we get a weight of 4×3/(2×2) =
3.
• For the second term in parentheses in (9.11) we have four legs of the vertex
to connect with the external line to y; this gives 4 possibilities. The external
line to z can then still be connected with 3 remaining vertex legs. Since
there is an exchange symmetry between y and z we get an additional factor
1/2 (as in S0 ), so that the weight of this vertex becomes 6.
100 9 Perturbative φ4 Theory

• The last graph, finally, is obtained by joining one of the four legs of the
vertex to one of the external points, say z. This generates 4 possibilities.
Next we join any one of the 3 remaining free legs of the vertex to the
external point y; there are obviously 3 ways to do this. The remaining 2
legs can be joined in 2 different ways with the two external points v and w.
Thus, there are in total 4! = 24 possibilities. However, since all the external
points v,w,y and z can be exchanged without changing any of the physics
(v,w,y and z are integration variables), there are also 4! identical terms so
that the last graph in the parentheses in Fig. 9.11 carries the weight 1.

These weights (symmetry factors) have to be multiplied for each graph


to the analytical expression obtained by following the rules given above for
the translation of the pictorial representation into an analytical one. Indeed,
using the symmetry factors just given and following the Feynman rules for
the graphs (9.11) gives the expression (8.51) (together with (9.4) and (9.8)).
The phases of the three terms in (9.8) follow directly from observing the
proper factors i contained in the propagators and in the source (rules 1) and
2)).

9.1.1.2 Vacuum Contributions We now consider the normalization term


ln Z0 . Since we are working with normalized generating functionals, Z0 is
given by W [0]−1
  4 ε 2
Z0−1 = Dφ ei d x(L0 −V +i 2 φ ) (9.12)
 4 1 δ  i  1
−i d xV i δJ(x) − J(x)DF (x−y)J(y)d4x d4y 1
=e e 2 J=0 .

We can now treat this expression in exactly the same way as we just did for
S[J]; the only change being that we have to take the final result at J = 0.
This gives (see (9.10))

(−ig)
ln Z0 = −iS[0] = −iS0 [0] − S̃1 [0] + O(g 2 ) . (9.13)
4!
In the graphical
 representation this reads, using S0 [0] = 0 and S1 [0] =
−3DF2 (0) d4x (cf. (9.8)),

ln Z0 = − (9.14)

The normalization constant, or – in other words – the denominator of the


generating functional, thus contains just the vacuum graph. Inserting (9.14)
into (9.11) then removes the vacuum contribution giving, finally, for Z[J] the
graphical representation up to terms of first order in the interaction

= + + (9.15)
9.2 Two-Point Function 101

Although we have shown here only for first-order coupling that the de-
nominator in Z[J] (see (5.15)) just removes the vacuum contributions, this
is a general result that holds to all orders of perturbation theory.

9.2 Two-Point Function


The connected n-point function can now be obtained from its definition in
(8.7) 1
 n−1 1
1 δn S 1
Gc (x1 , . . . , xn ) = 1 . (9.16)
i δJ(x1 ) . . . δJ(xn ) 1
J=0
In this section we work out the connected two-point function in the lowest
orders of the coupling constant.

9.2.1 Terms up to O(g 0 )


There is only one connected Green’s function in the free case which is just
given by the Feynman propagator (8.31)

d4q −iq(x1 −x2 ) i
DF (x1 − x2 ) = e . (9.17)
(2π)4 q − m2 + iε
2

9.2.1.1 Momentum Representation We now evaluate the lowest order


(in g) two-point function in momentum space (cf. (8.8)). Taking the Fourier-
transform of the two-point function (8.31) gives

e−i(p1 x1 +p2 x2 ) Gc (x1 , x2 ) d4x1 d4x2 (9.18)
  
d4q −iq(x1 −x2 ) i
= e−i(p1 x1 +p2 x2 ) e d4x1 d4x2 .
(2π)4 q 2 − m2 + iε
We first perform the integrations over x1 and x2 and obtain, according to the
definition (8.1.1), for the rhs
i
(2π)4 δ 4 (p1 + p2 )Gc (p1 , p2 ) = (2π)4 δ 4 (p1 + p2 ) . (9.19)
p21 − m2 + iε
From this equation we can read off the momentum representation of the prop-
agator. The momenta p1 and p2 are incoming momenta that point towards a
vertex. Thus, the momentum representation of the free propagator is given
by
i
G0 (p, p = −p) = 2 , (9.20)
p − m2 + iε
pictured in Fig. 9.1. Note that here the second momentum appears with a
negative sign. This is due to our notation to take all momenta as incoming
(see Fig. 8.1).
102 9 Perturbative φ4 Theory

p −p

Fig. 9.1. Two-point function of a scalar theory

9.2.2 Terms up to O(g)

Up to terms linear in the coupling strength we obtain from (9.5)


1
1 δ2 S 1
Gc (x1 , x2 ) = 1 (9.21)
i δJ(x1 )δJ(x2 ) 1J=0
1
δ 2 S0 (−ig) δ 2 S1 1
= −i − 1 + O(g 2 ) .
δJ(x1 )δJ(x2 ) 4! δJ(x1 )δJ(x2 ) 1
J=0

With S0 [J] from (9.3) and S1 [J] from (9.8) this gives for the two-point func-
tion

Gc (x1 , x2 ) = iDF (x1 − x2 )


  
ig
− − 12i d4xDF (x − x)DF (x − x1 )DF (x − x2 ) + O(g 2 )
4!

g
= iDF (x1 − x2 ) − d4xDF (x2 − x)DF (x − x)DF (x − x1 )
2
+ O(g 2 ) . (9.22)

This is the connected propagator, up to terms of O(g), of the interacting


theory.
We can represent this equation in the following graphical form, where
denotes the “dressed” propagator

= + (9.23)
x1 x2 x1 x2
x1 x2
with the rules developed above. Since the n-point functions involve derivatives
with respect to the source current, taken at zero source, the external lines
of all Feynman graphs do not contain crosses, that depict sources, anymore.
They are instead given by free propagators.
The weight factors for these diagrams have been explained at the end
of Sect. 8.2.2. Since we deal here with Green’s functions with definite, fixed
external points, the exchange symmetry factors must not be divided out
here. Thus, the first diagram on the rhs in (9.23) carries the weight 1 and the
second, the so-called tadpole diagram, the weight 12.
9.2 Two-Point Function 103

9.2.2.1 Momentum Representation In Sect. 8.1.1 we have introduced


the momentum representation of the Green’s function and in Sect. 9.2.1.1 we
have already evaluated it for the free case. Here we now determine it for the
φ4 theory up to terms of order O(g).
Taking the Fourier-transform of the two-point function (9.22) gives

e−i(p1 x1 +p2 x2 ) Gc (x1 , x2 ) d4x1 d4x2 (9.24)
   
d4q e−iq(x1 −x2 )
= e−i(p1 x1 +p2 x2 ) i d4x1 d4x2
(2π)4 q 2 − m2 + iε
  ( 
g −i(p1 x1 +p2 x2 ) d4q1 d4q2 d4q3
− e 4
dx
2 (2π)4 (2π)4 (2π)4
−iq2 (x2 −x) −iq1 (x−x1 )
)
e e
× 2 d4x1 d4x2 .
(q1 − m2 + iε)(q22 + m2 − iε)(q32 − m2 + iε)
As in Sect. 9.2.1.1 we first perform the integrations over x1 , x2 and x and
obtain for the rhs
i
(2π)4 δ 4 (p1 + p2 )Gc (p1 , p2 ) = (2π)4 δ 4 (p1 + p2 ) 2 (9.25)
p1 − m2 + iε

g 1 d4q 1 1
− (2π)4 δ 4 (p1 + p2 ) 2 .
2 p1 − m + iε
2 (2π) q − m + iε p2 − m2 + iε
4 2 2 2

This equation is used to read off the momentum representation of the propa-


gator (see (8.8)). The momenta p1 and p2 are incoming momenta that point
towards a vertex. Thus, the momentum representation of (9.22) is given by
i
Gc (p, p = −p) = (9.26)
p2 − m2 + iε
  
i −ig d4q i i
+S 2
p − m2 + iε 4! (2π)4 q 2 − m2 + iε p2 − m2 + iε
where the symmetry factor is S = 12. Equation (9.26) gives the momentum
representation of the two-point function up to terms of order g. The first
term on the rhs gives the free propagator (9.20) already obtained in Sect.
9.2.1.1, whereas the second term gives the contribution of the interaction to
this two-point function.

Momentum space Feynman rules. The Feynman rules for (9.26) are now
i
1) each free propagator line gives a factor .
q2 − m2 + iε
−ig
2) each vertex gives a factor .
4!
3) there is four-momentum conservation for the sum of all momenta flowing
into a vertex.
104 9 Perturbative φ4 Theory

d4q
4) each internal line gives an integration .
(2π)4
5) to each diagram a weight factor has to be multiplied as explained above.

Self-energy. With the abbreviation



g d4 q i
Σ= (9.27)
2 (2π) q − m2 + iε
4 2

and the free two-point function G0 from (9.20) we can write the two-point
function (9.26) as
   −1
Σ Σ G0
Gc (p, −p) = G0 + G0
G0 = G0 1 + G0 ≈ G0 1 − Σ
i i i
i 1
= 2
p − m2 + iε 1 − Σ p2 −m1 2 +iε
i
= . (9.28)
p2 − m2 − Σ + iε

Here we have consistently kept terms up to O(g).


The quantity Σ appears like an additional mass term in the final result.
It is, therefore, called a self-energy and the second graph on the rhs in (9.23)
is called a self-energy insertion. The appearance of this self-energy is a first
indication that the mass m appearing in the Lagrangian is the mass of the
particle only in a classical theory. In quantum theory it gets changed by the
interactions.

9.2.3 Terms up to O(g 2 )

As noted at the end of Sect. 8.3.1 there are two distinct contributions to
the second order term, one being a genuine term of second order in V and
the other just being an iteration of the first order term. With the help of the
Feynman rules we can now construct the corresponding Feynman graphs. For
the two-point function these are given in Fig. 9.2.
Graph (a) in Fig. 9.2 obviously just represents an iteration of the first
order tadpole graph in (9.23). Its contribution to the two-point function is
given by
  
i −ig d4q i
Ga (p, −p) = Sa 2 (9.29)
p − m2 + iε 4! (2π)4 q 2 − m2 + iε
  
i −ig d4q i i
× 2 ,
p − m2 + iε 4! (2π)4 q 2 − m2 + iε p2 − m2 + iε

Sa is the symmetry factor; it is simply given by the product of the corre-


sponding factors for the one-loop graphs, Sa = 12 · 12 = 144.
9.2 Two-Point Function 105

a) b) c)

Fig. 9.2. Feynman graphs for the two-point function up to O(g 2 )

It is evident from Fig. 9.2a, as well as from its algebraic representation


in (9.29), that the graph can be cut into two parts, each representing a first
order process. This reflects the appearance of the last term in (8.51) that
is simply the square of the first order term. Such a graph that falls apart
into 2 unconnected parts, if one internal line is cut, is called one-particle-
reducible; otherwise it is one-particle-irreducible (1PI). The reducible graph
here is generated by the square of the first order term ∼ S̃12 in (9.5).
In order to facilitate the following discussions we introduce now the ver-
tex function Γ (p1 , p2 , . . . , pn ), sometimes also called connected proper vertex
function, which describes only 1PI graphs and in which the propagators for
the external lines are missing. The n-point vertex function is, therefore, given
by

Γ (p1 , p2 , . . . , pn ) (9.30)
−1 −1 −1
=G (p1 , −p1 )G (p2 , −p2 ) . . . G (pn , −pn )Gc (p1 , p2 , . . . , pn ) .

The free 1PI 2-point function is then given by the inverse propagator
1 2 
Γ (p, −p) = p − m2 = G−1 (p, −p) . (9.31)
i
Note that the product of inverse two-point Green’s functions and the n-body
function is just the combination that appears in the reduction theorem (cf.
(7.37)).
The 1PI part of the Green’s function for the graph 9.2a reads

Γa (p, −p) = G−1 (p, −p)G−1 (−p, p)Ga (p, −p)



−ig d4q i
= Sa (9.32)
4! (2π) q − m2 + iε
4 2

with Sa = 144.
106 9 Perturbative φ4 Theory

For the graph in Fig. 9.2b, which is one-particle irreducible, we get


 2 
−ig d4q d4u i i(2π)4 δ 4 (q − u)
Γb (p, −p) = Sb
4! (2π) (2π) q − m + iε u2 − m2 + iε
4 4 2 2

d4r i
× (9.33)
(2π)4 r2 − m2 + iε

with Sb = 12·12 = 144. For the graph in Fig. 9.2c (also 1PI) we finally obtain
 2 
−ig d4q d4r d4s
Γc (p, −p) = Sc δ(p − (q + r + s))
4! (2π)4 (2π)4 (2π)4
i i i
× 2 , (9.34)
q − m + iε r − m + iε s − m2 + iε
2 2 2 2

with Sc = 4 · 4! = 96.

9.3 Four-Point Function


It is easy to see that the three-point function vanishes for the model consid-
ered here since the third functional derivative of the action (9.5) at J = 0
vanishes.

9.3.1 Terms up to O(g)

The four-point function up to terms of O(g) is given by


 3 1
1 δ4 S 1
Gc (x1 , x2 , x3 , x4 ) = 1 (9.35)
i δJ(x1 ) . . . δJ(x4 ) 1J=0
1 1
δ 4 S0 1 (−ig) δ 4 S1 1
= i 1 + 1
δJ(x1 ) . . . δJ(x4 ) 1 4! δJ(x1 ) . . . δJ(x4 ) 1
J=0 J=0
+ O(g 2 )

where S is given by (9.5) and Fig. 9.2. Because Gc involves the fourth deriva-
tive with respect to the source and S0 depends on J only quadratically (see
(9.3)), only the last term of S1 in (9.8) can contribute to the Green’s function.
Thus we get

Gc (x1 , . . . , x4 ) = −ig d4xDF (x − x1 )DF (x − x2 )DF (x − x3 )DF (x − x4 ) .
(9.36)
In momentum space this is simply the product of the four propagators (9.20)
times the factor −ig. The corresponding Feynman graph is given in Fig.
9.3. It carries the symmetry factor 4!, corresponding to a symmetry under
permutation of all external legs.
9.3 Four-Point Function 107
3 4

1 2

Fig. 9.3. Feynman graph for the four-point function

Unconnected graphs. From the connected graphs calculated so far, we could


reconstruct also the unconnected graphs. Up to terms linear in the coupling
constant g we get in symbolic notation
1 2
Z[J] = e = 1+ + ( ) + ...
 2! 
= 1+ + +
 2
1
+ + +
2!
+ O(g 2 ) . (9.37)
The four-point function is generated by diagrams with four external legs (each
external leg corresponds to a factor J in the generating functional), because
G is given by a fourth functional derivative. Therefore, only the last diagram
in the second line of (9.37) and the square of the first term in the last line
can contribute to the four-point function in this order.
We thus have for the four-point function up to terms linear in the cou-
pling

G(x1 , x2 , x3 , x4 ) = + + (9.38)

In both of the first two diagrams the two particles just move by each other,
without interaction. These unconnected graphs thus do not contribute to any
interaction processes.

9.3.2 Terms up to O(g 2 )

In order to become more familiar with Feynman graphs, we now construct


the connected four-point function up to terms of order g 2 in a graphical way.
This four-point function is shown in Fig. 9.4.
The first line in Fig. 9.4 gives the four-point function just constructed,
with a symmetry factor S1 = 4!. The four diagrams on the second line are
108 9 Perturbative φ4 Theory

3 4
3 4

1 2
1 2

3 4 3 4 3 4 3 4

+ + + +

1 2 1 2 1 2 1 2

3 4
3 4
3 4
q1

+ + + + O(g 3 )
q2
1 2
1 2
1 2

Fig. 9.4. Four-point function of φ4 theory up to terms ∼ g 2

just the basic vertices with self-energy insertions on each of the external
legs; these insertions carry the extra symmetry factor S2 = 12, as we have
seen in Sect. 9.1.1.1. The last three diagrams in the third line are of a new
topological structure. They represent modifications of the basic interaction
vertex through the insertion of internal lines.
Each one of these graphs has the same external lines. We thus have an
overall factor for all graphs


4
i
S1 , (9.39)
p2k − m2
k=1

with the external symmetry factor S1 = 4! so that the full four-point function
is given by
9.3 Four-Point Function 109


4
i
Gc (p1 , p2 , p3 , p4 ) = S1 (G1 + G2 + G3 ) , (9.40)
p2 − m2
k=1 k

where Gi denotes the contribution from the i-th line in Fig. 9.4 without the
external line symmetry factor.
The first basic vertex then just gives the contribution
−ig
G1 = . (9.41)
4!
The graphs of the second line have one loop in addition. We thus have
 
4
−ig 2 d4q i i
G2 = ( ) S2 (9.42)
4! (2π)4 q 2 − m2 p2l − m2
l=1

for their contribution, with S2 = 12. The internal momenta pl here are those
between the loop and the four-point vertex. Since the loop carries no mo-
mentum away they are the same as the corresponding incoming momenta.
The three graphs of the third line, finally, have two internal lines, which
are, however, related by energy - and momentum conservation at the incoming
vertex. They give

−ig 2 d4q1 d4q2 i i
G3 = ( ) S3
4! (2π) (2π) q1 − m q2 − m2
4 4 2 2 2


× (2π)4 δ 4 (q1 + q2 − (pk + pl )) (9.43)
kl

where the sum over (k, l) runs only over the pairs of numbers (1,2), (1,3) and
(1,4), i.e. the external legs at one of the vertices in each of the three graphs.
The δ-function appears because the net momentum running into the dressed
vertex has to be zero (see (8.8)).
The symmetry factor S3 can, for example for the middle graph of Fig.
9.4, be obtained as follows. The external leg 1 can be connected with the left
vertex in 4 different ways; the same holds for the leg 2 with the right vertex.
Once these connections (4 × 4 possibilities) have been done, each of the 3
free legs of the left vertex can be connected with the external leg 3; the same
holds for the connections of the right vertex to the external point 4 (3 × 3
possibilities). Finally, each of the remaining 2 legs of the left vertex can be
connected to one leg of the right vertex; for the remaining leg there is then no
freedom left (2 possibilities). Thus, the symmetry factor for the last 3 graphs
in Fig. 9.4 is
(4 · 4)(3 · 3)2 (4!)2 4!
S3 = = = . (9.44)
S1 2S1 2
110 9 Perturbative φ4 Theory

9.4 Divergences in n-Point Functions


Many of the expressions obtained in the preceding sections for two- and four-
point functions are actually ill-defined because they diverge, as we will show
in this section. We start with a rather general discussion of divergences in φ4
theory and then evaluate explicitly the two- and four-point functions.
To illustrate the divergence of the Green’s functions obtained we consider,
as an example, the two-point function (9.26)
i
Gc (p, −p) = (9.45)
p2 − m2 + iε
 g i i
+ −i iDF (0) 2 .
2 p2 − m2 + iε p − m2 + iε
The loop contribution between the two Feynman propagators in the second
term on the rhs is given by the self-energy

g g d4q i
Σ = iDF (0) = . (9.46)
2 2 (2π)4 q 2 − m2 + iε
The integral here diverges: after integrating over q0 we obtain integrals of the
form (cf. Sect. 6.1.2) 
d3q
 . (9.47)
q 2 + m2
By introducing an upper bound Λ for the integral over |q| and then taking
Λ → ∞ we see that the integral diverges as Λ2 ; this is called a quadratic
divergence. Because this divergence happens here for large q one also speaks
of a quadratic ultraviolet divergence.

9.4.1 Power Counting

A way to characterize the degree of divergence of (9.46) is the so-called


“power-counting”: There are 4 powers of q in the integration measure, but
only two powers of q in the denominator of (9.46); the degree of divergence
D is then given by the net power of q, i.e. D = 2.
The power-counting just illustrated for the case of the tadpole diagram
can be generalized to any Feynman graphs with an arbitrary number of loops.
In order to see this we consider a theory with an interaction ∼ φp in n
dimensions.
 Since each loop contributes according to the Feynman rules an
integral dnq to the total expression and since each internal propagator gives
a power q −2 , we have for the degree of divergence D in a diagram with L
loops and I internal lines
D = nL − 2I . (9.48)
Note that each loop has also at least 1 internal line. For example, the tadpole
diagram has L = 1 and I = 1, giving D = 2 in four dimensions. D > 0 clearly
9.4 Divergences in n-Point Functions 111

diverges, D = 0 corresponds to a logarithmic divergence, and D < 0 seems


to be convergent.
In order to become familiar with this power counting for φ4 theory in four
dimensions we give three examples in Fig. 9.5. According to our earlier con-
siderations we expect that D ≥ 0 diverges. However, this does not guarantee
that graphs with D < 0 actually converge. This is illustrated by the lowest
example in Fig. 9.5, that has L = 1, I = 3, and thus D = 4 · 1 − 2 · 3 = −2, but
diverges because of the loop on the externals legs. Only when each possible
subgraph has also D < 0, then the whole expression converges (Weinberg’s
Theorem).

D = 4 · 1 − 2 · 1 = +2

D =4·1−2·2=0

D = 4 · 0 − 2 · 1 = −2

D = 4 · 1 − 2 · 3 = −2

Fig. 9.5. Examples for graphs in φ4 theory with different degree of divergence. On
the right equation (9.48) is illustrated for n = 4

In φp theory the total number of lines is pV if a graph has V interaction


vertices. This is so because each vertex has p legs; these can be either external
or internal lines. If they are internal, they have to be counted twice because
each internal line originates and disappears at a vertex. Thus we have
pV = E + 2I , (9.49)
112 9 Perturbative φ4 Theory

where E is the number of external lines. In addition, the number of loops, L,


is related to the number of vertices, V , by

L=I −V +1 . (9.50)

Combining equations (9.49) and (9.50) with (9.48) allows us to eliminate L


and I to obtain
  p 
n(p − 2)
D =n+ −p V − −1 E . (9.51)
2 2

The degree of divergence of a connected graph in this φp theory in n dimen-


sions thus depends on the number of external legs and, in general, also on
the number of vertices.
In the perturbative treatment of field theory derived in Sect. 8.3 we have
seen that the order of perturbation theory directly equals the number of
vertices, V , in a Feynman diagram. Thus, the degree of divergence becomes
larger and larger with increasing order of a perturbative treatment, if the
factor of V in (9.51) is positive. On the other hand, D is independent of this
order if that factor is zero and D becomes even smaller when going to higher
orders of perturbation theory if the factor is negative.
Thus the perturbative treatment leads to an infinite number of divergent
terms if the factor multiplying V in (9.51)

n(p − 2)
v≡ −p>0 . (9.52)
2
In this case, for a fixed number of external legs, E, the degree of this di-
vergence gets worse with increasing order of perturbation theory. Looking
at (9.51) shows that with each higher order new divergent n-point functions
with the same D, but a larger number E of external lines, appear. For the
physically interesting case of n = 4 condition (9.52) amounts to p > 4. Thus
classically well defined theories with φ5 , φ6 , . . . selfcouplings have more and
more divergent Green’s functions (i.e. Feynman graphs) pop up with each
order of perturbation theory.
For the theory under discussion here, with a φ4 coupling in n = 4 di-
mensions, the factor v multiplying V in (9.51) is v = 0 and the degree of
divergence D is given by D = 4 − E, i.e. it is given only by the number
of external legs E. In this case there can still be infinitely many divergent
terms, but in each order of perturbation theory only E = 4, i.e. the four-point
function, and E = 2, i.e. the two-point function, are divergent with the same
D. All higher n-point functions lead to D < 0 and thus convergence if also
all subdiagrams have D < 0.
If v < 0, finally, then there is only a finite number of divergent n-point
functions, because from a certain order of perturbation theory on all the
n-point functions will converge.
9.4 Divergences in n-Point Functions 113

As we will show later, in both of the latter two cases one can add a finite
number of so-called counter terms to the Lagrangian that just remove these
divergent terms. In this case a theory is called renormalizable 1 .

9.4.2 Dimensional Regularization of φ4 Theory

Regularization serves to make all divergent expressions convergent, at the


expense of introducing a parameter that has no physical meaning and ulti-
mately has to be removed. If the integrals can be performed and evaluated as
a function of this parameter then the infinite contributions can be separated
from the finite ones.
We show this procedure by evaluating now the divergent integrals by
a modern technique called dimensional regularization. In this method the
’parameter’ is the deviation of the actual dimension from the physical value 4.

Dimensions of coupling terms. The technique starts by considering the La-


grangian of the theory in n dimensions
1 μ  g
L= ∂ φ ∂μ φ − m2 φ2 − φ4 . (9.53)
2 4!
Since the action 
S= L dnx (9.54)

is a dimensionless quantity (in units in which h̄ = 1), L must have the


dimension −n ( is a length). The kinetic energy term in L thus has also
dimension −n and – since [∂μ ] = −1 – we get [φ] = 1−n/2 . Any potential
term of the form gφp must have the same dimension as the Lagrangian, i.e.
n
−n ; we thus get [g] = −n /p(1−n/2) = −(n+p(1− 2 )) . The mass term (p = 2),
2 −2
for example, has [m ] =  in n dimensions, as it should.
Note that the dimension of g is determined by an exponent which is just
the factor of the number of vertices, V , in (9.51). Thus, the dimension of the
coupling constant and the renormalizability of a theory are closely connected:
if the dimension of g is that of a positive or zeroth power of mass, then the
theory is renormalizable.
These considerations show that for p = 4

[g] = n−4 = mass4−n . (9.55)

In four dimensions the coupling constant of a φ4 theory is thus dimensionless


and the theory contains only a finite number of divergent Green’s functions,
i.e. the two-point and the four-point function. The theory is thus renormal-
izable. In n = 4 dimensions, however, this is no longer true.
1
Theories with v < 0 are sometimes called superrenormalizable.
114 9 Perturbative φ4 Theory

If we want to keep the coupling constant dimensionless in order to ensure


renormalizability also in n dimensions we have to modify the φ4 term in
the Lagrangian (8.52) such that an additional factor with the dimension of
(mass)4−n absorbs the dimension so that g can remain dimensionless
g 4−n 4
L = L0 − μ φ . (9.56)
4!
Note that μ here is an arbitrary mass.

9.4.2.1 Two-point Function The two-point function is completely deter-


mined once we know the self-energy. We, therefore, start by calculating this
quantity in lowest order in the coupling by evaluating explicitly the contri-
bution of the tadpole diagram.
To do so we first go to n dimensions so that in lowest order the self-energy
(9.46) becomes 
g dnq i
Σ = μ4−n n
. (9.57)
2 (2π) q − m2 + iε
2

This integral can be obtained analytically by going into a space of n-dimen-


sional polar coordinates (see App. C). The result is (C.17)
g 4−n 1 n
 n
Σ= μ n
mn−2 π 2 Γ 1 − . (9.58)
2 (2π) 2

The divergence of this expression is now manifest, since the Γ -function has
poles at 0 and the negative integers and thus also for n = 4.
We now expand Γ around this pole. For that purpose we write
 n  ε
Γ 1− = Γ −1 + (9.59)
2 2
with ε = 4 − n and expand in powers of ε (cf. (C.4))
 ε 2
Γ −1 + = − − 1 + γ + O(ε) , (9.60)
2 ε
where γ is the Euler–Mascheroni constant (γ ≈ 0.577..). We thus get for n
close to 4, i.e. ε ≈ 0
 
g με 2−ε 4−ε 2
Σ= m π 2 − − 1 + γ + O(ε) (9.61)
2 (2π)4−ε ε
 ε  
m2 4πμ2 2 2
=g − − 1 + γ + O(ε) .
32π 2 m2 ε

We now use
ε→0
xε = eε ln x −→ 1 + ε ln x (9.62)
and obtain
9.4 Divergences in n-Point Functions 115
(  )  
ε→0 m2 ε 4πμ2 2
Σ −→ g 1 + ln − − 1 + γ + O(ε)
32π 2 2 m2 ε
2
(  2
)
m 2 4πμ
= g − − 1 + γ − ln + O(ε)
32π 2 ε m2
(  )
m2 1 m2 4πμ2
= −g − g 1 − γ + ln + O(ε) . (9.63)
16π 2 ε 32π 2 m2

The self-energy (9.63) is a regular, well defined function for all |ε| > 0, i.e. for
all dimensions n = 4. By going away from the physical case n = 4 we have
thus regularized the self-energy.
For n = 4 (ε → 0) the self-energy Σ diverges as 1/ε, but this divergence is
well separated from the rest of the expression; it is only the first term in (9.63)
that diverges as n → 4. The second term is finite and depends on the arbitrary
mass μ that was originally introduced only to keep the coupling constant free
of dimension. The appearance of the arbitrary mass μ in the finite part is
related to the arbitrariness in separating an overall infinite expression into a
sum of an infinite and a finite contribution.
Note that in this order Σ is independent of p. In next higher order (∼ g 2 )
the same is true for the self-energy contribution of Fig. 9.2b whereas that of
Fig. 9.2c depends quadratically on p [12].

9.4.2.2 Four-point Function In this section we will now evaluate the four-
point function and regularize it. By looking at Fig. 9.4 we expect that the
four graphs in the second line just contribute again to the self-energy. On
the other hand, we expect that the three diagrams in the lowest line can all
graphically be contracted such that they contain only one interaction point
with possibly modified interaction strength.
As an example, we now evaluate the middle graph in the last line of Fig.
9.4. According to the Feynman rules its contribution to the 1PI vertex in n
dimensions is (see (9.43))
 2 
−ig 4−n dnq i i
ΔΓ (p1 , p2 , p3 , p4 ) = μ S3 S 1
4! (2π) q − m (p − q)2 − m2
4 2 2

(9.64)
with p = p1 + p3 = p2 + p4 . The symmetry factor is S1 · S3 = 4! · 12 (9.44).
The two denominators in the integrand can be combined into one by a
mathematical trick due to Feynman that is very often used for the evaluation
of such expressions. This trick starts from the elementary integral relation

b
dx 1 b 1 1 b−a
2
= − |a = − + = . (9.65)
x x b a ab
a

By now substituting
x = az + b(1 − z) (9.66)
116 9 Perturbative φ4 Theory

we obtain
b 0
dx dz
= (a − b) 2 . (9.67)
x2 [az + b(1 − z)]
a 1

Combining (9.65) and (9.67) gives


1
1 dz
= 2 . (9.68)
ab [az + b(1 − z)]
0

We now apply this to the integrand in (9.64) and obtain


1
1 1 dz
=
q − m (p − q)2 − m2
2 2
{(q 2 − m2 )z + [(p − q)2 − m2 ] (1 − z)}
2
0
1
dz
= 2 .
[q 2 − 2pq(1 − z) + p2 (1 − z) − m2 ]
0
(9.69)

The first three terms in the denominator can be combined by substituting

q  = q − p(1 − z) . (9.70)

This gives for expression in the denominator

q 2 − 2pq(1 − z) + p2 (1 − z) − m2
= [q − p(1 − z)]2 − m2 − p2 (1 − z)2 + p2 (1 − z)
= q 2 − m2 − p2 z(z − 1) . (9.71)

The 1PI vertex now reads


 1
1 dnq  1
ΔΓ (p1 , p2 , p3 , p4 ) = g 2 (μ2 )4−n dz
2 (2π)n [q 2 − m2 − sz(z − 1)]
2
0
(9.72)
with s = p2 = (p1 + p3 )2 .
We now interchange the order of integration and evaluate the integral
over q  first with the help of (C.17) in Appendix C

dnq  1
(2π)n [q 2 − m2 − sz(1 − z)]2
 
i / 2 0 n−4 n Γ 2 − n2
= m + sz(1 − z) 2
π2 . (9.73)
(2π)n Γ (2)
The four-point function thus reads
9.4 Divergences in n-Point Functions 117

1 2 2 4−n π
n
2
 n
ΔΓ (p1 , p2 , p3 , p4 ) = g (μ ) i n
Γ 2−
2 (2π) 2
1
/ 0 n−4
× dz m2 + sz(1 − z) 2 . (9.74)
0

We now introduce again ε = 4 − n. This gives

1  ε  1 ( m2 + sz(1 − z) )− 2ε
2 ε
ΔΓ (p1 , p2 , p3 , p4 ) = ig μ Γ dz . (9.75)
32π 2 2 4πμ2
0

Now the four-point function is in a form that allows the limit ε → 0 to be


taken. Using again
xε = eε ln x → 1 + ε ln x (9.76)
and (C.2)
ε 2
Γ = − γ + O(ε) (9.77)
2 ε
gives
ΔΓ (p1 , p2 , p3 , p4 )
⎡ ⎤
ε
( ) 1  2 
iμ 2 ε m + sz(1 − z)
= g2 − γ + O(ε) ⎣1 − ln dz ⎦
32π 2 ε 2 4πμ2
0
⎡ ⎤
ε ε 1  
iμ 1 iμ ⎣γ + ln m 2
+ sz(1 − z)
= g2 − g2 dz ⎦ + O(ε)
16π 2 ε 32π 2 4πμ2
0

with s = (p1 + p3 ) = (p2 + p4 )2 .


2
(9.78)
Now the four-point function has been separated into a divergent part and a
convergent term (for ε → 0). The integral appearing is a function of p, m,
and μ; the latter dependence remains even when ε → 0.
So far, we have only calculated the middle graph in the last line of Fig.
9.4. It is evident, however, that the result can also be directly taken over
to the other 2 graphs in the last line by taking for p the appropriate total
momentum at one of the vertices. We, therefore, introduce now – in addition
to s – the three Lorentz-invariant Mandelstam variables
s = (p1 + p3 )2 = p2
t = (p1 + p2 )2
u = (p1 + p4 )2 , (9.79)
with the property s + t + u = 4m2 . These variables represent the squared
total four-momentum at the vertex that involves p1 in each of the graphs in
the last line of Fig. 9.4.
118 9 Perturbative φ4 Theory

We thus get for the sum of all three diagrams in the last line of Fig. 9.4,
the vertex correction diagrams,

Γv (p1 , p2 , p3 , p4 ) (9.80)
2 ε 2 ε
3ig μ 1 ig μ
= − [3γ + F (s, m, μ) + F (t, m, μ) + F (u, m, μ)] .
16π 2 ε 32π 2
Here F (s, m, μ) denotes the integral in (9.78). In this form Γv has been split
up again into a divergent (for the physical ε = 0) and a convergent part.
Of the diagrams in Fig. 9.4 only the first and the three last ones are
1PI graphs. The four diagrams in the second line are all 1P reducible; they
just differ by the propagators on the external legs that are left out when we
consider the 1PI four-point function. The complete 1PI four-point function
is, therefore, given by

Γ (p1 , p2 , p3 , p4 ) = −igμε + Γv (p1 , p2 , p3 , p4 )


= −igμε
3iμε 1 2 iμ
ε
+ g2 2
− g 2
[3γ + F (s, m, μ) + F (t, m, μ) + F (u, m, μ)]
16π
 ε ( 32π
3 1
= − igμε 1 − g
16π 2 ε
)
1
− [3γ + F (s, m, μ) + F (t, m, μ) + F (u, m, μ)] . (9.81)
32π 2

Equation (9.81) gives the effective, “dressed” interaction vertex. By going to


ε = 0 we have regularized the normally divergent four-point function.
By comparison with the free 1PI four-point function, (9.81) suggests to
absorb all effects of the loop graphs, contained in the curly brackets, on the
four-point function into a new effective coupling constant g̃

g {1 − g δΓ (s, t, u, m, ε, μ)} −→ g̃ , (9.82)

with
1
δΓ = Γv , (9.83)
ig 2 με
so that
Γ (p1 , p2 , p3 , p4 ) = −ig̃με . (9.84)
Thus, ‘measuring’ Γ , for example in a physical ( = 0) 1+2 → 3+4 scattering
experiment, determines g̃.
This effective coupling constant depends on s, t, u and m and, as the self-
energy, separates into a divergent and a finite term; it contains all the effects
of the loops up to terms of order g 2 . We could thus use g̃ in our original
Lagrangian and calculate only so-called tree graphs without any vertex loops
in them. These would be automatically taken care of through the dependence
of the effective coupling constant g̃ on the kinematical variables s, t and u.
9.4 Divergences in n-Point Functions 119

9.4.3 Renormalization

In the preceding two subsections we have seen that both the 1PI two-point
and the 1PI four-point functions are for ε = 0 regular functions of ε. These
quantities, that diverge in four dimensions, have thus been regularized. For
n → 4 both have divergent and finite contributions from higher order dia-
grams. In this section we now show how to handle these divergences by the
renormalization technique. Any renormalization procedure requires first a
regularization, either by a cut-off for the upper bounds of diverging integrals
or by going to n = 4 dimensions, to be followed by a procedure in which the
dependence on these artifacts is removed.
Since the separation of a divergent quantity into a finite and an infinite
contribution is arbitrary there are various so-called renormalization schemes.
They all have in common that they either add terms to the Lagrangian or
scale the fields and coupling constants such that the original form of the La-
grangian is maintained. The most obvious scheme is the so-called minimal
subtraction scheme that removes just the pole contributions in the dimen-
sional regularization, i.e. the terms that go like powers of 1/ε. This scheme
is straightforward and well-suited for the dimensional regularization, but it
leads to expressions in which the parameters m and g have no direct relation
to measurable quantities. They just characterize the n-point functions.
Here we discuss another scheme, in which we require that the parame-
ters of the Lagrangian assume their physical, measured values, i.e. m is the
physical, observed mass and g the physical, observed coupling constant. Any
observation, of course, will determine the renormalized parameters. The mass
m is given by the pole position of the renormalized two-point function and the
coupling constant g is determined by the renormalized four-point function.

9.4.3.1 Two-point Function We start by looking at the structure of the


most general two-point function. When we go to terms that depend on g 2 and
higher orders of the interactions we always encounter 1P reducible graphs,
such as the one in Fig. 9.2a. The general structure of the (reducible) two-point
function is of the form (9.28).

i
Gc (p, −p) = . (9.85)
p2 − m2 − Σ + iε
The self-energy Σ is in general a function of the momentum p of the
particle. We can therefore expand Σ(p2 ), for example, around the on-shell
point p2 = m2 , where m is the physical, observable mass

Σ(p2 ) = Σ(m2 ) + (p2 − m2 )Σ1 + Σ2 (p2 ) . (9.86)

Here
∂Σ
Σ1 = | 2 2 and Σ2 (m2 ) = 0 . (9.87)
∂p2 p =m
120 9 Perturbative φ4 Theory

In (9.86) only the first two terms of the expansion of Σ have been written
out explicitly; Σ2 (p2 ) denotes the whole remainder of the expansion. Inserting
this expansion into (9.85) gives for the propagator

i
Gc (p, −p) =
p2 − m2 − Σ(m2 ) − (p2 − m2 )Σ1 − Σ2 (p2 ) + iε
1 i
= . (9.88)
1 − Σ1 p2 − m2 − Σ(m2 )+Σ2 (p2 ) + iε
1−Σ1

The pole of this propagator should be at the physical mass and its residuum
should be i. However, this is in general not the case, if we start with the ob-
servable, physical mass in the Lagrangian because the self-interactions Σ(m2 )
and Σ1 contribute to the self-energy and can shift the pole away from its phys-
ical location. In addition, the residue is changed by the factor 1/(1 − Σ1 ).

Counter terms. In order to get the pole to the correct, physical location we
therefore have to add a so-called counter term to it
1 1
Lcm = − δm2 Zφ2 with Z= . (9.89)
2 1 − Σ1
The purpose of this counter-term is to ’neutralize’ the shifts, induced by the
self-energy, of the pole of the propagator away from the physical, observed
mass m.
Z is usually called “field renormalization constant” for reasons that will
become obvious a little later (see (9.104)). Note that this counter-term has
exactly the same form as the mass term in the original Lagrangian. It will
thus also add a term −Zδm2 in the denominator of the dressed propagator
(9.88).
We determine the unknown δm2 by the requirement

δm2 + Σ(m2 ) = 0 (9.90)

so that the new term just cancels the self-energy contribution at the physical
on-shell point. With the counter-term added, the propagator then becomes
1 i
G(p, −p) = . (9.91)
1 − Σ1 p2 − m2 − ZΣ2 (p2 ) + iε

Since Σ2 (m2 ) = 0 by definition, this propagator has the correct pole at the
physical mass m.
Its residue, however, still is – instead of being simply i –
i
= iZ . (9.92)
1 − Σ1
This deficiency can be cured by adding another, additional counter-term
9.4 Divergences in n-Point Functions 121

1  
Lcφ = (Z − 1) ∂μ φ ∂ μ φ − m2 φ2 (9.93)
2
to the Lagrangian. Then the propagator becomes
i
G(p, −p) = Z
p2 − m2 − ZΣ2 (p2 ) + (Z − 1)(p2 − m2 ) + iε
i
= 2 . (9.94)
p − m2 − Σ2 (p2 ) + iε

This propagator has the pole at the correct, physical mass m (because of
Σ2 (m2 ) = 0) and the correct residue i. By adding the given counter-terms
we have thus removed the divergent quantities Σ(m2 ) and Σ1 from the prop-
agator. Note that the counter-terms all have the structure of terms already
present in the original Lagrangian.

φ4 theory. We now specify all of these general considerations to the example


of φ4 theory. As we have discussed towards the end of Sect. 9.4.1 in φ4 theory
only a finite number of Green’s functions, the two-point and the four-point
function, diverge. These are just the terms we have calculated in the last
two subsections, namely the self-energy contribution (9.63) and the vertex
function (9.81).
In φ4 theory Σ, being the self-energy insertion of a two point function,
is quadratically divergent. Consequently, Σ1 as the first derivative of Σ with
respect to p2 is logarithmically divergent since taking the derivate always
adds one more power of q 2 in the denominator of the Feynman propagators.
The one remaining quantity Σ2 (p2 ) involves second and higher derivatives
of the self-energy with respect to p2 and is thus convergent; by definition it
vanishes at the on-shell point.
For the self-energy we have, up to one-loop diagrams, (cf. (9.63)),
(  )
m2 1 m2 4πμ2
Σ(p ) = −g
2
−g 1 − γ + ln + O(ε) . (9.95)
16π 2 ε 32π 2 m2

Thus, in this order Σ does not depend on p2 and consequently Σ1 = Σ2 =


0 and thus Z = 1. Thus in this one-loop approximation, there is no field
renormalization, but already in the two-loop approximation we would get
Σ1 = 0 because the graph in Fig. 9.2c is p-dependent [12]. To be general we,
therefore, keep the factor Z in the following expressions.
The renormalized 1PI two-point function is thus

Γ (p, −p) = p2 − m2 + O(g 2 ) . (9.96)

Note that it contains only finite terms and does not contain the artificial
mass μ.
122 9 Perturbative φ4 Theory

9.4.3.2 Four-point Function In Sect. 9.4.2.2 we have seen that all the loop
corrections in the last line of Fig. 9.4 can be subsumed in the effective coupling
constant g̃ (9.82). With this g̃ only the elementary four-point function in
the first line of Fig. 9.4 has to be calculated. In other words, in an actual
experiment for a 2 → 2 reaction g̃ and not g itself is determined.
We now want to identify the coupling constant g appearing in the La-
grangian with the observed value of the coupling constant g̃ at a fixed kine-
matical point r = (s, t, u). With this aim in mind we define a vertex renor-
malization constant Zg by
Zg = (1 − g δΓ (s, t, u, m, ε, μ))−1 |r , (9.97)
where r denotes an in principle arbitrary renormalization point. Since we
want to relate the coupling to a measurable quantity we use the so-called
symmetric point  
4 1
pi · pj = m 2
δij − (9.98)
3 3
with i, j = 1, . . . , 4 labelling the external lines. At this point we have again
all particles on shell with s = t = u = 4m2 /3.
We can then introduce the additional counter-term
gμε
Lcv = − (Zg − 1) φ4 . (9.99)
4!
In this way we effectively replace the coupling constant g by gZg . Calcu-
lating with the latter the vertex corrections to the four-point function gives
according to (9.82)
g̃ = gZg {1 − gZg δΓ (s, t, u)}
 
g g
= 1− δΓ (s, t, u)
1 − gδΓ (r) 1 − gδΓ (r)
4 3
= g 1 − g [δΓ (s, t, u) − δΓ (r)] + O(g 2 ) , (9.100)
where we have dropped, for ease of notation, the dependencies of δΓ on
m, ε and μ. Equation (9.81) together with (9.82) shows that in (9.100) the
infinite term ∼ 1/ε is removed because it does not depend on the kinematical
variables s, t, u. Thus g̃ no longer depends on ε and is thus finite and well
defined even for the physical case of n = 4 dimensions, i.e. ε → 0. Closer
Inspection of δΓ , more specifically of the functions F (r, m, μ) appearing in it
(cf. (9.80) and (9.78)), also shows that the dependence on the arbitrary mass
scale μ drops out, so that the renormalized 1PI four-point function is given
by
Γ (p1 , p2 , p3 , p4 )
 ( )
g 4 2 2
= −ig 1 − G(s, m ) + G(t, m ) + G(u, m ) − 3G( m , m )
2 2 2
32π 2 3
+ O(g 3 ) . (9.101)
9.4 Divergences in n-Point Functions 123

Here  1  
G(s, m2 ) = ln m2 + sz(1 − z) dz . (9.102)
0
is no longer μ-dependent. The μ-dependence of Γ , that was still present in
(9.81), has dropped out in the subtraction in (9.101).
Equation (9.101) shows that measuring the four-point function at the
symmetrical point s = t = u = 4m2 /3 gives indeed Γ (r) = −ig and thus the
coupling constant g in the original Lagrangian. It also shows that at other
kinematical points (s, t, u) the coupling constant depends on the kinematics
of the interaction; it is then called a running coupling constant. Note that the
renormalized four-point function is independent of the arbitrary mass scale μ.
However, the counter-term (9.99) depends on this mass scale, even at n = 4,
through the μ-dependence of Zg (9.97).

9.4.3.3 Renormalized Lagrangian The Lagrangian with all the counter-


terms now reads
1/ 0 gμε
L= (∂φ)2 − m2 φ2 −
2 4!
1 / 0 1 gμε
+ (Z − 1) (∂φ) − m2 φ2 − δm2 Zφ2 −
2
(Zg − 1) φ4
2 2 4!
Z/   0 gμε
= (∂φ)2 − m2 + δm2 φ2 − Zg φ4 . (9.103)
2 4!
This Lagrangian leads, in second order perturbation theory, to the correct
physical behaviour for the 2-point Green’s function with the physical mass m
and the proper residue. It also gives the ‘measured’ coupling constant g at the
symmetric point. The fields φ, in terms of which the propagator is defined, are
thus the physical fields, which include already the effects of self-interactions.
By introducing the “bare field” φ0 and the “bare mass” m0 by

φ0 = Zφ
m20 = m2 + δm2 (9.104)

and a “bare coupling constant” by


Zg
g0 = gμε , (9.105)
Z2
the complete Lagrangian can be expressed in terms of bare quantities only
1/ 0 g0
L= (∂φ0 )2 − m20 φ20 − φ40 . (9.106)
2 4!
This bare Lagrangian has the same form as the original one, because all the
counter-terms had the same form as terms already appearing in the original
Lagrangian. As a consequence it leads to finite physical quantities. While we
124 9 Perturbative φ4 Theory

have only shown that this is the case in second order perturbation theory
this result actually holds in all orders. If this is the case, then the theory is
well defined, i.e. it is said to be renormalizable. The bare Lagrangian is really
considered to be the ‘true’ Lagrangian of the theory because it leads only to
finite physical quantities.
In the preceding considerations we have chosen 2 arbitrary renormaliza-
tion points; we have required the propagator to have a pole at p2 = m2 , where
m is the physical mass (cf. (9.94)), and for the vertex we have chosen the
symmetric point r (s = t = u = 4m2 /3) (cf. (9.98)). Choosing other renor-
malization points will lead to other values for the mass and the coupling.
This arbitrariness in the renormalization point reflects the fact that the di-
vergent expressions for the self-energy and the vertex both separate into a
divergent and a finite term (cf. (9.81) and (9.95)). Any scheme, that removes
the infinite terms, will lead to finite expressions for the physical observables,
independent of what it does with the finite contributions. It will also lead to
the same n-point function, and – with the help of the reduction theorem –
to the same observable transition rates.
In the dimensional renormalization discussed here all the counter-terms
and renormalization factors also depend on the arbitrary mass μ, originally
introduced to keep the coupling constant g dimensionless, even after ε → 0.
This arbitrariness, however, is nothing else than the arbitrariness we have
already encountered in the choice of the renormalization point. The physics
must be independent of μ. This observation is the starting point for the
development of the renormalization group method.
It is essential to realize that even in a massless theory, which contains
no dimensional scale parameters, μ introduces a scale that determines the
momentum dependence of the coupling constant. Thus, at the quantum level,
a bare Lagrangian is not enough to specify a theory, but a renormalization
scheme must be added that introduces necessarily a scale into the theory.
10 Green’s Functions for Fermions

For simplicity of notation we have so far in this book discussed only the
path integrals and generating functions for scalar fields. All this formalism
can also be easily generalized to vector fields which obey the Proca equation
(4.26) and thus fulfill component by component the Klein–Gordon equation.
For fermion fields, however, there is a problem. The main idea in using path
integrals is to express quantum mechanical transition amplitudes by integrals
over classical fields; the values of these fields at the discrete coordinate-sites
were taken to be commuting numbers. Such a formalism can, however, not
“know” about the Pauli principle. For example, with the formalism developed
so far, a fermion could be propagated to a point in configuration space which
is already occupied. In nature, however, this propagation is Pauli-forbidden.
For the description of fermions it is, therefore, necessary to extend the
theory developed so far such that the Pauli principle is taken into account.
This can be achieved by using an anticommuting algebra for the classical
fields, the so-called Grassmann algebra.

10.1 Grassmann Algebra


In this section we outline the basic mathematical properties of the Grassman
algebra as far as we will need them in the later developments.
We define the n generators i , . . . , n of an n-dimensional Grassmann al-
gebra by the anticommutation relations

{i , j } ≡ i j + j i = 0 . (10.1)

Let us now consider series expansions in these variables. Since

21 = 22 = . . . = 2n = 0 , (10.2)

because of (10.1), any series in i must have the form


 
φ() = φ0 + φ1 (i)i + φ2 (i, j)i j (10.3)
i i<j

+ φ3 (i, j, k)i j k + . . . + φn (1, 2, . . . n)1 2 . . . n .
i<j<k

U. Mosel, Path Integrals in Field Theory


© Springer-Verlag Berlin Heidelberg 2004
126 10 Green’s Functions for Fermions

Here the φi are ordinary, commuting c-numbers. Note that the expansion in
a space of n Grassmann numbers actually terminates at the n-th order term
because of (10.2). Every element of the algebra can be written in the form
(10.3); thus, the various powers of  in this expansion constitute a basis of
the algebra.
If we choose the function φi to be totally antisymmetric in all their vari-
ables, then we can also write
 1 
φ() = φ0 + φ1 (i)i + φ2 (i, j)i j
i
2! ij
1 
+ φ3 (i, j, k)i j k + . . . . (10.4)
3!
ijk

From now on we can assume that the φi are indeed antisymmetric since any
symmetric part of φ would not contribute anyway.
As a consequence of this relation any analytical function in 1 dimension
has the form
φ(1) () = φ0 + φ1  , (10.5)
and in 2 dimensions

φ(2) () = φ0 + φ1 (1)1 + φ1 (2)2 + φ2 (1, 2)1 2 . (10.6)

These relations imply for the Taylor expansion of a Gaussian

e− = 1
2
(10.7)

and
e−i j = 1 − i j . (10.8)

10.1.1 Derivatives

Since functions of Grassmann variables can, because of (10.2), be at most


linear in ε, the derivative operator is defined as an algebraic operation that,
when applied to a Grassmann variable, simply replaces that variable by a 1;
when applied to a product the variable first has to be commuted to a position
next to the derivative before it can be removed. We then get, for example,
 
∂ ∂ ∂1
, i 1 = (i 1 ) + i
∂j ∂j ∂j
∂i ∂1 ∂1
= 1 − i + i = δij 1 . (10.9)
∂j ∂j ∂j

The “−” sign in the second line appears, because the factor 1 must be first
brought to the left before the derivative can act. This result is valid in general.
We thus can define the derivative also by the anticommutation relation
10.1 Grassmann Algebra 127
 

, i = δij . (10.10)
∂j

This relation can be used to obtain the derivative of a function φ().


Applying the lhs of (10.10) to φ, which can always be written as polynomial
(see (10.4)), yields
 
∂ ∂ ∂
, i φ() = [i φ()] + i φ() (10.11)
∂j ∂j ∂j
 
∂  1 
= δij φ() + φ0 − φ1 (k)k + φ2 (k, l)k l
∂j 2!
k kl

1  ∂
− φ3 (k, l, m)k l m + . . . i + i φ() .
3! ∂j
klm

Here all the terms in φ() with an odd number of ’s have changed sign when
they were brought to the left next to the derivative operator. With

(k l ) = δkj l − δlj k (10.12)
∂j
and

(k l m ) = δkj l m − δlj k m + δmj k l (10.13)
∂j
we get
  
∂ 1  1 
, i φ() = δij φ() + −φ1 (j) + φ2 (j, l)l − φ2 (k, j)k
∂j 2! 2!
l k
1  1 
− φ3 (j, l, m)l m + φ3 (k, j, m)k m
3! 3!
lm km

1  ∂
− φ3 (k, l, j)k l + . . . i + i φ( )
3! ∂j
kl
= δij φ()
 
 1 
− φ1 (j) − φ2 (j, k)k + φ3 (j, k, l)k l − . . . i
2!
k kl

+ i φ( ) . (10.14)
∂j
The last step is possible because all the φi are totally antisymmetric under ex-
change of their arguments. Because of (10.10) this has to be equal to δij φ();
the last two terms in (10.14) thus have to cancel each other. Commuting i
through to the left side of the square parentheses changes again the signs of
all odd terms so that we have
128 10 Green’s Functions for Fermions

∂  1 
φ() = φ1 (j) + φ2 (j, k)k + φ3 (j, k, l)k l + . . . . (10.15)
∂j 2!
k kl

This gives the derivative of a general function φ(). Equation (10.15) could
also have been obtained by differentiating the expansion (10.4) directly.
The second derivatives can also be defined
∂ ∂
φ() = φ2 (j, i)
∂i ∂j
1  1 
+ φ3 (j, i, l)l − φ3 (j, k, i)k + . . .
2! 2!
l k

= φ2 (j, i) + φ3 (j, i, l)l + . . . . (10.16)
l

Because the φi are antisymmetric this gives immediately

∂2 ∂2
φ() + φ() = 0 , (10.17)
∂i ∂j ∂j ∂i
or  
∂ ∂
, =0. (10.18)
∂i ∂j
Thus we have, in particular,
∂2
=0. (10.19)
∂2i
This relation implies that there is no inverse to the derivative. This can
be seen by multiplying the defining equation for the inverse
 −1
∂ ∂
φ() = φ() (10.20)
∂ε ∂ε

from the left by ∂ε . This gives
 −1  −1
∂ ∂ ∂ ∂ ∂
φ() = 0 · φ() = φ() (10.21)
∂ε ∂ε ∂ε ∂ε ∂ε

and thus the inverse does not exist.

10.1.2 Integration

Because an inverse to the derivative does not exist integration in the space
of Grassmann variables can be defined only in an operational sense. This is
achieved by the following relations
10.1 Grassmann Algebra 129

di = 0

di i = 1 . (10.22)

The symbols di obey the commutation relations

{di , dk } = {di , k } = 0 , (10.23)

but note that d is not an infinitesimal interval and, in particular, not a


member of the Grassmann algebra of the i .
Integration over Grassmann variables has thus the same effect as dif-
ferentiation. The integral in (10.22) is the only nonvanishing integral over
functions of i . It has the property of translational invariance
 
d φ(1) () = d φ(1) ( + α) = φ1 (10.24)

with the definition of φ(1) in (10.5); α is also a Grassmann number.


For Grassmann variables we can also define a δ-function. In 1 dimension
we have
 
− d ( −  )φ( ) = − d ( −  )(φ0 + φ1  )
  

= φ0 + φ1  = φ() . (10.25)

We can thus identify the δ function as

δ( −  ) = −( −  ) ; (10.26)

note that δ( −  ) is an odd function.


There is also a Fourier-representation for the δ-function
 

δ( −  ) = −( −  ) = dζ [1 − ζ( −  )] = dζ e−ζ(− ) ;
  
(10.27)

here (10.8) has been used.

10.1.2.1 Multiple Integrals In multiple integrals the integration variable


first has to be commuted to a position next to the integration measure
    
d1 d2 1 2 = − d1 d2 2 1 = − d1 1 = −1 . (10.28)

We now determine the Jacobian J that appears when we perform a linear


transformation
 = O (10.29)
of the integration variables in n dimensions; O is a general matrix. J is defined
by
130 10 Green’s Functions for Fermions
 
d1 . . . dn 1 . . . n = d1 . . . dn J (O)1 (O)2 . . . (O)n (10.30)

= d1 . . . dn J O1α O2β . . . Onν α β . . . ν .

The product of the Grassmann variables α , . . . , ν vanishes, if any of the


indices appears twice. Thus only the permutations of 1, . . . , n survive. We
can therefore bring each of them into the ordered form by writing

α β . . . ν = (−)P 1 . . . n , (10.31)

where (−)P is the sign of the permutation. We thus have


  
   
d1 . . . dn 1 . . . n = d1 . . . dn J (−)P O1α . . . Onν 1 . . . n
P

= d1 . . . dn J det(O) 1 . . . n . (10.32)

On the other hand, we can also evaluate the integrals over the ’s directly.
This gives  
d1 . . . dn 1 . . . n = d1 . . . dn 1 . . . n , (10.33)

because of (10.22). Comparison with (10.33) yields

J = [det(O)]−1 . (10.34)

Note that this is the inverse of the ordinary result!

10.1.2.2 Gaussian Integration For an exponential integral we get


 
di dj e−i j = di dj (1 − i j )
   
= 0 − di dj i j = + di dj j i

= di i = 1 . (10.35)

We next consider the more general Gaussian integral



T
I(n) = d1 . . . dn e− M  (10.36)

where M is an antisymmetric (n×n) matrix with ordinary c-number elements


and  is a column vector with n elements. Expanding the exponential function
we obtain
T 1  T 2
e− M  = 1 − T M  +  M − · · · . (10.37)
2!
10.1 Grassmann Algebra 131

For simplicity let us consider the case n = 2 first. We have


 
T  
I(2) = d1 d2 e− M  = d1 d2 1 − T M 

= 0 − d1 d2 T M  (10.38)

All higher order terms in the expansion of the exponential vanish because
they necessarily lead to higher-order-powers of  which vanish (cf. (10.2)).
For example, for the second order term in the expansion we get

T
 M  = d1 d2 i Mij j k Mkl l . (10.39)

The indices i, j, k and l all have to be different, because otherwise the prod-
uct of the  s vanishes. This, however, is impossible because we have only 2
different variables; the integral I2 (2) thus vanishes. The same holds for all
higher order terms in (10.38).
For the integral we now obtain
 
T
I(2) = − d1 d2  M  = − d1 d2 i Mij j

= − d1 d2 (1 M11 1 + 1 M12 2 + 2 M21 1 + 2 M22 2 )

= − d1 d2 (1 M12 2 + 2 M21 1 ) , (10.40)

because the integrands of the other two terms contain squares of Grassmann
variables which vanish. We therefore have (because M is antisymmetric)
2
I(2) = M12 − M21 = 2M12 = 2 2 det(M ) . (10.41)

For 3 variables we obtain immediately



T
d1 d2 d3 e− M  = 0 , (10.42)

because the exponent is quadratic in  and, therefore, the series expansion


involves only even powers. The term of second order in  gives 0, because of
d = 0; the term of fourth order also vanishes, because, with only three
variables, it has to contain a square of one of them, which vanishes.
A similar reasoning leads one to the conclusion that with four variables
only the term of fourth order in  can contribute. Thus we have

T
I(4) = d1 d2 d3 d4 e− M 

1  2
= d1 . . . d4 T M 
2
132 10 Green’s Functions for Fermions

1
= d1 . . . d4 i Mij j k Mkl l
2

1 
4
= Mij Mkl d1 . . . d4 i j k l . (10.43)
2
ijkl=1

Here only the terms with all four indices different can contribute. We thus get
4! = 24 nonvanishing terms. Of these 24 terms many are equal to each other:
the antisymmetry Mij = −Mji gives Mij Mkl = Mji Mlk and Mij Mlk =
−Mij Mkl and thus a factor 22 , the symmetry Mij Mkl = Mkl Mij another
factor 2, so that only 3 essentially different terms remain. We then have

I(4) = 4M12 M34 d1 . . . d4 1 2 3 4

+ 4M13 M24 d1 . . . d4 1 3 2 4

+ 4M14 M23 d1 . . . d4 1 4 2 3

= 4 (M12 M34 − M13 M24 + M14 M23 )


4
= 2 2 det(M ) (10.44)

The last step can be seen by explicitly expanding the (4 × 4) determinant of


M in terms of its elements in the first row.
The result just obtained holds in general
  n
−T M  2 2 det(M ) n even
d1 . . . dn e = for . (10.45)
0 n odd

The proof is given in Appendix D. Equation (10.45) is the equivalent of


(B.13) for commuting numbers. Note that here the squareroot of the deter-
minant appears in the numerator whereas for commuting numbers it is in the
denominator.
The other integral relations for commuting variables, derived in Appendix
B.2.1, can also be obtained for Grassmann variables. One gets
 
1 T 1 T −1
d1 . . . dn e− 2  M +ηi i = det(M ) e 2 η M η , (10.46)

where the ηi are also Grassmann variables with

{ηi , ηj } = 0 = {ηi , j } (10.47)

(there is no factor 2(2n)/2 here because the exponent contains an extra fac-
tor 1/2). Equation (10.46) corresponds to (6.27) for scalar fields; again the
determinant appears here in the numerator.
10.1 Grassmann Algebra 133

Complex integration. We now enlarge the algebra of the n by adding the


conjugate elements ∗n . Here conjugation is defined by the properties

(i )∗ = ∗i
(∗i )∗ = i
(λi )∗ = λ∗ ∗i
(1 2 · · · n )∗ = ∗n ∗n−1 · · · ∗1 . (10.48)

The latter condition ensures that ∗  is real. The n Grassmann generators


∗1 , . . . , ∗n are then combined with the n generators 1 , . . . , n into a 2n-
dimensional algebra with generators κ1 , . . . , κ2n such that

{κ1 , . . . , κ2n } = {∗1 , 1 , ∗2 , 2 , . . . , ∗n , n } . (10.49)

All the κi anticommute with each other.


With this definition we evaluate a Gaussian integral containing also con-
jugate Grassmann numbers


Z ∼ d∗1 d1 d∗2 d2 . . . d∗n dn e− M  . (10.50)

The matrix M in (10.50) can now be a general, complex matrix without


any specific symmetries. In contrast to the c-number integrals discussed in
Appendix B.2.1 it does not have to have special properties since convergence
of the integrals is guaranteed here by the Grassmann algebra. In order to be
able to apply the results obtained above (see (10.45)) to this integral we now
enlarge the general (n × n)-dimensional matrix M into an antisymmetric
(2n × 2n)-dimensional matrix N , so that the exponent in (10.50) can be
rewritten
n
 1 
2n
1
† M  = ∗i Mij j = κk Nkl κl = κT N κ . (10.51)
i,j=1
2 2
k,l=1

For example, for the case n = 2 the matrix N reads


⎛ ⎞
0 M11 0 M12
⎜ −M11 0 −M21 0 ⎟
N =⎜ ⎝ 0 M21 0 M22 ⎠ .
⎟ (10.52)
−M12 0 −M22 0

One then gets


2
det(N ) = [det(M )] , (10.53)
a result that is valid in any dimension. Using now (10.45) gives
 
1 T
Z ∼ dκ1 . . . dκ2n e− 2 κ N κ = det(N ) , (10.54)
134 10 Green’s Functions for Fermions

and with (10.53) we finally get




Z ∼ d∗1 d1 . . . d∗n dn e− M  = det(M ) . (10.55)

This result can also easily be obtained for a diagonal matrix M . Even though
the integral is a 2n-dimensional one, the matrix M on the rhs is only n-
dimensional. Note again that the matrix M here is a general complex matrix.
A slightly more general form follows when we add linear terms in the
exponent and write an expressively imaginary M . Then we obtain
 
† † † † −1
d∗n dn ei M +iη +i η = det(−iM ) e−iη M η . (10.56)
n

In all these formulas for Grassmann integrals the determinant appears in


the numerator in contrast to the Gaussian integrals over ordinary numbers
(B.22).

10.2 Green’s Functions for Fermions


In this section we now apply the Grassmann algebra to a path integral de-
scription of fermions. We do so by introducing classical Grassmann fields Ψ
for the fermions. These appear in the continuum limit when the number of
generators n of the Grassmann algebra goes to infinity. Remembering that in
field theory the spatial coordinates assume the role of the coordinate indices
the relations corresponding to (10.1),(10.10) and (10.22) are, respectively,

{Ψ (x) , Ψ (y)} = 0
 
δ
, Ψ (y) = δ(x − y)
δΨ (x

dΨ (x) = 0

dΨ (x) Ψ (x) = 1 . (10.57)

With these fields we can construct path integrals of the type (10.55)
 
Z ∼ DΨ̄ DΨ ei Ψ̄ M Ψ d x
4
(10.58)

where the Ψ̄ are Grassmann fields conjugate to Ψ .

10.2.1 Generating Functional for Fermions

The Lagrangian for Dirac fields is


10.2 Green’s Functions for Fermions 135

L = Ψ̄ (iγμ ∂ μ − m) Ψ − V (Ψ̄ , Ψ ) (10.59)

where Ψ and Ψ̄ are to be considered as independent fields. They are now


taken to be elements of an infinite-dimensional Grassmann algebra. The path
integral is then – in direct generalization of our results for boson fields – given
by   4
Z[η, η̄] = Z0 DΨ̄ DΨ ei d x [L+η̄(x)Ψ (x)+Ψ̄ (x)η(x)] . (10.60)

Here Z0 is the normalization factor, equal to the inverse of the path integral
without sources. The functions η(x) and η̄(x) are four-component source
functions corresponding to the classical fields Ψ̄ and Ψ , respectively. They
are also taken to be Grassmann fields that anticommute among themselves
as well as with the fields Ψ and Ψ̄

{Ψ (x), Ψ (x )} = {Ψ (x), Ψ̄ (x )} = {Ψ̄ (x), Ψ̄ (x )} = 0

{η(x), η(x )} = {η(x), η̄(x )} = {η̄(x), η̄(x )} = 0


{η(x), Ψ (x )} = {η(x), Ψ̄ (x )} = {η̄(x), Ψ (x )} = 0 . (10.61)
While (10.60) is an obvious generalization of our earlier results obtained
in Chap. 5 for scalar fields, its actual (numerical) evaluation may not seem to
be straightforward, because of the Grassmann nature of the fields. In order to
define more stringently what is actually meant by the path integral in (10.60)
we can proceed as in Sect. 5.1 for scalar fields and Fourier-expand the fields.
This is carried through in detail in [2]. For our present purpose it is enough
to mention that the fields Ψ and Ψ̄ can be expanded in terms of Grassmann
numbers α and ∗α such that

Ψ (x) = wα (x)α (t)
α

Ψ̄ (x) = w̄α (x)∗α (t) , (10.62)
α

where the wα are x-dependent Dirac spinors for positive and negative energies
(u(pα , sα )eipα ·x or v(pα , sα )e−ipα ·x , resp., see [2], p. 517-521).
Equation (10.62) just amounts to an x-dependent change of the basis from
α , ∗α to the Grassmann numbers Ψ and Ψ̄ . The fermionic action, e.g. for free
fields, then reads in terms of the Grassmann numbers (with k ≡ γμ k μ )

S = Ψ̄ (i ∂ − m) Ψ d4x

= w̄α ∗α (i ∂ − m) wβ β d4x , (10.63)
αβ

and the path-integral integration measure is really


136 10 Green’s Functions for Fermions

DΨ̄ DΨ = dΨ̄ (xk , tj ) dΨ (xk , tj )
k j

∼ d∗α (tj ) dα (tj ) . (10.64)
α j

Here the first index (α) denotes the Grassmann generator, the second (j) the
time interval. The last step here was possible because of the orthonormality
of the spinor-functions wα and w̄α (cf. (4.46)).

10.2.1.1 Feynman Propagator for Fermions In the case of free boson


fields we succeeded in writing the generating functional as a term involving
only space–time integrals over the sources (see (6.11) and (6.39)). This had
the advantage that the two-point function could be directly read off from
that expression.
We achieve here the same for fermions by using (10.56). This gives for the
generating functional for free fermions
  4
Z0 [η, η̄] = Z0 DΨ̄ DΨ ei d x(L+η̄Ψ +Ψ̄ η)
−1
= Z0 det(−iM ) e−iη̄M η
, (10.65)

where the matrix M is obtained by identifying



/ 0
i† M  ←→ i d4x Ψ̄ (x)(i ∂ − m)Ψ (x)

/ 0
= − i d4 x d4 y Ψ̄ (x) (i ∂y + m)δ 4 (x − y) Ψ (y) (10.66)

after partial integration. Thus M is given by

M (x, y) = − (i ∂ + m) δ 4 (x − y) . (10.67)

Since Z0 [0, 0] = 1 we have Z0 = [det(−iM )]−1 . We thus get


 −1
Z0 [η, η̄] = e−i η̄(x)M (x,y)η(y)d x d y .
4 4
(10.68)

The inverse operator appearing here can be obtained as at the end of


Sect. 6.1.3
 
/ 0
M (x, z)M −1 (z, y) d4z = − (i ∂ z + m) δ 4 (x − z) M −1 (z, y) d4z

= + δ 4 (x − z) (i ∂ z − m) M −1 (z, y) d4z

= (i ∂ x − m) M −1 (x, y) = δ 4 (x − y) .
!

(10.69)
10.2 Green’s Functions for Fermions 137

The last equation is fulfilled by

M −1 (x, y) = (i ∂ + m) DF (x − y) , (10.70)

since

(i ∂ − m)(i ∂ + m)DF (x − y) = − (2 + m2 )DF (x − y)


= δ 4 (x − y) , (10.71)

because of (6.7); here we have used {γμ , γν } = 2gμν (4.41).


We therefore have now the desired result. For free fields the normalized
generating functional is

Z0 [η, η̄] = e−i η̄(x)SF (x−y)η(y) d xd y ,
4 4
(10.72)

with the Feynman propagator for fermions

SF (x − y) = (i ∂ + m)DF (x − y) (10.73)

being the propagator of the Dirac equation

(i ∂ − m) SF (x − y) = δ 4 (x − y) . (10.74)

Note that SF is a (4 × 4) matrix because of the Dirac matrix structure of γμ .


For SF we can also find an integral representation

d4k e−ik(x−y)
SF (x − y) = (i ∂ + m)
(2π)4 k 2 − (m − i)2

d4k −ik(x−y) k + m
= e
(2π) 4 k − (m − i)2
2

d4k −ik(x−y) 1
= e (10.75)
(2π) 4 k − (m − i)

because of ( k + m)( k − m) = k 2 − m2 .
Note that SF (x − y) is not symmetric under exchange x ↔ y, while
DF (x − y) is. However, it is antisymmetric under an operation that corre-
sponds in Dirac theory to simultaneous x ↔ y exchange and hermitian con-
jugation since
γ0 SF† (y − x)γ0 = SF (x − y) , (10.76)
which can easily be derived with the help of the relation

γ0 γμ† γ0 = γμ . (10.77)

As we have discussed in Sect. 6.1.2 for scalar fields the Feynman propaga-
tor moves positive-energy solutions forward in time and those with negative
energy backwards. The same is true for the propagator for fermions as can
be shown following steps very similar to those leading to (6.20).
138 10 Green’s Functions for Fermions

10.2.2 Reduction Theorem for Fermions

For completeness and easier reference we give here the Reduction Theorem
(cf. Sect. 7.2) also for fermions. In this case the normal mode expansion of
the free in-field is – in analogy to (7.16) – given by

√ d3k   † ∗

Ψ (x) = 2m a ks u(k, s)fk (x) + b ks v(k, s)fk (x) . (10.78)
(2π)3 s

Here the functions fk are defined in (7.17) and u(k, s) and v(k, s) are eigen-
spinors of the free Dirac Hamiltonian to positive and negative eigenvalues,
respectively (4.43).
The further derivation of the reduction theorem for fermions then pro-
ceeds in complete analogy to the case of scalar fields in Sect. 7.2.2. The
result, the reduction theorem for fermions, looks more complicated than the
one for the uncharged, scalar field treated in Sect. 7.2.2 because there are
now also antiparticle degrees of freedom involved. We have thus in- and out-
operators for both particles and antiparticles. Equation (10.78) together with
(10.87) shows that an incoming fermion is in the propagator connected with
the spinor u(k), an outgoing one with the spinor ū(k). Correspondingly, an
incoming anti-fermion is connected with a spinor v̄(k), and an outgoing one
with v(k)1 .
The reduction theorem for fermions (with m particles with momenta k
and m̄ antiparticles with momenta q in the in state and n particles with
momenta k  and n̄ antiparticles with momenta q  in the out state) then reads

Sβα = β, n k  , n̄ q  out|α, m k, m̄ q in


  m  m̄   n 

m+m̄+n+n̄
=i 4 4
d xi d yı̄ d4xj d4yj̄
i=1 ı̄=1 j=1 j̄=1

× fk∗ (xj )fq∗j̄ (yj̄ ) fki (xi )fqı̄ (yı̄ )


j
 
× ū(kj ) i ∂xj − mj v̄(qı̄ ) (i ∂yı̄ − mı̄ )

ˆ (x ) · · · Ψ̄
× 0|T Ψ̂ (ym̄ ) . . . Ψ̂ (y1 )Ψ̂ (xm ) . . . Ψ̂ (x1 ) Ψ̄ ˆ (x )Ψ̄
ˆ (y  ) . . . Ψ̄
ˆ (y  ) |0
m 1 n̄ 1
   
← ←
× −i ∂ xi −mi u(ki ) −i ∂ yı̄ −mı̄ v(qj̄ ) . (10.79)

To simplify the notation the helicity quantum numbers of the spinors have
not been written out here.
Again, the structure of (10.79) is as for the scalar field case. The S-
matrix elements again contain a vacuum expectation value of a time-ordered
product of fermion field operators, i.e. a correlation or n-point function. The
Dirac operators remove the propagators on the external lines and replace
1
The ’bars’ over u and v arise since the two-point function contains the fields Ψ̄ .
10.2 Green’s Functions for Fermions 139

them with the eigenspinors u(k)fk (x) and v̄(q)fq (x) for the incoming particle
and antiparticle states, respectively, and ū(k  )fk∗ (x) and v(q  )fq∗ (x) for the
outgoing ones.

10.2.3 Green’s Functions

The n-point functions for fermions entering the S matrix can now be ob-
tained, as in the case of bosons, as functional derivatives of the functional
Z[J]. For this purpose we modify the definition (8.4) such that the presence
of two fermion fields, Ψ and Ψ̄ , is taken into account.
We define the 2n-point function for fermions by

Gα,β,...,2ν (x1 , x2 , . . . , x2n ) (10.80)


 2n 1
δ 2n Z 1
1 1
= 1 ,
i δη2ν (x2n ) · · · δην+1 (xn+1 )δ η̄ν (xn ) · · · δ η̄α (x1 ) 1
η=η̄=0

where Z[η, η̄] is given by (10.60)


 
d4x [L+η̄(x)Ψ (x)+Ψ̄ (x)η(x)]
Z[η, η̄] = Z0 DΨ̄ DΨ ei . (10.81)

Here the order of the functional derivatives with respect to η and η̄, respec-
tively, is fixed by convention (see discussion below). Note that the Green’s
function now depends also on the Dirac spinor-indices, in addition to the
space-time coordinates.
As in (8.2) for bosons, the Green’s function can be written as a vacuum
expectation value of a time-ordered product of field operators

G(x1 , x2 , . . . , x2n ) (10.82)



DΨ̄ DΨ Ψ (xn ) · · · Ψ (x1 )Ψ̄ (x2n ) · · · Ψ̄ (xn+1 ) eiS
= 
DΨ̄ DΨ eiS

ˆ (x ) · · · Ψ̄
= 0|T Ψ̂ (xn )Ψ̂ (xn−1 ) · · · Ψ̂ (x1 ) Ψ̄ ˆ (x
2n n+1 ) |0 ,

where S denotes the action. The special ordering of fields in (10.82) is a con-
sequence of our ordering of the derivatives in (10.80); it ensures that there is
no additional phase present. This can be seen as follows: In calculating the
derivatives of Z with respect to η̄ no phase appears, because first the deriva-
tive is anticommuted through all field operators already pulled down from
the exponential, then acts without a further sign change on the exponential
and then is anticommuted back, thus creating the same phase backwards as
on the way forward. The extra phase that appears from the derivative with
respect to η due to the necessary reordering in the exponent is cancelled by
the same phase caused by anticommuting the derivative operator δ/δη with
all the Grassmann fields Ψ and Ψ̄ , that were already pulled down.
140 10 Green’s Functions for Fermions

2-point function. The two-point function for free fermions is defined by


1
δ 2 Z[η, η̄] 11
Gαβ (x1 , x2 ) = − . (10.83)
δηβ (x2 )δ η̄α (x1 ) 1η=η̄=0

Using (10.65) for the generating functional of the free theory we obtain
  4
DΨ̄ DΨ Ψα (x1 )Ψ̄β (x2 )ei L d x
Gαβ (x1 , x2 ) =   (10.84)
DΨ̄ DΨ ei L d x
4

Here α and β are Dirac-Spinor indices.


This expression can, as for bosons, be written as an expectation value
over a time-ordered product of field operators

Gαβ (x1 , x2 ) = 0|T Ψ̂α (x1 )Ψ̄ˆβ (x2 ) |0 (10.85)

where, now for fermions, we have


*
 ˆ β(x )
Ψ̂α (x1 )Ψ̄
ˆ 2 t > t2
T Ψ̂α (x1 )Ψ̄ β (x2 ) = ˆ for 1 . (10.86)
−Ψ̄ β (x2 )Ψ̂α (x1 ) t 2 > t1

Note the extra minus sign compared to the boson case. This sign appears
because of the Grassmann nature of the fermion fields when we reorder the
fields as in the developments leading to (3.47).
For free fields the two-point function can be most easily obtained from
(10.72)
  −ik(x−y) 
d4k e
Gαβ (x1 , x2 ) = i[SF (x1 − x2 )]αβ = i . (10.87)
(2π) 4 k − (m − i) αβ

Again, as in scalar field theory (8.18), the 2-point function is – up to a phase


– just equal to the propagator of the relevant equation of motion.
The free propagator of our fermionic theory is thus just the inverse of the
operator appearing between the fields Ψ̄ and Ψ in the Lagrangian
L = Ψ̄ (i ∂ − m)Ψ
G(x, y) = i(i ∂ − m)−1 = iSF (x − y) . (10.88)
For boson fields this was also the case (see the discussion at the end of Sect.
6.1.3)
1
L = − φ(2 + m2 )φ
2
G(x, y) = − i(2 + m2 )−1 = iDF (x − y) . (10.89)
We can thus in general read off the propagator directly from the Lagrangian2 .
2
The factor 1/2 in the Lagrangian for the bosons is a special feature of the un-
charged field and would not be there if we were working with charged fields.
11 Interacting Fields

So far we have treated only the case of free fermion fields. In the case of an
interacting theory, expression (8.44) can directly be taken over for the case
of fermions so that we have
  
, i δη̄ −i  η̄(x)SF (x − y)η(y)d4x d4y
δ 1 δ
−i d4x V 1i δη
Z[η, η̄] = Z0 e e .
(11.1)
Here the interaction V is defined by the Lagrangian of the interacting theory

L = L0 − V (Ψ̄ , Ψ ) . (11.2)

If the theory involves both boson and fermion fields and their couplings,
i.e. if the Lagrangian is given by
φ
L = LΨ
0 + L0 − V1 (Ψ̄ , Ψ ) − V2 (φ) − Vint (Ψ̄ , Ψ, φ) , (11.3)

then Z[J, η, η̄] is given simply by (cf. (8.44))


  
δ 1 δ 1 δ
−i d4x Vint 1i δη , i δη̄ , i δJ
Z[J, η, η̄] = Z0 e
  
δ 1 δ   
−i d4x V1 1i δη , i δη̄ −i d4x V2 1 δ
×e e i δJ

×e−i η̄(x)SF (x − y)η(y)d x d y
4 4


×e− 2 J(x)DF (x − y)J(y)d x d y .
i 4 4
(11.4)

This functional can generate all the n-point functions of the interacting the-
ory and these can again be graphically represented in the form of Feynman
diagrams.

11.1 Feynman Rules

The Feynman rules of Sect. 9.1 can be extended by the following rules (in
momentum space)

U. Mosel, Path Integrals in Field Theory


© Springer-Verlag Berlin Heidelberg 2004
142 11 Interacting Fields

1) each fermion line gives a factor


p + m
SF (p) = =
p2 − (m − i)2

This rule follows directly from the 2-point function (10.87). The fermion lines
now carry an arrow that points into the direction into which the fermion’s
charge flows. This is a consequence of the fact that the fermion propagator
is no longer symmetric in its arguments (cf. (10.76)) so that we have to give
the direction of motion explicitly. For fermions we deal with two fields, ψ and
ψ̄, that carry different charges. In a Feynman graph, in which time runs from
left to right, particle lines are always rightwards directed whereas antiparticle
lines move leftwards. In other words, the fermion lines move from ψ̄ to ψ,
because ψ̄ creates particles and ψ annihilates them. Since the opposite holds
for antiparticles their lines run leftwards, reflecting the fact that the Feynman
propagator propagates them backwards in time.
In a coupled theory a new class of diagrams can appear because of the
coupling of fermion and boson fields. For example, a coupling term of the
form
Vint = g Ψ̄ Ψ φ = g Ψ̄α Ψα φ , (11.5)
the so-called Yukawa coupling, will generate vertices of the form shown in Fig.
11.1. The corresponding Feynman rule can be obtained by first calculating

Fig. 11.1. Fermion-boson vertex (see (11.5)). The solid line denotes the fermion,
the dashed one the boson. The dot denotes the interaction vertex

the generating functional (11.4) in lowest order in the coupling g, and then
the 3-point function. This gives as a new Feynman rule dealing with fermion-
boson vertices
2) each boson–fermion vertex of Yukawa theory gives a factor −ig.
Another important example is the coupling of a fermion field to a vector
field through the fermionic current (4.72)

Vint = g Ψ̄ γμ Ψ Aμ . (11.6)
11.1 Feynman Rules 143

Such a coupling appears in QED (4.69) and, more generally, in all gauge field
theories to be treated later in this book. In this case the Feynman rule reads
3) each fermion–vector vertex gives a factor −igγμ .

11.1.1 Fermion Loops

A change in the rules appears when we consider – again for simplicity in


Yukawa theory – fermion loop insertions in the boson propagator, e.g. the
graph shown in Fig. 11.2, where the solid line denotes the fermions and
the dashed line the bosons. Such a term is obviously of second order in the
fermion–boson coupling constant (indicated by the two vertices) and con-
tributes to the bosonic 2-point function. It can thus be generated by the
second order term in the expansion of (11.4), properly generalized to fermion
fields. Because the graph in Fig. 11.2 is one-particle irreducible only the true
second order term (∼ S2 in (8.51),(8.52)) can contribute. This term has the
structure

iS2 [η, η̄, J] (11.7)


(   )2
1 −iS0 [η,η̄,J] 1 δ 1 δ 1 δ
= e −i d4x Vint , , e+iS0 [η,η̄,J] .
2 i δη(x) i δ η̄(x) i δJ(x)

Here S0 [η, η̄, J] is given by



i
iS0 [η, η̄, J] = − d4v d4w J(v)DF (v − w)J(w)
2

− i d4x d4y  η̄(x )SF (x − y  )η(y  ) . (11.8)

Because the graph in Fig. 11.2 has only bosonic, but no fermionic external
legs only those terms in the generating functional can contribute to the boson
propagator in Fig. 11.2 that are of second order in the bosonic current J
and of zeroth order in the fermionic currents η and η̄. Therefore, the only
contributing term in (11.7) with the interaction (11.5) is

(x α) (y β)

Fig. 11.2. Fermion loop contribution to the boson propagator


144 11 Interacting Fields

(−ig)2
iS2loop = −
2

δ6
× e−iS0 d4x d4y e+iS0
δηα (x)δ η̄α (x)δJ(x)δηβ (y)δ η̄β (y)δJ(y)
 
(−ig)2 −iS0 4 4
= e d xd y d4v d4w (11.9)
2
( )
δ4
× DF (x − v)J(v)DF (y − w)J(w) e+iS0 ,
δηα (x)δ η̄α (x)δηβ (y)δ η̄β (y)

after performing the derivative with respect to J.


We now concentrate on the fermionic derivative (for notational conve-
nience we drop here the index F for the propagators)

δ3 δ     4  4 
e−i η̄(x )S(x −y )η(y ) d x d y
δηα (x)δ η̄α (x)δηβ (y) δ η̄β (y)
 
δ2 δ
= (−i) Sβγ (y − y  )ηγ (y  ) d4y  e−i ···
δηα (x)δ η̄α (x) δηβ (y)

δ δ
= − iSββ (y − y)
δηα (x) δ η̄α (x)
    

− Sβγ (y − y )ηγ (y ) d y  4 
η̄δ (x )Sδβ (x − y) d x e−i ···
  4 

 
δ
= − Sββ (y − y) Sαγ (x − y  )ηγ (y  ) d4y 
δηα (x)
   
⊕ Sβγ (y − y )ηγ (y ) d y Sαβ (x − y) + O(η η̄) e−i ··· . (11.10)
  4  2

Here the symbol ⊕ denotes a ‘plus’ sign that is due to the Grassmann nature
of the source functions and would have been opposite, if we had worked
with bosons instead of fermions. The terms denoted by O(η 2 η̄) stand for
expressions that involve products of the two source functions.
Performing now the last derivative gives for the expression (11.10)

= [−Sββ (y − y)Sαα (x − x) ⊕ Sβα (y − x)Sαβ (x − y) + O(η η̄)]



η̄(x )SF (x −y  )η(y  ) d4x d4y 
× e−i . (11.11)

The sign of the second term would have been opposite, if we had worked with
bosons.
We now combine this result with (11.9). Since we are interested in partic-
ular in the fermion loop insertion to the boson propagator shown above, we
need to construct a bosonic 2-point function with no external fermion lines.
11.2 Wick’s Theorem 145

Since the n-point function is given by a functional derivative at a vanishing


source, these terms do not contribute to the 2-point function we are after.
Therefore, only those terms of the fermionic part (11.11) can contribute to
this graph that contain no sources η or η̄, except in the exponential. All con-
tributions from terms ∼ η η̄ or higher powers vanish after setting η = η̄ = 0.
Disregarding the vacuum contributions (first term in (11.11)) we thus
have as the only term of (11.11) that contributes to the loop diagram

⊕ tr [SF (y − x)SF (x − y)] e−i ··· . (11.12)

The trace here is one over the Dirac indices of SF . Inserting this result into
(11.9) gives for the loop contribution of S2

(−ig)2
iS2loop = ⊕ d4x d4y d4v d4w tr [SF (y − x)SF (x − y)]
2

× DF (x − v)DF (y − w)J(v)J(w)e−i ··· (11.13)

and, correspondingly, for the 2-point function

δ 2 (iS2loop )
Gloop
boson (x1 , x2 ) = − | (11.14)
δJ(x1 )δJ(x2 ) J=0,η=η̄=0

=  (−ig)2 d4x d4y DF (x1 − x)tr [SF (x − y)SF (y − x)] DF (y − x2 ) .

When we retrace this detailed calculation we find that the positive sign ⊕ in
(11.12) and thus the negative sign  in (11.14) is due to the fact that η and
η̄ are Grassmann fields.
Thus, we get the additional Feynman rule
4) Any fermionic loop graph is associated with an additional (−) sign and a
trace over the Dirac indices. The trace is due to the fact that the scalar
particle that the fermion couples to here carries no Dirac index.

11.2 Wick’s Theorem


In Sect. 8.2.1 we had expressed the n-point function for bosons G(x1 , . . . , xn )
in terms of a symmetrized product of n/2 2-point functions (Wick’s theorem).
The same can now be done also for the general theory that involves bosons
and fermions.

Mixed n-point functions. Boson and fermion fields commute because they
describe distinguishable particles; the corresponding field operators act in
different Hilbert spaces so that the groundstate expectation value separates
146 11 Interacting Fields

into a product of vacuum expectation values over fermion and boson operators
separately, e.g.
ˆ (4)|0 = 0|Ψ̂ (1)Ψ̄
0|Ψ̂ (1)φ̂(2)φ̂(3)Ψ̄ ˆ (4)|00|φ̂(2)φ̂(3)|0 . (11.15)
We, therefore, need to formulate here Wick’s theorem only for the fermions.
Wick’s theorem for fermions. We now derive a simple method, known as
Wick’s Theorem, that we already encountered for bosons in Sect. 8.2.1, for
the evaluation of a vacuum expectation value of a time-ordered product of
free field operators. In that case we can use the form (10.68) of the generating
functional for the evaluation of the functional derivative. Performing first the
functional derivatives with respect to η̄ on that expression gives1

δ 2n
e−i η̄(x)SF (x−y)η(y) d x d y |η=η̄=0
4 4

δη(x2n ) · · · δη(xn+1 )δ η̄(xn ) · · · δ η̄(x1 )


 
δn n
= (−i) p SF (xn − y)η(y) d4y
δη(x2n ) · · · δη(xn+1 )
 
× SF (xn−1 − y)η(y) d y × · · · × SF (x1 − y)η(y) d4y
4

 1
1
−i η̄(x)SF (x−y)η(y) d4x d4y 1
×e 1 . (11.16)
1
n=η̄=0

Here p is a phase due to the commutation of the derivative operator with the
source functions already pulled down from the exponent. Now the differen-
tiation with respect to η is performed. Since the whole expression has to be
taken at η = η̄ = 0, only the prefactors of the exponential can contribute.
This gives

= (−i)n (−)P SF (x1 − xp2n )SF (x2 − xp2n−1 ) . . . SF (xn − xpn+1 ) , (11.17)
P

where the sum over P runs over all permutations of the indices n + 1, . . . , 2n.
The sign of the permutations is such that (−)P = +, if always the outermost
operators in (10.80) are combined into the 2-point function, i.e. if (1, 2n),
(2, 2n − 1) · · · (n, n + 1) are combined. The Dirac indices, that have not been
written down here, have to be combined in the same way.
Inserting this result into the definition of the Green’s function in (10.80)
gives
 2n
1 δ 2n Z
G(x1 , x2 , . . . , x2n ) =
i δη(x2n ) . . . δη(xn+1 )δ η̄(xn ) . . . δ η̄(x1 )

= in (−)PSF (x1 −xp2n ) · · · SF (xn −xpn+1 ). (11.18)
P
1
To facilitate the notation the Dirac indices are not written down here.
11.3 Bosonization of Yukawa Theory 147

Equation (11.18) is Wick’s theorem for fermions. Written as a vacuum ex-


pectation value of time-ordered field operators it reads

ˆ (x ) · · · Ψ̄
0|T Ψ̂ (xn ) · · · Ψ̂ (x1 ) Ψ̄ ˆ (x
2n n+1 |0
  
= in (−)P 0|T Ψ̂ (x1 )Ψ̄ ˆ (x ) |0 · · · 0|T Ψ̂ (x )Ψ̄
ˆ (x
p2n n pn+1 ) |0 .
P
(11.19)

Wick’s theorem for fermions (11.18) has a form that is analogous to that for
bosons (8.21). The essential difference is the appearance of the sign of the
permutation that reflects the antisymmetric of fermionic states. In addition
the degeneracy factor 1/2n is missing here because for fermions we have no
longer a symmetry under exchange: SF (x − y) = SF (y − x) (see (10.76)).
This form also exhibits clearly the important consequence of using anti-
commuting Grassmann fields for the fermions. The exchange of any two of
the coordinates appearing in the fields Ψ or of any two in the fields Ψ̄ gives a
“−” sign to the Green’s function. For example, the 2-particle, 4-point Green’s
function has the property

Gαβγδ (x1 , x2 , y1 , y2 ) = −Gβαγδ (x2 , x1 , y1 , y2 ) = −Gαβδγ (x1 , x2 , y2 , y1 )


= Gβαδγ (x2 , x1 , y2 , y1 ) . (11.20)

This implies that if all the Dirac-indices are the same we must have x1 = x2
and y1 = y2 for G not to vanish, thus reflecting the Pauli principle.

11.3 Bosonization of Yukawa Theory


In this section we illustrate a widely used method to transform a Lagrangian
that involves different kinds of fields into one in which only one field appears
explicitly.
We use as an example of this technique a coupled fermion–boson theory
with a simple Yukawa coupling. Its Lagrangian reads
1
L = − φ2φ − V (φ) + ψ̄ (i ∂ − mF ) ψ − g ψ̄ψφ
2
1
= − φ2φ − V (φ) + ψ̄ (i ∂ − mF − gφ) ψ . (11.21)
2
The term V (φ) may contain both the mass term and self-interactions up to
terms of order O(φ4 ). Since the coupling constant g here is dimensionless we
expect that this theory is renormalizable (cf. the discussions in Sect. 9.4.2).
The second line of (11.21) shows that the Yukawa coupling contributes to
148 11 Interacting Fields

the fermion mass and, for mF = 0, can even generate the entire mass of a
fermion. This mechanism plays an important role in the theory of electroweak
interactions, to be discussed in Chap. 14.
The generating functional of this theory is given by
 (  )
  4
Z[η, η̄, J] = Z0 Dψ̄ Dψ Dφ exp i L[ψ̄, ψ, φ] + Jφ + η̄ψ + ψ̄η d x .
(11.22)
If we are interested in processes that do not contain external, physical
fermions then it is advantageous to rewrite Z such that the fermion fields no
longer appear explicitly. This can indeed be achieved because the Lagrangian
(11.21) is quadratic in the fermion fields and the integrand is thus of a Gaus-
sian form. The path integral over the fermionic degrees of freedom can thus
be performed. According to (10.67) this gives
  1
−1
Z[η, η̄, J] = Z0 Dφ det(M ) e−iη̄M (φ)η e−i [ 2 φ2φ+V (φ)−Jφ] d x (11.23)
4

with (see (10.66))

M (φ) = i ∂ − m∗ (φ) and with m∗ (φ) = mF + gφ ; (11.24)

the factor ‘i’ before M in the determinant has been absorbed into the nor-
malization constant Z0 . Note that the fermion determinant M is a function
of φ. Even though (11.23) no longer contains a path integration over fermion
fields it is, nevertheless, completely equivalent to (11.22).
We now concentrate on processes that contain only bosons on the external
legs and thus require only bosonic n-point functions for the calculation of
transition rates etc. We can then set η and η̄ to zero. In doing so we give
up the possibility of calculating any n-point function with fermionic external
legs because these are obtained as functional derivatives with respect to η.
However, through the presence of the fermion determinant M we do keep the
effects of fermionic loops in our theory for the scalar field φ.
Using now (6.55)
det [M (φ)] = etr ln M (φ) (11.25)
gives
  1
Dφ e−i [ 2 φ2φ+V (φ)−Jφ] d x+tr ln M (φ) .
4
Z[J] = Z0 (11.26)

Equation (11.26) shows that now formally all the fermion degrees of freedom
have been removed. Alternatively, we could also say that we are now working
with a new, effective action for the bosonic sector
  
1
S̃B = − φ2φ + V (φ) d4x − i tr ln (i ∂ − m∗ (φ)) . (11.27)
2
11.3 Bosonization of Yukawa Theory 149

The trace here has to be performed over Dirac indices and over space–time co-
ordinates (or, equivalently, in momentum space). The corresponding effective
Lagrangian density reads
1
LB = − φ2φ − V (φ) − i Tr ln (i ∂ − m∗ (φ)) , (11.28)
2
where we have used 
tr = Tr d4 x ; (11.29)

Tr is a trace over the Dirac (and possibly other internal) indices. In this form,
which is still exact, the original Lagrangian (11.21) is said to be bosonized.
The true groundstate (vacuum) of the bosonic sector is, of course, deter-
mined by the potential V (φ), but it may not coincide with the minimum of
this potential. Nevertheless, we may take here the constant, translationally
invariant field φ0 which minimizes V as a reference field around which we
can perform a perturbative expansion. In order to normalize the action, and
thus the physics, to this state we subtract the action corresponding to φ0 .
This gives (with m∗0 = mF + gφ0 )
  
1
SB = − φ2φ − V (φ) d4x − i tr ln (i ∂ − m∗ ) + i tr ln (i ∂ − m∗0 ) .
2
(11.30)
Next we use the identity

tr ln (i ∂ − m∗ ) = tr ln (−i ∂ − m∗ ) (11.31)

which holds because of the symmetry of the eigenvalue spectrum of the oper-
ator and the fact that the trace runs over all eigenvalues; here we remember
from our discussion in Sect. 6.1.3 that the determinant of an operator is just
the product of its eigenvalues. We thus get
1
tr ln (i ∂ − m∗ ) =tr ln [(i ∂ − m∗ ) (−i ∂ − m∗ )]
2
1  
= tr ln 2 + m∗ 2 − ig ∂φ . (11.32)
2
Thus the effective bosonic action becomes
  
1
SB = − φ2φ + V (φ) d4x (11.33)
2
i   i  
− tr ln 2 + m∗ 2 − ig ∂φ + tr ln 2 + m∗0 2
2 2
    
1 i m∗ 2 − m∗0 2 − ig ∂φ
= − φ2φ + V (φ) d x − tr ln 1 +
4
2 2 2 + m∗0 2

and the Lagrangian density now reads


150 11 Interacting Fields
 
1 i m∗ 2 − m∗0 2 − ig ∂φ
LB = − φ2φ − V (φ) − Tr ln 1 −
2 2 −2 − m∗0 2
1 i
= − φ2φ − V (φ) − Tr ln (1 − DW ) (11.34)
2 2
with the operator D and the field-dependent quantity W
1
D=− and W = m∗ 2 − m∗0 2 − ig ∂φ . (11.35)
2 + m∗0 2
Note that D is nothing else than the Feynman propagator (cf. (6.7)) for a
field with a free mass m∗0 = mF − gφ0 . So far, no approximations have been
made and the theory is still exact. Notice also the similarity of the extra term
in (11.34) to the effective potential (6.61) derived in Sect. 6.2.1.

11.3.1 Perturbative Expansion


We now develop a perturbative treatment of this bosonized Lagrangian that
uses an expansion around the reference field φ of the bosonic sector.
For that purpose we expand the logarithm (− ln(1 − x) = x + x2 /2 + . . .)
in (11.34) and obtain
∞
1 n
−Tr ln(1 − DW ) = Tr (DW ) (11.36)
n=1
n

The trace in (11.36) indicates that each term in the expansion contains one
closed fermion loop (with n vertices W coupling to the field φ). Equation
(11.36) is therefore called the one-loop expansion, it represents a perturba-
tive expansion of the effective action. Integrating out the fermion degrees
of freedom has thus led to an effective theory that takes all fermionic one-
loop diagrams into account. The expansion is expected to converge the more
rapidly the closer the actual field φ is to the chosen reference field φ0 and the
smaller the gradients of φ are.
Our aim in the following developments is to write the trace term as an
integral over a space–time density which could then be interpreted as a cor-
rection term to the Lagrangian density, thus yielding an effective Lagrangian
which contains the effects of the fermions on the one-loop level.
In order to be able to do so we have to separate the terms under the trace
into a product of operators that are local in x and p. In this case the trace
can be separated

d4p
tr [F (p̂)G(x̂)] = Tr p|F (p̂)G(x̂)|p
(2π)4
   
d4p d4p
= Tr d 4
x d4x
(2π)4 (2π)4
×p|F (p̂)|p p |x x |G(x̂)|xx|p
 
d4p
= Tr F (p) d4x G(x) . (11.37)
(2π)4
11.3 Bosonization of Yukawa Theory 151

The expressions in (11.36) are, however, not in a form that allows direct
application of (11.37). While W is diagonal in x and D in p, these two op-
erators are mixed in the higher order terms in (11.36). We thus first reorder
each term such that all propagators are moved to the left. To do so we start
from the identity / 0
W D = DW + D D−1 , W D (11.38)
/ 0
and iterate this equality by applying it again to the quantity W  = D−1 , W
/ 0
W  D = DW  + D D−1 , W  D . (11.39)

Combining now (11.38) and (11.39) gives


/ 0 / / 00
W D = DW + D2 D−1 , W + D2 D−1 , D−1 , W D . (11.40)

This procedure can be continued and yields the operator identity


/ 0 / / 00
W D = DW + D2 D−1 , W + D3 D−1 , D−1 , W + . . . . (11.41)

The basic commutator appearing here is given by


/ −1 0
D , W = [−2, W ] = −2W − 2 (∂ μ W ) ∂μ . (11.42)

Since W contains a derivative of the field this term is of order O((∂μ φ)3 ) and
the next higher order commutator in (11.41) is of order O((∂μ φ)4 ). Equation
(11.41) thus represents an expansion in terms of gradients of the field. We also
see that it contains increasing powers of D so that we expect to encounter
only a finite number of ultraviolet, i.e. for large momenta, divergent terms.
We now use the identity (11.41) to rewrite the expansion (11.36). Up to
terms of second order in W the expansion reads
1
−Tr ln(1 − DW ) = tr(DW ) + tr(DW DW ) + O(W 3 )
2
1 1  / 0 
= tr(DW ) + tr(D2 W 2 ) + tr D3 D−1 , W W
2 2
+ O(W 3 ) . (11.43)

Note that all the terms on the rhs have the momentum dependence and the
x-dependence separated so that – according to (11.37) – they can be written
as an integral over the momentum times a spatial integral over a Lagrangian
density. Equation (11.43) is an expansion in terms of powers of W .
We now discuss the terms in the expansion (11.43) and start with the
term tr(DW ). This is explicitly given by
  
tr(DW ) = tr D m∗ 2 − m∗0 2 − ig ∂φ . (11.44)
152 11 Interacting Fields

Since
tr γμ = 0 (11.45)
and
  
m∗ 2 − m∗0 2 = g 2mF (φ − φ0 ) + g φ2 − φ20 (11.46)
we get
4 /  03
tr(DW ) = g tr D 2mF (φ − φ0 ) + g φ2 − φ20 . (11.47)

The trace operation can now be performed. This gives


 
d4k 1 /  0
tr(DW ) = 4g ∗
d4x 2mF (φ(x) − φ0 ) + g φ2 (x) − φ20 ,
(2π) k − m0
4 2 2

(11.48)

where we have used (11.37) with the trace over the Dirac indices Tr = 4.
This term is now in the desired form, i.e. it is given by a space–time integral
over fields.
However, the momentum-integral appearing here actually diverges quad-
ratically as power counting shows us. Also the next term in (11.43) diverges.
It is given by

d4k 1
tr(D2 W 2 ) = g 2 Tr   (11.49)
(2π) k 2 − m∗ 2 2
4
0
 ( )2
 
× d4x 2mF (φ(x) − φ0 ) + g φ2 (x) − φ20 − i ∂φ(x)

and diverges logarithmically.


The second term proportional to W 2 in (11.43)
 / 0  
tr D3 D−1 , W W = tr (D) (−2W − 2(∂ μ W )∂μ ) W
3
(11.50)

and all higher order terms are clearly ultraviolet convergent because they
contain higher and higher powers of the propagator D.
This means that we have only two divergent terms ((11.48) and (11.49)) in
the expansion (11.36). These can be removed by introducing proper counter
terms. We can thus define a finite, renormalized effective Lagrangian density
in the boson sector of our theory by just adding the two divergent terms
(11.48) and (11.49) to the Lagrangian
1
L̃B (x) = − φ(x)2φ(x) − V (φ(x))
2
i   
− Tr ln 1 − D(0) m∗ 2 (x) − m∗0 2 − ig ∂φ(x)
2
11.3 Bosonization of Yukawa Theory 153

d4k 1 4 / 03
− i2g ∗
2mF [φ(x) − φ0 ] + g φ2 (x) − φ20
(2π) k − m0
4 2 2

  2
g2 d4k 1
− i Tr
4 (2π)4 k 2 − m∗0 2
4 / 032
× 2mF [φ(x) − φ0 ] − i ∂φ(x) + g φ2 (x) − φ20 . (11.51)

Here it is understood that the two diverging integrals are regularized, either
by a cutoff or by dimensional regularization. The counter-terms in the 3rd
to 5th line of this equation, after taking the trace, reduce to powers of the
field φ and the d’Alembert operator and thus contribute to a change of the
potential V (φ) and a field renormalization.
Writing (11.51) in a somewhat more compact form gives

1 i
L̃B = − φ(x)2φ(x) − V (φ(x)) − Tr ln(1 − DW )
2 ( ) 2
i 1 2
− Tr DW + (DW ) (11.52)
2 2
1 1  3 / −1 0 
= − φ(x)2φ(x) − V (φ(x)) + Tr D D , W W
2 2
+ O(W 3 ) . (11.53)

where Tr denotes a trace over the Dirac indices only and where it is under-
stood that we have to use the expansion (11.36) for the Tr ln term. The two
counter terms in the second line have the structure of terms already present
in the original boson Lagrangian; they just remove the first two terms in the
Taylor expansion of the logarithmic term in the first line. The Yukawa theory
is thus indeed renormalizable, provided the potential V (φ) is ‘well-behaved’
and contains at most terms of fourth order in φ.
Part III

Gauge Field Theory


12 Path Integrals for QED

The fundamental interactions of nature like, e.g. the electroweak and the
strong interactions are described by so-called gauge field theories. In these
theories the action is invariant under certain space–time dependent unitary
“gauge” transformations. As a consequence for each field configuration there
are infinitely many others that can all be obtained by a gauge transformation
from the original one. The physics remains unchanged under such transfor-
mations.
Special problems arise when such gauge theories are quantized. The reason
is quite easy to see in the framework of path integrals. If we consider a field
Aμ , the generating functional would be naively written down as

W [0] ∼ DA eiS[Aμ ] , (12.1)

where the action S is a functional of Aμ . The integration measure stands for


a straightforward generalization of the measure used for scalar fields
n 
 4
DA ∼ dAμ (x, tj ) (12.2)
x j=1 μ=1

in the limit n → ∞ (cf. (5.11)); the measure so defined is gauge invariant as


we will see later. For a fixed field Āμ there are infinitely many other fields,
connected to Āμ by gauge transformations, that leave the action S invariant
and thus all give the same contribution to the integrand. Thus the path
integral as written down here cannot converge.
In this chapter we develop methods to deal with this difficulty in the
well-known case of Quantum-Electrodynamics. After a discussion of gauge-
invariance in this theory, we develop a method to integrate only over phys-
ically distinct electromagnetic fields. We end this chapter with a derivation
of the Feynman rules in QED.

12.1 Gauge Invariance in Abelian Free Field Theories


We start this section with a discussion of gauge invariance in a free field
configuration without currents.

U. Mosel, Path Integrals in Field Theory


© Springer-Verlag Berlin Heidelberg 2004
158 12 Path Integrals for QED

For such a field we have already discussed in Sect. 4.1.1 the Lagrangian
that gives the free Maxwell equations as equations of motion for the fields.
This Lagrangian is given by
1 1
L = − Fμν Fμν = − (∂μ Aν − ∂ν Aμ ) (∂ μ Aν − ∂ ν Aμ ) ; (12.3)
4 4
it is clearly invariant under a gauge transformation (4.23)

Aμ (x) −→ Aμ (x) = Aμ (x) + ∂μ (x) . (12.4)

Note that L has only a kinetic term. Indeed any massterm of the form m2 A2
would violate this gauge invariance. The corresponding equation of motion
for the fields Aμ is

∂ν Fνμ = ∂ν (∂ ν Aμ − ∂ μ Aν ) = (δ μ ν 2 − ∂ν ∂ μ )Aν = 0
−→ (gμν 2 − ∂μ ∂ν ) Aν = 0 , (12.5)

which is also gauge invariant; here δ μ ν is the usual Kronecker-Delta written


as a Lorentz-tensor.
We now proceed to obtain the propagator for the electromagnetic field.
As mentioned at the end of Sect. 10.2.3 the propagator of the theory could
be read off from the term in the Lagrangian that is quadratic in the field.
The Lagrangian (12.3) can indeed also be converted into a form quadratic in
the fields by writing
 
1
Ld x = −
4
d4x (∂ μ Aν − ∂ ν Aμ )(∂μ Aν − ∂ν Aμ )
4

1
= − d4x [ (∂ μ Aν )(∂μ Aν ) − (∂ μ Aν )(∂ν Aμ )
4
− (∂ ν Aμ )(∂μ Aν ) + (∂ ν Aμ )(∂ν Aμ )] . (12.6)

By partial integration we obtain


 
1
L d4x = d4x (Aμ 2Aμ − 2Aμ ∂ν ∂μ Aν + Aμ 2Aμ )
4

1
= d4x Aμ (gμν 2 − ∂μ ∂ν ) Aν , (12.7)
2
so that a physically equivalent L is given by
1 μ
L= A (gμν 2 − ∂μ ∂ν ) Aν . (12.8)
2
This Lagrangian is very similar to that of a scalar field theory (4.35). Since
it is quadratic in the fields it can be treated in analogy to the developments
in Sect. 6.1.3. According to our normal procedure (cf. (10.89)) we thus expect
to obtain the propagator of the electromagnetic field as
12.1 Gauge Invariance in Abelian Free Field Theories 159
−1
Dμν ∼ (gμν 2 − ∂μ ∂ν ) . (12.9)

The problem arises when we realize that this operator D does not exist.
We see this by applying its inverse to an arbitrary four-gradient ∂ ν G
−1 ν
Dμν ∂ G = (2∂μ − ∂μ 2) G = 0 . (12.10)

Thus, D−1 has a zero eigenvalue and, loosely speaking, its inverse D is infinite.
This infinity is linked to that in the path integral over gauge fields dis-
cussed at the start of this chapter. This can be seen by a straightforward
application of the integration formula for quadratic Lagrangians (B.18) to
the naive path integral for the gauge field Aμ
 4
W [0] = ei d x L (12.11)

with L given by (12.8). Equation (6.27) yields det(D−1 ) = 0 in the denomina-


tor1 . This infinity of the path integral is an indication that we are integrating
over too many degrees of freedom. Indeed, so far no gauge condition has en-
tered anywhere as a restriction on the fields.
Therefore, we expect that one way to overcome this difficulty is to fix a
gauge, for example by imposing the so-called covariant gauge, sometimes also
called Lorentz -gauge, condition

∂ ν Aν = 0 . (12.12)

If we then consider only potentials that fulfill this condition and integrate
only over them, we expect the path integral to be well-behaved.
In order to investigate this expectation we now construct a Lagrangian
that incorporates the gauge condition (12.12), at the price of loosing its man-
ifest gauge invariance. A hint how this Lagrangian might look like can be ob-
tained by inspecting the equation of motion (12.5) with the gauge-constraint
(12.12) taken into account. This is given by the free wave equation for Aμ

2Aμ = 0 . (12.13)

This equation of motion can directly be obtained from the Lagrangian


1 1
L̃ = − Fμν Fμν − (∂μ Aμ ) ;
2
(12.14)
4 2
neither (12.13) nor (12.14) are gauge invariant.
This result can be generalized by considering now
1 1
L̃ = − Fμν Fμν − (∂μ Aμ ) .
2
(12.15)
4 2λ
1
Remember here that the determinant of an operator is given by a product of its
eigenvalues.
160 12 Path Integrals for QED

This Lagrangian, that for λ = 1 agrees with (12.14), contains a so-called


gauge-fixing term as a quadratic constraint, coupled in by means of a La-
grange-multiplier (1/(2λ)). This additional term could be considered as a
potential that becomes the more repulsive the more the actual potential Aμ
differs from the Lorentz-gauge (12.9); the constraint thus drives the system
towards that gauge.
Performing the steps leading to (12.8) once again, we find that (12.15) is
equivalent to the Lagrangian
(   )
1 μ 1
L̃ = A gμν 2 − 1 − ∂μ ∂ν Aν (12.16)
2 λ

which leads to the same action. The equations of motion following from
(12.16) are (   )
1
gμν 2 − 1 − ∂μ ∂ν Aν = 0 . (12.17)
λ
We can thus either start from a manifestly covariant Lagrangian, derive the
equations of motion from it and then impose a suitable gauge condition or
we can, alternatively, incorporate the gauge condition into the Lagrangian,
at the price of giving up its gauge invariance, and then obtain the gauge-fixed
equations of motion directly from it. The physics must obviously be the same
in both methods.
Reading again – as at the end of Sect. 10.2.3 – the propagator off from
the Lagrangian (12.16) gives
(   )−1
1
Dμν = gμν 2 − 1 − ∂μ ∂ν (12.18)
λ

with the momentum space representation


(   )−1
1
Dμν (k) = −gμν k + 1 −
2
kμ kν . (12.19)
λ

The inverse of this operator no longer has a zero eigenvalue as we can easily
see by performing again the steps in (12.10). The propagator D thus exists.
In momentum space it is given by
 
1 kμ kν
Dμν (k) = − 2 gμν + (λ − 1) 2 . (12.20)
k k

This is the propagator for the electromagnetic field2 .


The constant λ in (12.20) is arbitrary, but note that the propagator does
not exist for λ → ∞; this is just the case of the unconstrained Lagrangian
2
From here on the denominator k2 in (12.20) should really read k2 + iε in order
to ensure the proper boundary conditions for the propagator.
12.2 Generating Functional 161

discussed earlier in this section (cf. (12.3) and (12.15)). Choosing λ = 1 gives
the so-called Feynman gauge, whereas λ = 0 describes the Landau-gauge. The
Feynman gauge is the one that we are used to from classical electrodynamics;
in it the equations of motion for the electromagnetic field (12.17) read

2Aν = 0 . (12.21)

This is the equation of motion that we obtained earlier for the Lorentz gauge.

12.2 Generating Functional


Now that we have constructed a propagator for the electromagnetic field, we
can write down the generating functional as
  4 1 μ −1 ν μ
W [J] = DA ei d x( 2 A Dμν A +Jμ A )
  4 μ
= DA ei d x (L+LGF +Jμ A ) (12.22)

4
where DA = μ=1 DAμ , L is given by (12.8) and LGF is the gauge-fixing
term
LGF = −1/(2λ)(∂μ Aμ )2 . (12.23)
The action of the gauge-fixing term can most clearly be seen by going into
the Euclidean representation (cf. Sect. 5.1.1. Then the functional becomes
  4E 1 μ   E μ 2 4 E
WE [J] = DA e− d x [ 2 A (δμν 2 +∂μ ∂ν )A +Jμ A ] e− 2λ
ν μ 1
E E E
∂μ A dx

(12.24)
The gauge-fixing term thus suppresses the contributions of fields that do not
fulfill the Lorentz gauge condition to the path integral.
The normalized generating functional for the free field theory can then
be obtained (cf. (B.18))
 μ ν
Z0 [J] = e− 2 J (x)Dμν (x − y)J (y) d xd y ,
i 4 4
(12.25)

and the two-point function is, as usual (see (10.88), (10.89)), found to be

Gμν (x, y) = iDμν (x − y) (12.26)

This two-point function obviously depends on the gauge chosen, as can be


seen by the presence of the parameter λ in (12.20); it depends in an even wider
sense on it because the covariant gauge has been chosen from the outset.
The path integral (12.22) does not suffer from the problems mentioned at
the start of this section. Now the matrix in the quadratic term of L (12.8)
can be inverted and the propagator exists. Furthermore, all fields that differ
162 12 Path Integrals for QED

from a specific field by a gauge transformation that leads out of the covariant
Lorentz gauge have a higher action and thus contribute less to the path
integral.
While the Lagrangian and the corresponding action themselves are gauge-
independent the integral over the gauge-fixing term and the source terms
do depend on the gauge. Thus, the generating functional itself is gauge de-
pendent and, consequently, also the Green’s functions obtained from it by
functional derivatives. All of the gauges corresponding to the various val-
ues of λ, however, lead to the same physics because in calculating a tran-
sition amplitude Dμν will couple to conserved currents only which fulfill
∂μ j μ = 0 → kμ j μ (k) = 0. As a consequence the λ-dependent term in (12.20)
always drops out and the physically observable transition amplitudes are all
gauge independent. This also holds for gauges other than the covariant gauge
treated here.

12.3 Gauge Invariance in QED


We now generalize all the considerations of the preceding section to a system
of coupled fermions and electromagnetic fields. The full Lagrangian then con-
sists of a part that describes the free electromagnetic field (cf. (12.3)), the free
fermions (10.88) and an interaction between both fields that is determined by
the product of the conserved fermion current and the electromagnetic field
(cf. (4.18))
1
L = − Fμν Fμν + Ψ̄ (i ∂ − m) Ψ − eΨ̄ γμ Ψ Aμ
4
1
= − Fμν Fμν + Ψ̄ [iγ μ (∂μ + ieAμ ) − m] Ψ . (12.27)
4
In the second line the interaction term has been rewritten so that the so-
called minimal substitution of the four-derivative by the so-called covariant
derivative
∂μ → Dμ = ∂μ + ieAμ (12.28)
becomes apparent (e = −|e|).
Due to the presence of the coupling term this Lagrangian is invariant
under the gauge transformation
1
Aμ (x) −→ Aμ (x) = Aμ (x) + ∂μ (x) (12.29)
e
only if also the fermion field is changed simultaneously
Ψ (x) −→ Ψ  (x) = e−i(x) Ψ (x) . (12.30)
The latter represents a x-dependent phase transformation, a so-called local
gauge transformation. The photon field Aμ is correspondingly called a gauge
field.
12.3 Gauge Invariance in QED 163

In Sect. 4.2.2.1 we have seen that invariance of the fermion Lagrangian


under a global phase transformation, i.e. one with a constant phase , leads to
current conservation. Here we see that it is just this current that couples to
the electromagnetic gauge field through the covariant derivative (12.28). We
also see that by requiring invariance of the photon field Lagrangian under the
transformation (12.29) we are led to postulate an invariance of the fermion
field under the local phase transformation (12.30).
We can now turn this argument around: postulating invariance of the
fermion field under the local phase transformation (12.30) necessitates the
introduction of a vector field Aμ (x), needed to cancel the extra terms gen-
erated by the derivative in the Dirac Lagrangian acting on the x-dependent
phase (x). The coupling of this vector field to the fermions is fixed by this
requirement.
Formally, this can be seen by requiring invariance of the free Dirac La-
grangian (4.47) under a unitary transformation of the fermion field

Ψ (x) −→ Ψ  (x) = U (x)Ψ (x) with U (x) = e−i(x) . (12.31)

If we then introduce a vector field Aμ and have


 
Dμ = U −1 (x) ∂μ + ieAμ (x) U (x) = U −1 (x)Dμ U (x) , (12.32)

then
LF = Ψ̄ (iγ μ Dμ − m) Ψ (12.33)
remains invariant under the transformation (12.31). Condition (12.32) fixes
the required transformation properties of the vector field Aμ (x)
 
i i
Aμ (x) −→ Aμ (x) = − U (x)Dμ U −1 (x) = U (x) Aμ (x) − ∂μ U −1 (x) .
e e
(12.34)
It is easy to see that with the transformation U given by (12.31) the latter
relation is identical to (12.29). Noting that the field tensor can also be written
as
i
Fμν = − [Dμ , Dν ] (12.35)
e
we get with the help of (12.32) under the gauge transformation

Fμν −→ Fμν = U (x)Fμν U −1 (x) = Fμν . (12.36)

Starting from (12.31) we have thus recovered the gauge field transforma-
tion (12.34) and the gauge invariance of the field tensor. Consequently, also
the free electromagnetic field Lagrangian is invariant under the local gauge
transformation U (x).
164 12 Path Integrals for QED

Generating functional. Due to the presence of the gauge-fixing term and the
source terms also the generating functional of the interacting theory is not
gauge invariant
  μ 2 μ
W [η, η̄, J] = DA ei (L− 2λ (∂μ A ) +Jμ A +η̄ψ+ψ̄η)d x ,
1 4
(12.37)

where L is now given by (12.27), nor is the two-point function following from
it. However, again, physical transition rates will be independent of the special
gauge if the photons couple only to conserved currents.

12.4 Feynman Rules of QED


The Feynman rules in the covariant gauge can then easily be obtained.
• The free photon propagator is – in the covariant Feynman gauge – given
by (12.20). It is represented by a wiggly line
−gμν
iDμν (k) = i =
k 2 + iε μ ν

• The free fermion propagator is given by (11.1). It is represented by a line


with an arrow on it that denotes the flow of charge (fermion number); p is
the momentum in that direction.
p + m
SF (p) = =
p2 − (m − i)2

• The fermion-photon vertex can be obtained as described in Sect. 11.1 for


the coupling of a scalar field. It is then given by

−ieγ μ = μ

• Four-momentum is conserved at every vertex. All undetermined loop mo-


menta have to be integrated out with

d4k
4 (12.38)
(2π)
• Closed fermion loops contribute a factor (−1).
The latter rules follow from our discussions in Sect. 11.1.
These rules are sufficient to determine the n-point Green’s function. In
order to get the S matrix elements we have to remove the propagators for
12.4 Feynman Rules of QED 165

the external legs and multiply with the corresponding free fields (cf. Sects.
7.2.2 and 10.2.2).

Free field operators. For easier reference we give here now these free field
operators and the additional Feynman rules for S matrix elements.

1. The free electromagnetic field can – in analogy to the corresponding ex-


pansion for a scalar field (7.16) – be written in a normal mode expansion
as

d3k   (p) 
3
(p) (p) † (p) ∗ ∗
Aμ (x) = aμ (k)μ fk (x) + aμ (k)μ fk (x) (12.39)
(2π)3 p=0

with the functions fk (x) defined in (7.17). Here the four-momentum of


(p)
the free photon is given by kμ = (ωk = k, k). The vectors μ form a four-
dimensional basis of polarization vectors. All these vectors are space–like
(p) ∗ μ
and normalized to μ (p) = −1. They can be chosen to have the form
μ
(p) = (0, ) in a given Lorentz-frame. For real incoming or outgoing
photons the polarization vectors have to be transverse to the photon
momentum.
Following the considerations at the end of Sect. 7.2.2 we thus get the
Feynman rule for an external photon with polarization μ : if the photon
is outgoing then it yields a factor ∗μ (k), if it is ingoing then it gives μ (k),
as can be seen from (12.39).
2. The free fermion field can similarly be Fourier-expanded (cf. (10.78))

√ d3k   † ∗

Ψ (x) = 2m aks u(k, s)f k (x) + b ks v(k, s)f k (x) .
(2π)3 s
(12.40)
An outgoing external fermion in a Feynman graph is then represented by
a factor ū(k), an incoming by u(k). Correspondingly, (10.78) shows that
an outgoing antiparticle is represented by v(k) and an incoming one by
v̄(k).
13 Path Integrals for Gauge Fields

In this chapter we discuss the more general case of gauge-fields also in non-
Abelian gauge field theories. We start this chapter with a short discussion of
the most important features of gauge-field theories and then develop the path-
integral formalism for them. In doing so we will recover the developments of
the last chapter for QED as a special case.

13.1 Non-Abelian Gauge Fields


In Sect. 12.3 we have discussed how the Lagrangian of QED can be ob-
tained from the requirement of so-called gauge-invariance of the fermion La-
grangian, i.e. invariance under a local phase transformation. The field theory
thus generated is called gauge field theory; it has been constructed by ‘gaug-
ing’ the symmetry of the free fermion Lagrangian (4.47) under a global phase
transformation. The latter is also called a U (1) symmetry since the phase
transformations (4.71) obviously represent a group of transformations that
depend on one parameter, the phase, and are unitary.

Fundamentals of Lie groups. This concept can be generalized to other groups


of unitary transformations. In more general gauge field theories the starting
point is again an invariance of the Lagrangian under the transformation of
the particle fields (see [4])

ψ → ψ  = U ψ(x) . (13.1)

The particle fields are described by N -component spinors ψ where the com-
ponents represent internal degrees of freedom. Here U is a unitary transfor-
mation - 2 N
k T k k
Tk
U = e−i k=1 ≡ e−i (13.2)
where the T k are N × N dimensional complex matrices and the k are N 2
continuous parameters. The matrices T k , also called generators of the trans-
formation, are hermitian so that the transformations generated by U are uni-
tary. Consequently, also all transformations U can be represented by N × N -
dimensional unitary matrices. Transformations with these properties form
the group of unitary transformations in N dimensions, called U (N ), with N 2

U. Mosel, Path Integrals in Field Theory


© Springer-Verlag Berlin Heidelberg 2004
168 13 Path Integrals for Gauge Fields

generators. From each of these transformations we can separate out a simple


phase transformation U (1). What remains is a group of special unitary trans-
formations SU (N ) (with U (N ) = SU (N )×U (1)) that have unit determinant.
The latter implies that the matrix representations of the N 2 − 1 generators
T k of SU (N ) have vanishing trace [4]. All these groups with unitary ma-
trix representations that depend on continuous parameters are examples of
so-called Lie groups.
Note here that the dimension d of the representations of the generators
T k must not necessarily be equal to N . If it is possible to change the basis on
which the T k act such that the generators become blockdiagonal, then the
matrix representation is called reducible; otherwise it is irreducible. The irre-
ducible representation of the generators of SU (N ) with dimension d = N is
called the fundamental representation. Its basis states form the fundamental
basis.
The generators T k form a Lie algebra with the defining commutation
relations
[T l , T m ] = if lmn T n (l = 1, 2, . . . N 2 − 1) ; (13.3)
the f lmn in (13.3) are the completely antisymmetric, real structure constants
of the group. The generators are normalized such that
  1
tr T l T m = δ lm (13.4)
2
in the fundamental representation. An important matrix representation of
the generators is the so-called regular representation
 j kl
T = −if jkl (13.5)

with dimension N × N .
Since the generators do not commute the group SU (N ) is a non-Abelian
one. For the case N = 1 there is only one generator and the group becomes
Abelian, called U (1).

Gauge field theories. So far, we have not made any assumptions on the contin-
uous parameters k in (13.2). We know from our considerations in Sect. 4.2.2
that invariance of the Lagrangian under such transformations with constant
parameters leads to the conservation of associated currents.
We now generalize our discussions of gauge invariance in QED in Sect.
12.3. There we had seen that the Lagrangian of QED follows from a gauge
principle in which a global U (1) phase invariance of the free particle La-
grangian is replaced by a local one. Thus, ‘gauging’ the global U (1) symmetry
yields automatically the coupling of the particles to the gauge field.
We now take this principle over to higher-dimensional, non-Abelian global
symmetries and require invariance of the particle Lagrangian under
- k k
U (x) = e−i k  (x)T (13.6)
13.1 Non-Abelian Gauge Fields 169

with x-dependent parameters k (x). Formally, many of the discussions of


Sect. 12.3 can directly be taken over here with the one esssential change that
now there is one separate gauge field Akμ for each one of the N 2 −1 generators
T k.
Invariance under the transformation (13.6) can be achieved only if the
derivative ∂μ is replaced by the so-called covariant derivative

Dμ = ∂μ + igT l Alμ (13.7)

where g is the coupling constant and Alμ is a vector (gauge) field; the su-
perscript l (= 1, . . . , N 2 − 1) indicates that now to each internal degree of
freedom there is a separate gauge field. This “minimal substitution” fixes the
interaction between particle and gauge fields, just as in QED.
The gauge transformation of these fields must then be given as in (12.34)
by
 
i i
Aμ (x) → AUμ (x) = U (x) A μ (x) − ∂μ U
−1
(x) = − U (x)dμ U −1 (x) ,
g g
(13.8)
if the Lagrangian in the particle sector is required to be invariant. Here we
have introduced the fields Aμ in (13.8) as contractions of the generators with
the fields Akμ
Aμ = T k Akμ . (13.9)
The change of A under an infinitesimal gauge transformation is given by
 jk k 1
δAjμ = −il T l Aμ + ∂μ j
g
1 j
= (Dμ ) . (13.10)
g
Equation (13.10) shows that the field Aμ transforms under a global phase
transformation as an irreducible regular representation of SU (N ).
As for Abelian gauge fields we define the field tensor by (12.35)
i
Fμν = − [Dμ , Dν ] . (13.11)
g
Because of the presence of the non-commuting fields Aμ in the covariant
derivatives D, the tensor now assumes a more complicated structure

Fμν = ∂μ Aν − ∂ν Aμ + ig [Aμ , Aν ] . (13.12)

Because of the structure of the field Aμ , which has been obtained by con-
tracting the fields with the generators, also the field tensor F is a scalar
in the intrinsic space, obtained by contracting the field tensor Flμν with the
generators (Fμν = Flμν T l ). These scalars change in the gauge-transformation
just as under any unitary transformation
170 13 Path Integrals for Gauge Fields

Fμν → FU
μν = U Fμν U
−1
, (13.13)

but notice that for a non-Abelian gauge transformation F itself is not gauge-
invariant.
However, a gauge-invariant Lagrangian can be constructed by starting
from the Lorentz scalar Fμν Fμν which – because of the matrix structure of
the generators in the Aμ – is still an N × N matrix in the internal degrees of
freedom. By taking the trace over these degrees of freedom we obtain
 μν   
tr FU FU μν = tr U Fμν F U
μν −1
= tr (Fμν Fμν ) ; (13.14)

the last step is possible because the trace is invariant under unitary transfor-
mations. The contraction of the field tensor with itself is thus gauge-invariant.
Thus, for non-Abelian gauge fields the gauge-invariant free Lagrangian is
given by
1
L = − tr (Fμν Fμν ) . (13.15)
2
We can exhibit the internal indices explicitly by multiplying F with T l and
then taking the trace, using (13.4). This gives

Flμν = ∂μ Alν − ∂ν Alμ − gf lmn Am n


μ Aν (13.16)

and
1 μν
L = − Flμν Fl . (13.17)
4
Gauge invariance again requires that the gauge fields are massless, since a
mass term of the form m2 A · A would clearly violate gauge invariance.
Equation (13.17), together with (13.16), shows that the free gauge field
is selfinteracting, with terms ∼ A3 and ∼ A4 in L
1  
L= − ∂μ Alν − ∂ν Alμ ∂ μ Alν − ∂ ν Alμ
4
1 lmn m n  μ lν 
+ gf Aμ Aν ∂ A − ∂ ν Alμ
2
1 2 lmn lop m n oμ pν
− g f f Aμ Aν A A . (13.18)
4
Here the first line on the rhs formally looks like the Lagrangian of an Abelian
gauge field (12.3) and can be replaced by the equivalent Lagrangian (12.8).
The interaction terms in the last two lines are generated by the non-Abelian
piece of Fμν and are absent for an Abelian theory (f lmn = 0). Notice also that
the coupling constant g, determining the interaction strength, is the same for
gauge field–gauge field (13.16) and gauge field–particle (13.7) interactions.
The complete ‘generic’ Lagrangian of a non-Abelian gauge field theory
with several distinct fermion fields then reads
13.2 Generating Functional 171

1 / 0
L = − Flμν Fl μν + Ψ̄f (iγ μ Dμ − mf )Ψf . (13.19)
4
f

The sum over f runs over all fermions, but there is only one set of N 2 − 1
gauge fields common to all fermions. Since there is a common gauge field for
all fermions, all these different fermions must couple with the same coupling
constant g to the gauge fields because the coupling constant appears both in
the fermion–gauge field coupling and in the gauge field–gauge field coupling.
This universality is a special property of non-Abelian gauge field theories.
An example for the relevance of a non-Abelian gauge field theory is pro-
vided by Quantum Chromodynamics (QCD), the theory of the strong inter-
actions between quarks. Here the relevant internal degree of freedom of the
quarks is color and the symmetry group is SU (3). Another example is pro-
vided by the theory of the electroweak interaction where the gauge group is
U (1) × SU (2); in this case the relevant degrees of freedom are the electrical
charge (U (1)) and the weak isospin (SU (2)). Both of them will be discussed
in some more detail in Chap. 14.

Abelian gauge field theories. For the case of QED U (x) = e−i(x) , T k = 1 and
f lmn = 0; thus U is Abelian. Indeed (13.8) then reduces to the well known
gauge transformation of electrodynamics
1
Aμ → Aμ = Aμ + ∂μ  (13.20)
g
and also the field tensor assumes its well known form. The case of QED is
thus contained as a special case in the developments in this section. Since in
this case there is no coupling constant g in the field tensor F, universality
does not hold for non-Abelian gauge field theories.

13.2 Generating Functional

We now want to develop a path integral formulation also for non-Abelian


fields and, therefore, restrict ourselves first to the gauge field sector. We start
again with the naive expression for a path integral over gauge fields

W [0] = DA eiS[A] . (13.21)

Here the action is that of the original theory without any gauge-fixing term

S = d4x L (13.22)

and the integration measure DA stands for


172 13 Path Integrals for Gauge Fields

  N−1 
2
4
DA = d Alμ (x, tj ) (13.23)
x j l=1 μ=1

(cf. (5.11)). This integral, of course, diverges as in the case of the Abelian
theory because, as discussed in the last section, the path integral runs over
all field configurations, i.e. both over essentially distinct ones and those that
differ only by gauge transformations from each other.
It is then tempting to just add a gauge-fixing term to L as we have done
in (12.22). However, while this problem of the gauge-equivalent potentials ap-
pears for both Abelian and non-Abelian gauge field theories, there is another
difficulty that is specific for non-Abelian theories. The path integral in the
form given above, in which an integration only over the fields appears, was
shown in Chap. 1 to be valid only for quadratic Hamiltonians with constant
coefficients in front of the kinetic term. This, however, is in general no longer
the case for a non-Abelian gauge theory, where the field tensor F depends
not only on the derivatives of the fields, but also on the fields themselves.
Effectively this leads to a kinetic energy term in which – in certain gauges
– the coefficient of the momentum-dependent term depends on the fields
themselves. When considering the nonrelativistic analogue of this case in
Sect. 1.3.2 we have started from the Hamiltonian formulation of the path in-
tegral. There we have found that a new factor involving the squareroot of the
coordinate dependent coefficient appears under the path integral (see (1.57)).
It is therefore natural that in the treatment of non-Abelian gauge field theo-
ries we would also have to start from the Hamiltonian representation of the
path integral (1.32) (see [2] for a detailed treatment). At the expense of some
mathematical rigor, however, the concept of a path integral over the fields
only can be maintained and this is what we are going to do in this chapter.
In the following we will now develop a method to split the path integral
up into two factors, one containing the divergent part and the other one being
convergent. The divergent part will be seen to be an infinite constant that
drops out when we work with the normalized functional. The divergence is
here – as in the Abelian case – connected with the presence of zero eigenvalues
of the matrix in the quadratic term of the gauge field action. It is just those
zero eigenvalues that we have to factor out by integrating only over those
fields that correspond to nonzero eigenvalues of the inverse propagator.
Since we have seen in the Abelian case that zero eigenvalues of the in-
verse propagator were connected to an infinite integral over the gauge degree
of freedom (cf. the discussion around (12.11)) we start by dividing up the
total configuration space of all possible Aμ (x) into equivalence classes. Each
of these equivalence classes contains all the possible fields AU μ that can be
obtained from a specific initial field Aμ (x) by gauge transformations. The in-
tegrand of (13.21) is constant along the fields AU μ , i.e. within an equivalence
class. The path integral (13.21) is then proportional to an infinite constant
which is just the “volume” of the gauge group U and the physically relevant
13.2 Generating Functional 173

part of the path integral just runs over the essentially distinct fields Aμ (x).
This restriction can be achieved by requiring that the fields that contribute
to this integration all fulfill a gauge condition which we write in the form

F l (Akμ (x)) = 0 . (13.24)

Examples of such gauge conditions are given by the

Lorentz gauge : Fk = ∂ μ Akμ = 0


Coulomb gauge : Fk = ∇ · Ak = 0
(13.25)
Axial gauge : Fk = Ak3 = 0
Temporal gauge : Fk = Ak0 = 0 .

The challenge is now to include these conditions1 in the path integrals such
that the path integrals are well defined.
With this aim in mind we now define a functional integration only over
fields that are connected by a gauge transformation, i.e. over the elements of
the symmetry group SU (N ). The latter is determined by the transformation
k
Tk
U (x) = e−i , (13.26)

and denote the integration measure by DU ; in Fig. 13.1 this integration over
the gauge group corresponds to an integration along the solid lines. This
integration measure is gauge invariant

DU  = D(U  U ) = DU (13.27)

because of the group-property of the transformations U ; this can also be seen


by noting that we integrate over all gauge transformations U . To factor out
an integral over DU is the aim in the following paragraphs.
With this integration measure we now consider the integral

U
M−1 [Alμ ] = DU δ[F l (Akμ )] (13.28)

Here F l (Akμ ) = 0 is a gauge-fixing condition (13.24). The superscripts l, k


stand for internal degrees of freedom of the symmetry group SU (N ); for
notational convenience, we will write them out only when necessary. The
δ-functional in (13.28) is really a product of Dirac δ-functions at each space–
time point (x, tj ). AU is the gauge-transformed field. While (13.28) looks at
first sight quite formidable, we can easily get a first insight into its meaning
by considering the one-dimensional Integral

−1
M [f ] = du δ(f (u)) (13.29)

1
The Lorentz gauge is often also called covariant gauge.
174 13 Path Integrals for Gauge Fields

AUμ


Aμ(x)
Fig. 13.1. Lines of constant integrands. The horizontal axis contains the physi-
cally distinct fields, the vertical one the gauge-transformation degree of freedom.
The solid lines depict the fields within a given equivalence class; the dashed line
represents the gauge condition F (AU μ ) = 0 for an Abelian theory, the dotted line
for a non-Abelian theory exhibits the appearance of Gribov copies for large fields

which is a functional of the function f (u). According to a well-known property


of the δ-function a Jacobian appears when evaluating the integral

−1 δ(u − ui ) 1
M [f ] = du 11 11 = 11 11 . (13.30)
df df
1 du 1 1 du 1
i i

M is thus just a Jacobian. Generalizing this result to the more complicated,


higher-dimensional integral (13.28) we expect that also M represents a Ja-
cobian ( )
δF
M[A] = det . (13.31)
δU
This simple argument is indeed correct as we will see when we evaluate M
explicitly in one of the following sections.
Since the δ-functional in (13.28) restricts the gauge transformations to
U ≈ 1 it suffices to consider only infinitesimal gauge transformations
U (x) ≈ 1 − il (x)T l with   1 . (13.32)
In this case the measure can be written as
 N−12

DU = dl (x, tj ) (13.33)


x,j l=1
13.2 Generating Functional 175

in a discrete space–time representation. For the measure over infinitesimal


transformations in (13.33) we have
l l
U  = U  U ≈ (1 − i T l )(1 − il T l ) ≈ 1 − i( + l )T l
l
= 1 − i T l , (13.34)

so that the Jacobian from  to  =  +  is


 l 
δ
det =1. (13.35)
δ k
Thus the measure (13.33) is indeed gauge-invariant.
The condition F (AU μ ) = 0 defines a surface that crosses that of each
equivalence class (dashed line in Fig. 13.1); we assume that for an initial field
Aμ this is the case for one group element. This is certainly the case for an
Abelian gauge field theory, like e.g. QED. There, the gauge transformation
U reads (cf. (13.20))
1
Aμ (x) −→ AU
μ (x) = Aμ (x) + ∂μ (x) (13.36)
g
and the gauge condition becomes
1
F (AU μ μ
μ ) = ∂μ A (x) + ∂μ ∂ (x) = 0 (13.37)
g
where we have chosen as an example the covariant gauge ∂μ Aμ = 0. Equation
(13.37) represents an inhomogeneous wave equation for . With the boundary
condition Aμ (x) → 0 as |x| → ∞ it specifies a unique (x). As Gribov has
pointed out this is no longer the case if we are dealing with a non-Abelian
field theory. In this case there can be several equivalent fields, all fulfilling
the same gauge condition (open points in Fig. 13.1). Thus, in this case one
also has to remove these so-called Gribov copies from the path integration.
However, since these copies appear only at large field strengths we can neglect
them in a perturbative treatment, which involves only small quantum fields,
around either a vanishing or a classical field configuration.
Not only the integration measure, but the integral (13.28) as a whole is
gauge invariant. This can be seen by writing the Jacobian as an integral and

starting from a gauge-field AU
 (  U )

U
M−1 [AUμ ] = DU δ F A μ . (13.38)

We now change the variables from U to U  = U U  and get, because the


integration measure is gauge-invariant,
    
U
M [Aμ ] = DU  δ F AU
−1
μ = M−1 [Aμ ] , (13.39)
176 13 Path Integrals for Gauge Fields

since U  is just an integration variable that could be renamed into U . M−1


is thus indeed gauge-invariant.
We can now transform the original path integral (13.21) by inserting the
1 from (13.28). This gives
  
/  0 iS[A]
W [0] = DA e iS[A]
= DA M[Aμ ] DU δ F AU μ e . (13.40)
6 78 9
=1

We now look first at the integral over A. Its value is independent of U since the
U -dependence in the δ-function can be removed by a gauge transformation
AUμ → Aμ . Since S, M and DA are invariant under this transformation the
integrand no longer depends on U and we can take the integral over U out
of the remaining integral
 
W [0] = DU DA M[Aμ ] δ [F (Aμ )] eiS . (13.41)

This functional determines our theory in the F (Aμ ) = 0 gauge. The integral

DU (13.42)

is just the infinite factor we wanted to pull out from the expression.
Since this factor is cancelled when normalizing the functional, we can
from now on work with the generating functional

/  0  4 l lμ
Z[J] = DA M[Aμ ] δ F l Akμ ei d x (L+Jμ A ) . (13.43)

This functional can be used for a perturbation theoretical treatment because


it is finite. This was not possible for the original functional (13.21) which
diverges because of the integration over physically equivalent gauge fields.
In contrast, in (13.43) the integration over A is finite since the integration
extends only over fields that – because of the presence of the δ-functional –
all fulfill the gauge condition.
We have seen at the end of Sect. 13.1 that the case of an Abelian gauge
field is contained in the more general framework of non-Abelian gauge fields.
This then immediately raises the question of how the generating functional
(12.22), which looks quite different, is related to the general case (13.43).

13.3 Gauge Fixing of L


For a simplification of the generating functional (13.43) we can actually use
the easier example of an Abelian theory, treated in Sect. 12.1, for guidance.
There we had succeeded, in (12.22), to write down a meaningful generating
13.3 Gauge Fixing of L 177

functional that did not contain such a δ-functional, but instead a gauge-fixing
term in the Lagrangian.
Taking this result as a guideline we, therefore, in the following paragraphs
reformulate Z[J] into a more tractable form. In doing so we will find also an a
posteriori justification for the Abelian results (12.37) which at that point we
had written down guided only by the observation that it leads to a reasonable
propagator.
To achieve a simplification we consider somewhat more generalized gauge
conditions
F m (Aμ (x)) − C m (x) = 0 , (13.44)
where C m (x) is an arbitrary function of space–time. Note that for given A,
C m and F m this condition determines a gauge-transformation. The definition
of M[A] changes correspondingly

/ 0
M [Aμ ] = DU δ F m (AU
−1 m
μ (x)) − C (x) . (13.45)

Suppose, we looked at a field for which the argument of the δ-functional


with given C m (x) vanishes and then looked at this argument again, but with
a different function C̃ m (x), so that the argument no longer vanishes. We
could then find a gauge transformation that makes the argument of the δ-
functional zero again. Since the rest of the integral is invariant under gauge
transformations (and we integrate anyway over all gauge transformations)
the value of the integral does not change. Thus, the FP determinant M does
not change under this replacement and is even independent of the specific
function C m (x). Then, obviously, also
 
Z[J] = DA M[Aμ ] δ [F m (Aμ ) − C m (x)] ei LJ d x
4
(13.46)

with
LJ = L + J · A (13.47)
is independent of C(x). We are thus free to change the integrand by a weight-
ing factor  m
e − i
2λ [C (x)]2 d4x (13.48)
and integrate functionally over the function C; this will only change the
normalization of the integral. We then get
    4
Z[J] = DC DA M[Aμ ] δ [F m (Aμ ) − C m (x)] ei LJ − 2λ (C(x)) d x .
1 2

(13.49)
Now the functional integration over C can be performed since C appears
in the δ-functional. This gives
   4
m
Z[J] = DA M[Aμ ] e i L J − 1
2λ [F (Aμ )]2
dx . (13.50)
178 13 Path Integrals for Gauge Fields

In this form L has again picked up a ‘penalty potential’ (see discussion after
(12.15)) and the δ-functional has disappeared. This generating functional is
finite due to the presence of the gauge fixing term and can, therefore, be used
for a derivation of the Feynman rules for perturbation theory. As we have
seen in the last section for the case of QED the divergence of the original
path integral was linked to the zero eigenvalues of the inverse propagator.
The vanishing eigenvalues were connected to the infinite integration over
the gauge transformation degree of freedom. Here now this infinity has been
removed by fixing the gauge and the propagators are finite.
Notice, however, that the appearance of the gauge-fixing term in the ac-
tion is not the only change. An additional factor, the functional M[Aμ ],
appears in the integrand of the path integral. In QED such a factor was ab-
sent (cf. (12.37)). In the next Section we will show that indeed in Abelian
theories this functional does not depend on A; it can, therefore, be pulled
out of the generating functional. In non-Abelian theories, however, M[Aμ ] in
general does depend on the fields and thus cannot be pulled out of the path
integral. The deeper reason for this difference of Abelian and non-Abelian
theories lies in the different structure of the field tensors in both theories.
Due to the presence of the self-interaction term in the non-Abelian field ten-
sor (13.16) for general gauges the kinetic part of the field energy becomes
field-dependent and fields and conjugate momenta no longer decouple in the
Hamiltonian. This is then exactly the situation that we have encountered
already in Sect. 1.3.2 in the framework of nonrelativistic quantum mechanics
where a coordinate dependent term also led to an additional factor in the
path integral (cf. (1.57)).
While everything else in the integrand of the path integral (13.50) is gauge
invariant, the source term and the gauge-fixing terms break this invariance.
Thus, the specific rules of the perturbation theory, e.g. the Feynman rules,
will depend on the gauge. In particular also the Green’s functions, obtained
as functional derivatives of Z[J], are not gauge invariant. This, however, is
no problem since these Green’s functions only appear at intermediate stages
of the calculation. The S matrix elements, which are physical quantities, are
then gauge invariant again. We have seen an example for that in the last
section where we discussed the QED propagators; in calculating the S ma-
trix the gauge-dependent propagators are always contracted with conserved
currents and the gauge-dependent terms then drop out.

13.4 Faddeev–Popov Determinant


It now remains to evaluate M[Aμ ]. Since M[Aμ ] is gauge invariant we can
always choose a field Aμ that fulfills the condition F (Aμ ) = 0. This allows
us to replace the condition F = 0 for the derivative in (13.55) by the simpler
condition U = 1. We thus have to consider only infinitesimal gauge transfor-
mations
13.4 Faddeev–Popov Determinant 179

U ≈ 1 − il T l (13.51)
for the evaluation of M. For these integrations close to U = 1 we can use
the integration measure (13.33). In an abbreviated, but obvious notation,
the path integral over the gauge transformations then reads for n discretized
space coordinates2

M−1n = d1 d2 . . . dn δ n [F ()] (13.52)

with n = (xn ). In order to evaluate this integral we now change the inte-
gration variables from dn to dF . This gives

d1 d2 . . . dn det M = dF (x1 ) dF (x2 ) . . . dF (xn ) , (13.53)

where det M is the Jacobian for this transformation of the integral measure.
We thus get

−1
M−1n = dF (x1 ) dF (x2 ) . . . dF (xn ) δ n [F ] (det M ) . (13.54)
6 78 9
=1

The integral is identically equal to 1 so that Mn is just the Jacobian


 
δF (xi )
Mn = det Mij = det . (13.55)
δ(xj ) F =0

This is just the result we had obtained in a first rough argument in (13.31).
Rewriting this result back into a continuum representation and adding all
the relevant group indices gives
 m 
δF (Aλ (x))
M[Aμ ] = det M ml (x, y) = det . (13.56)
δl (y) =0

The determinant here, the so-called Faddeev–Popov (FP) determinant, has


to be calculated with respect to both the space–time indices (x, y) and the
N 2 − 1 SU (N ) group indices (m, l). Note that the FP determinant depends
in general on the gauge fields Akμ .
We can now replace M[Aμ ] in (13.50) by the Faddeev–Popov determinant
(13.56) and obtain as the generating functional for the gauge field sector

   m
 4
Z[J] = DA det M ml (x, y) ei LJ − 2λ [F (Aμ )] d x .
1 2
(13.57)

This generating functional is the central result of this chapter.


2
To facilitate the notation we drop the group indices here.
180 13 Path Integrals for Gauge Fields

13.4.1 Explicit Forms of the FP Determinant

The question still remains why no FP determinant is present in the generating


functional of QED (12.37). In this section we will, therefore, show that indeed
the presence of the FP-determinant in (13.57) is a typical feature of non-
Abelian gauge field theories and that it can be removed in Abelian theories.
In order to obtain an explicit expression for the FP determinant, i.e. the
functional derivative of the gauge condition F with respect to the gauge
transformation parameters  (13.56), we expand the gauge condition F (AU μ)
around a field that fulfills the gauge condition, i.e. F m (Aμ ) = 0 and U ≈ 1.
Under Aμ → AU μ = Aμ + δAμ the condition F changes into
 m
 l 
m U 4 δF Aλ (x)
F (Aμ ) = 0 + d z δAkμ (z)
δAkμ (z)
 m
 l 
1 4 δF Aλ (x) k
= dz (Dμ (z)) . (13.58)
g δAkμ (z)

Here we have used (13.10) to replace δA with the covariant derivative


1 k 1 / 0 k
δAkμ = (Dμ ) = ∂μ + igT n Anμ  . (13.59)
g g
Taking now the derivative according to (13.56) we get
1
ml δF m (Aλ (x)) 11
M (x, y) = 1
δl (y)
 m
=0 
1 4 δF Alλ (x) δ k
= dz k l
(Dμ (z))
g δAμ (z) δ (y)
  l  1
m 1
1 4 δF Aλ (x) kl 1
= dz D μ (z)δ 4
(z − y) 1 . (13.60)
g δAkμ (z) 1
F =0

This is a a general form for the FP-matrix. Its actual value depends through
the functional derivative on the gauge used.

13.4.1.1 Abelian Fields We have seen earlier that the non-Abelian gauge
field theory contains the Abelian U (1) symmetry group as a special case.
We therefore start by analyzing M for an Abelian gauge field theory such as
QED. We obtain directly from the definition (13.56)
  11  1
δF AU μ (x) 1 δF Aμ (x) + 1e ∂μ (x) 11
M (x, y) = 1 = 1 . (13.61)
δ(y) 1 δ(y) 1
F =0 =0
μ
Choosing, for example, the covariant gauge F = ∂μ A = 0 gives
1 2δ(x) 1
M (x, y) = = 2δ 4 (x − y) . (13.62)
e δ(y) e
13.4 Faddeev–Popov Determinant 181

Thus, M is independent of Aμ , and therefore its determinant can be pulled


out of the path integral in (13.57), changing only the normalization. The
generating functional can thus be taken as
   4
μ 2
Z[J] = DA e i L J − 1
2λ [∂μ A ] dx . (13.63)

This is indeed just the functional (12.22) for QED.


13.4.1.2 Non-Abelian Fields For a non-Abelian field the first term in the
gauge transformation (13.8) has a more complicated form so that it will in
general contribute an Aμ -dependent factor.
• Covariant gauge. The covariant (Lorentz) gauge is given by
F n (Aμ ) = ∂ μ Anμ = 0 . (13.64)
The FP-matrix is then obtained from (13.60) as
  λ m  1
1
ml 1 4 δ ∂ Aλ (x) kl 1
M (x, y) = dz D μ (z)δ 4
(z − y) 1 (13.65)
g δAkμ (z) 1
F =0
 1
1  μ 4  kl 1
= 4 mk
d zδ ∂x δ (x − z) Dμ (z)δ (z − y)11
4
.
g F =0

By using ∂xμ δ 4 (x − z) = −∂zμ δ 4 (x − z) and partial integration we obtain


 1
1  kl 1
ml
M (x, y) = d z δ δ (x − z)∂z Dμ (z)δ (z − y) 11
4 mk 4 μ 4
g F =0
1
1 μ  ml  1
= ∂x Dμ (x)δ 4 (x − y) 11 . (13.66)
g F =0
For the case of QED this expression reduces to the form already found in
(13.62). This is so because in this case δAμ = ∂μ /e in (13.58) and thus in
this case instead of Dμ simply ∂μ appears in (13.66).
• Axial gauge. It is interesting to note that the FP-determinant is inde-
pendent of the fields Aμ not only for Abelian gauge theories, but also in a
non-Abelian theory, if the so-called axial gauge is chosen. This axial gauge
is defined by the constraint
n·A=0 , (13.67)
where n is a fixed four-vector; this form comprises the last two cases in
(13.25). In this case we obtain from (13.60)
  λ m  1
1
ml 1 4 δ n Aλ (x) kl 1
M (x, y) = dz Dμ (z)δ 4
(z − y) 1
g δAkμ (z) 1
F =0
   1
1 mk μ   kl 1
= δ n kl z
d z δ (x − z) δ ∂μ + ig T
4 4 j
Aμ (z) δ (z − y)11
j 4
g F =0
1 ml μ x 4
= δ n ∂μ δ (x − y) , (13.68)
g
182 13 Path Integrals for Gauge Fields

because F = nμ Anμ = 0. Thus in this case the FP-determinant is again


independent of Aμ so that it can, as in the Abelian case, be pulled out of
the path integral and put into the normalization.

13.4.2 Ghost Fields

While we now know how to evaluate the FP determinant for various gauge
conditions we are still faced with the problem: to derive the Feynman rules
for a non-Abelian gauge field theory. The presence of the FP determinant in
Z[J] complicates the by now standard procedure to determine the Green’s
functions, i.e. the vacuum expectation values of time-ordered products of field
operators, by means of a functional derivative of Z.
The trick to bypass this problem is now to use (10.56) to express the
Faddeev–Popov determinant by another path integral over (hypothetical)
Grassmann fields g(x). We have
  m ml l
det(iM ) = Dḡ Dg e−i ḡ (x)M (x,y)g (y) d xd y .
4 4
(13.69)

Here the Grassmann fields g and ḡ carry the indices of the internal symmetry
group SU (N ). They have to be Grassmann fields because for commuting
fields the determinant would be inversely proportional to the integral on the
rhs. The normalized generating functional of the theory then becomes3 (cf.
(13.50))
  m
Z[J] = DA det(M ) ei (LJ − 2λ [F (Aμ )] )d x
1 2 4

   
i LJ − 2λ
1
[F m (Aμ )]2 − ḡM g d4y d4x
= DA Dḡ Dg e . (13.70)

The step performed here is the inverse to the bosonization of a Yukawa theory
in Sect. 11.3. There we had integrated out the fermionic degrees of freedom
and picked up a fermion determinant as a remainder. Here we have started
with the determinant and have converted it into an additional path integral.
If we remember that the FP determinant owes its existence to the inte-
gration over the gauge degree of freedom, then we see that the introduction
of the fields g(x) amounts to a dynamical treatment of the gauge invariance.
We therefore expect that this treatment generates also new Feynman graphs
involving the fields g(x). The so-called ghost field g is clearly unphysical. It
carries no Dirac indices and thus is a Lorentz-scalar. On the other hand, it
is an anticommuting field and thus has the ”wrong” spin-statistics relation.
Ghosts can therefore appear only on internal lines in a Feynman diagram.
This is in line with the fact that the ghosts can be integrated out thus gen-
erating the FP determinant.
3
A factor det(i · 1) has been absorbed into the normalization.
13.4 Faddeev–Popov Determinant 183

In order to be able to calculate vertex functions involving internal ghost


lines we now supplement this generating functional with sources also for the
ghost field g and obtain for the generating functional of the full theory in the
gauge-ghost sector
  eff 4
Z[J, γ, γ̄] = DA Dḡ Dg ei LJ d x (13.71)

with
1 m
J (x) = L(x) −
Leff [F (Aμ )]2
 2λ
− ḡ(x)M (x, y)g(y) d4y + Jμn (x)Anμ (x) + γ̄ n (x)gn (x) + ḡn (x)γ n (x) .

(13.72)

The second term here is the gauge-fixing term. As stressed earlier the gauge-
fixing term breaks the gauge-invariance and causes the path integral to be
well defined. As we have also seen this gauge fixing term is not enough to
specify the generating functional for non-Abelian theories. In these the FP
determinant appears in addition and manifests itself in the presence of the
fictitious “ghost field”, introduced in the third term in (13.72). Since the
ghosts carry the indices of the underlying symmetry group SU (N ) there are
as many ghost fields as there are gauge fields. Because in general M = M (A)
they couple to the gauge fields, and only to them. The last three terms are the
source terms for the gauge fields (J)and the ghost fields (γ, γ̄), respectively.
In the covariant gauge (cf. (13.66)) the ghost-part of the Lagrangian reads

Lghost = −ḡM g → ∂ μ ḡm (x)Dμml gl (x)


 
ml
= ∂ μ ḡm (x) δ ml ∂μ + ig (T n ) Anμ (x) gl (x)
 
ml
→ −ḡm (x) δ ml 2 + ig (T n ) Anμ (x)∂ μ gl (x) . (13.73)

after rescaling the ghost field (g → gg) in the first line; in the last line a
partial integration in the first part and the covariant gauge condition have
been used. Note that indeed the ghost field couples only to the gauge field
and not to any of the other fields, like e.g. fermions, in the theory.
As we have pointed out above, there is no need to introduce a ghost
field when we are dealing with an Abelian gauge theory. In this case M is
independent of the gauge field Aμ and the ghost-field g is decoupled from Aμ ;
it can, therefore, be left out from the beginning in this case. It is worthwhile
to remember that the same situation is encountered in a non-Abelian theory
if we work in the axial gauge. In this gauge, however, the propagator is rather
complicated because it involves explicitly the (arbitrary) vector nμ .
184 13 Path Integrals for Gauge Fields

13.5 Feynman Rules


Equations (13.71) and (13.72) give the final result for the generating func-
tional of a non-Abelian gauge-field theory. The generating functional now has
the same form as the one obtained earlier for QED (cf. (12.37)), except for
the presence of the ghosts which are a typical feature of non-Abelian gauge
field theories. If the system under study contains also physical fermions, then
the Lagrangian has to be supplemented by a fermion term along the lines
developed in Chap. 10 together with the proper source terms. The fermion-
gauge field coupling is mandated by local gauge invariance (cf. (13.7)) and
appears through the covariant derivative.
Since the generating functional (13.71) is finite and free of the overcount-
ing of gauge fields discussed at the start of this chapter it can be taken as
a starting point to derive the Feynman rules for non-Abelian gauge field
theories.

Gauge field couplings. For the gauge field and its couplings we obtain the
Feynman-rules by isolating the self-interaction terms in L. We get
The first term alone just looks like the Lagrangian of an Abelian field.
It is the only term quadratic in the fields, so that the gauge boson propa-
gator is just that of the photon found in (12.20), except for the additional
SU (N ) labels. We thus have for the gauge boson two-point function (cf.
(12.26),(12.20))
1)

μ  
k ν lm −iδ lm kμ kν
= iDμν = 2 gμν + (λ − 1) 2 .
l m k + i k
(13.74)
Note that D has to be diagonal in the internal SU (N ) quantum numbers
l and m because L involves a contraction over the group indices. In the
graph μ and ν are Dirac indices.
The second line on the rhs of (13.18) gives the cubic self-coupling terms
of the gauge field
1 lmn  μ lν   μ lν  m n
L3 = gf ∂ A − ∂ ν Alμ Am n
μ Aν = gf
lmn
∂ A Aμ Aν . (13.75)
2
It can be represented by the graph in Fig. 13.2. Its Feynman rules can be
obtained by considering the generating functional of lowest order in the cou-
pling constant g. The only term of Z that contributes to the three-point
Green’s function is that of order O(J 3 ); it is obtained by just replacing the
fields by those generated by the external source. This gives (cf. (8.45))
  
1 ∂ 1 ∂ 1 ∂
Z[J] = −i d4x (−g)f lmn ∂ μ eiS0 [J]
i ∂Jνl (x) i ∂J mμ (x) i ∂J nν (x)
13.5 Feynman Rules 185
   
= igf lmn d4x ∂ μ Dlk νλ (x − y  )Jλk (y  )d4 y 
 
mk
× Dμλ (x − y  )J kλ (y  )d4 y  nk
Dνλ (x − y  )J kλ (y  )d4 y 

+ terms of lower order in J
 jk
× e− 2
i
J jκ (x )Dκλ (x − y  )J kλ (y  )d4x d4y  . (13.76)

Here the Latin superscripts refer to the intrinsic SU (N ) degrees of freedom,


the Greek subscripts and superscripts are those of the Lorentz group. The

σj

πh
p q

ρi

Fig. 13.2. Cubic gauge coupling term. The first symbols at each gluon line give
the Lorentz indices, the second denote the SU (N ) labels

three-point function can now be obtained by differentiating Z[J]


 3 1
δ3 Z 1
hij 1 1
Gπρσ (x1 , x2 , x3 ) = 1 . (13.77)
i δJ hπ (x1 )δJ iρ (x2 )δJ jσ (x3 ) 1
J=0

This gives

Ghij
πρσ (x1 , x2 , x3 )
 *
lmn
  mj
= −igf d x ∂ μ iDπlh ν (x − x1 ) iDμσ
4 ni
(x − x3 ) iDνρ (x − x2 )
+
+ permutations of external legs . (13.78)

Equation (13.78) becomes in momentum space (with G = iD)



Ghij
πρσ (p, q, k) = −igf
lmn
(−ipμ )Glh ν mj ni
π (p)Gμσ (k)Gνρ (q)

+ permutations of external legs (13.79)
186 13 Path Integrals for Gauge Fields

= −gf hji pμ Gjj hh ν
μσ (k) Gπ (p)Gii
νρ (q)

+ permutations of external legs .

In the Feynman gauge we obtain the gauge field three-point function (after
commuting 2 indices in the totally antisymmetric f )

hij −i −i −i
Ghij
πρσ (p, q, k) = gf (13.80)
k 2 p2 q 2
 
× (k − q)π gρσ + (q − p)σ gρπ + (p − k)ρ gπσ .

From this expression the vertex factor can be read off. We then have the rule
2) The vertex factor for the three-point vertex in Fig. 13.2 is
 
gf hij (k − q)π gρσ + (q − p)σ gρπ + (p − k)ρ gπσ . (13.81)

τk σj

πh ρi

Fig. 13.3. Quartic gauge coupling graph

Similarly, the four-point function can be obtained.


3) The four-point vertex in Fig. 13.3 gives the factor

−ig 2 [ f his f jks (gπσ gρτ − gπτ gρσ ) (13.82)


+ f hjs f kis (gπτ gρσ − gπρ gστ )
+ f hks f ijs (gπρ gστ − gπσ gρτ ) ] .

Fermion-gauge coupling. If there are also fermions present, then the gauge
bosons couple to them through the minimal substitution (13.7)

Dλ = ∂λ + igT l Alλ , (13.83)

so that the coupling Lagrangian is given by


13.5 Feynman Rules 187

Lf A = −g Ψ̄ γλ Ψ Aλ . (13.84)

The Feynman rules for these couplings can directly be obtained from our con-
siderations in Chap. 11. In particular, any fundamental fermion field couples
to the gauge fields according to the rule
4) The fermion-gauge vertex (Fig. 13.4) for a gauge boson with Dirac index
λ and group index l is given by:
 
−ig (γλ )βα T l f i (13.85)

Fig. 13.4. Fermion-gauge boson vertex. The fermion is represented by a solid line,
the gauge-boson by the curly line. The indices λ and l give the Dirac index and the
internal SU (N ) index of the gauge boson, resp.

Ghost couplings. We now discuss the rules for handling those terms in the
Lagrangian that involve the ghost field. The Feynman rules for the ghost
fields can be read off from (13.73).
6) The ghost propagator is given by

k i
= δ lm . (13.86)
l m k 2 + i

Note that the ghost line carries an arrow; this reflects the presence of ghosts
and antighosts, because the FP matrix is in general not hermitian. The nota-
tion is then such that particles always run with the time axis and antiparticles
run against it. More loosely speaking, the ghost line runs from ḡ to g.
Finally, (13.73) gives the coupling of the ghost to the gauge field
ml
LgA = −igḡm (T n ) ∂ μ gl Anμ = −gf lmn ḡm ∂ μ gl Anμ , (13.87)

as depicted in Fig. 13.5.


188 13 Path Integrals for Gauge Fields

σj
k

Fig. 13.5. Ghost-gauge coupling. The wavy line denotes a gauge boson whereas the
dashed line denotes the ghost. The first symbol on the gauge field line (σ) denotes
the Lorentz index, the second (j) gives the SU (N ) label

The vertex factor is obtained by considering the ghost–ghost–gauge boson


three-point function. As for the gauge three-point function it can be obtained
by considering the lowest order (in the gauge–ghost coupling) expression

 1 δ 1 δ 1 δ
Z[γ, γ̄, J] = −i d4x −gf lmn ∂μ eiS0 [γ,γ̄,J]
i δJ nμ (x) i δγ m (x) i δγ̄ l (x)
 
= − igf lmn d4x nk
Dμν (x − y  )J kν (y  ) d4 y 
 
× γ̄ k Dkm (x − y  ) d4 y  ∂ μ Dlk (x − y  )γ k (y  ) d4 y 

+ terms of lower order in J eiS0 [γ,γ̄,J] (13.88)

This generating functional determines the following rule for the gauge boson–
ghost coupling vertex:
7) The gauge boson–ghost vertex (Fig. 13.5) has to be associated with a
factor
gf hij qσ (13.89)
where q is the outgoing momentum of the ghost. The four-momentum
conservation (p + k = q) is understood here; it is not written explicitly.
In addition, we have, as for real, physical fermions, the rule
8) All ghost loops carry a sign (−1).
We also have to remember that ghosts cannot appear on external lines.
14 Examples for Gauge Field Theories

In the preceding two chapters we have developed the the Feynman rules
explicitly only for QED. For non-Abelian gauge field theories they are of
a ‘generic’ nature: they apply to all such theories. The particular physics
content of a gauge field theory is determined by specifying the symmetry
group SU (N ) and the internal degrees of freedom on which it acts.
In the preceding chapter we have sometimes – besides Quantum Electro-
dynamics (QED) – , referred to Quantum Chromodynamics (QCD) and the
theory of electroweak Interactions. In this chapter we, very briefly, summarize
the physical contents and the Lagrangian of these latter two theories.

14.1 Quantum Chromodynamics


Quantum Chromodynamics is the theory of the strong interactions. Its foun-
dations, first discussed in the late 60’s, lie in the observation that the observed
hadron spectrum can be understood in terms of quarks, mesons being com-
posed of quark-antiquark pairs and baryons of three quarks; the quarks carry
spin 1/2 and a so-called flavor quantum number. For certain baryons, such
as the Δ++ , one obtains in such a scheme (flavor SU(3)) a completely sym-
metric wavefunction, i.e. a product of a symmetric space-, symmetric spin-
and symmetric flavor-part, for the three quarks, in glaring contradiction to
the Pauli principle. This problem was solved by the introduction of a new,
additional intrinsic degree of freedom for the quarks, called color, in which the
quarks can differ (for a detailed discussion see Ref. [4]). The Pauli principle
then requires a completely antisymmetric wavefunction in this new intrin-
sic space. In experiments that produce mesons, i.e. quark–antiquark pairs,
through lepton-annihilation the number of colors has been found to be three.
The completely antisymmetric color state has therefore been identified with
the completely antisymmetric singlet state of a color SU(3) group. Gauging
this group has led to the non-Abelian theory of Quantum Chromodynamics
(QCD).
The Lagrangian of QCD is then very simply given by
1 
LQCD = − Fcμν Fc μν + [q̄f (iγ μ Dμ − mf )qf ] . (14.1)
4
f

U. Mosel, Path Integrals in Field Theory


© Springer-Verlag Berlin Heidelberg 2004
190 14 Examples for Gauge Field Theories

Here the fundamental triplets of SU (3)C are given by the quark spinors
⎛ f⎞
qr
qf = ⎝ qgf ⎠ , (14.2)
qbf

which transform according to the fundamental (= lowest-dimensional) rep-


resentation of SU (3)C . The indices r, g and b stand for the three colors (red,
green and blue), the index f for the flavor. The covariant derivative in (14.1)
is given by (cf. (13.7))
λc
Dμ = ∂μ + ig Gcμ . (14.3)
2
Here the eight matrices λc are the so-called Gell-Mann matrices, the gener-
ators of SU (3), and the fields Gc represent the eight gauge-fields, here called
the gluon fields, which form an SU (3) octet.
The field tensors Fcμν have the general form of a non-Abelian theory
(13.16). The gluons thus also interact among themselves. One of the pre-
dictions of this theory is, therefore, the existence of so-called glueballs, that
consist only of gauge-field quanta.
The theory, furthermore, has the property of asymptotic freedom, i.e. its
running coupling constant (cf. the discussion at the end of Sect. 9.4.2.2)
becomes larger with decreasing gluon momentum. This leads to a quark–
quark potential that is Coulomb-like at small distances, but at large distances
increases linearly, and thus provides an explanation for the fact that free
quarks and gluons do not seem to exist, i.e. that they are confined inside the
hadrons.
Since the QCD Lagrangian is just that of a generic non-Abelian gauge
field theory its Feynman rules can directly be taken over from Chap. 13.

14.2 Electroweak Interactions


While the Lagrangian of QCD has the rather simple, generic form (14.1), that
of the electroweak interactions is considerably more complicated. Its founda-
tions go back to the early seventies when a gauge-field theory was found that
comprised both ‘classical’ Quantum Electrodynamics and the phenomena of
weak interactions, responsible, for example, for the nuclear β-decay.
This theory is based on a product of two gauge-groups, one an Abelian
U (1) and the other one a non-Abelian SU (2). For the latter the left-handed
parts of the observed fermions, quarks and leptons,
1
ψL = (1 − γ5 ) ψ (14.4)
2
are assumed to form doublets
14.2 Electroweak Interactions 191
   
νkL ukL
Ψk = or (14.5)
ekL dkL

for the left-handed fields of the kth family of leptons and quarks. In this
version of the Standard Model of electroweak interactions the neutrino is
assumed to be massless and to occur only in a left-handed state. In (14.5) we
have used generic abbreviations
       
νk νe νμ ντ
for , or (14.6)
ek e− μ− τ−

and        
uk u c t
for , or . (14.7)
dk d s b
The right-handed parts
Ψk = ekR or qkR (14.8)
are assumed to form SU (2) singlets. If the neutrino has indeed mass, as recent
experiments on solar and atmospheric neutrinos show, then a right-handed
neutrino would also exist and could be easily implemented by postulating
also a right-handed neutrino singlet.
The U (1) gauge group is very similar to that of QED, but it is not con-
nected with the electric charge, but instead a so-called weak hypercharge
y = 2q − 2t3 where q is the electrical charge of the fermion and t3 its weak
isospin (± 12 , depending on the position of the particle in the two-component
spinors (14.6), (14.7)), just as in the usual Pauli-spinors.
The complete Lagrangian of this theory is then given by
1 1
L = − Gμν · Gμν − Fμν Fμν
4 4
 ϕ 2 1 2  ϕ 2
2
+ MW Wμ† W μ 1 + + MZ Zμ Z μ 1 +
v 2 v

μ
+ Ψ̄k iγ Dμ Ψk + LY (14.9)
k
 
1 μ 1 ϕ 1  ϕ 2
+ (∂ ϕ) (∂μ ϕ) − m2H ϕ2 1 + + .
2 2 v 4 v

The boldfaced field tensors are vectors in the internal space and their product
contains a summation over the group indices.
Since there are now two gauge groups present, we also have two gauge
field tensors appearing in the Lagrangian: Gμν for the non-Abelian SU (2)L
and Fμν for the Abelian U (1)Y . Thus, also the covariant derivative contains
both of these fields with two different coupling constants
 
1
Dμ Ψk = ∂μ + igtl Wμl + ig  yk Bμ Ψk , (14.10)
2
192 14 Examples for Gauge Field Theories

where yk is the hypercharge quantum number of Ψk and the tl are the gener-
ators of SU (2). The three fields W l and the single field B are the gauge fields
for SU (2)L and U (1)Y , resp. By linearly combining them one obtains the 2
physical charged fields Wμ and Wμ† , the neutral field Zμ as well as the photon
field Aμ . With the covariant derivative (14.10) the first term in the third line
of (14.9) describes massless fermions interacting with the gauge fields in the
standard way.
The gauge fields W and B normally have to be massless, since an explicit
mass term breaks gauge invariance. Since these fields transmit the weak in-
teraction the latter would then be long ranged. This, however, presents an
immediate problem because the weak interaction, since Fermi’s theory, is
known to be very short-ranged. A way out of this difficulty is provided by
the so-called Higgs mechanism. The main ingredient of this mass generation
mechanism is the existence of a scalar Higgs field ϕ, yet to be discovered,
with non-linear self-couplings (last line in (14.9); these self-couplings lead to
a non-vanishing vacuum field. The second ingredient is a Yukawa coupling of
the gauge-fields to this Higgs field that generates their masses MW and MZ
in the same way as described in Sect. 11.3. Similarly, the term LY in (14.9),
generates masses for all the fermions through a Yukawa coupling to the Higgs
field. For the fermions it has the simple structure
v  ϕ
LY = −gf √ Ψ̄ Ψ 1 + , (14.11)
2 v

from which the fermion mass can be read off as


v
mf = gf √ . (14.12)
2
Notice that (14.9) also contains couplings of the Higgs field to the observed
W and Z bosons.
The Feynman rules for this theory can all be obtained from our derivations
in the former chapters, but the multitude of couplings leads to many different
rules. They are given explicitly in [14].
A Units and Metric

A.1 Units
While we work in this book in the first 3 chapters, the ’classical’ quantum
mechanics part, with the usual system of units, it is customary in elementary
particle physics and field theories to choose the system of units such that
the resulting expressions assume a simpler form. In particular, instead of
working with the usual three mechanical units for length, mass and time one
introduces instead three basic units for velocity, action and length.
The choice of the velocity of light, c, as the unit for the velocity implies
that
c=1. (A.1)
This in turn means that in such a system length and time have the same
dimensions and are equivalent units. With c = 1 the relativistic energy–
momentum relation assumes the simple form

E 2 = p2 + m2 . (A.2)

Choosing next the unit of action, h̄, such that

h̄ = 1 (A.3)

connects the dimensions of mass, time and length.


Since time and length are equivalent, mass assumes the dimension of an
inverse length; the same holds consequently for the momentum and the en-
ergy (A.2). The choice of 1 fm = 10−13 cm as a length unit, for example,
leads to masses, momenta and energies all given in fm−1 . Another often-used
unit for energy is that of 1 MeV, i.e. the energy that an electron acquires
when it is accelerated by the voltage of 1 MV. The transformation of the ex-
pressions back to the standard MKSA system can be achieved by multiplying
all quantities with the appropriate combinations of h̄ and c. A particularly
useful relation for this conversion is

h̄c ≈ 197.3 MeV fm . (A.4)


194 A Units and Metric

A.2 Metric and Notation


All vectors, both ordinary three-component vectors and those in some internal
space, are distinguished by boldface italic print in this book.
In relativistic expressions four-vectors, i.e. quadruples of numbers that
transform as space–time points under a Lorentz-transformation, are always
written as    
Aμ = A0 , A = A0 , A1 , A2 , A3 , (A.5)
where A denotes a usual three-dimensional vector

A = (Ax , Ay , Az ) = (A1 , A2 , A3 ) . (A.6)

Four-vectors with a superscript are called contravariant vectors. Covariant


vectors, denoted by subscripts, are then defined by

Aμ = gμν Aν (A.7)

with the metric tensor gμν that reads in Cartesian coordinates in Minkowski
space ⎛ ⎞
1 0 0 0
⎜ 0 −1 0 0 ⎟
gμν = ⎜⎝ 0 0 −1 0 ⎠ .
⎟ (A.8)
0 0 0 −1
The metric tensor with superscripts is used to raise the indices of four-vectors;
it is thus the inverse of g and is given by

g μν = gμν
−1
= gμν . (A.9)

The product
g μν gνλ = g μ λ (A.10)
yields a unit matrix. It is often also abbreviated by the Kronecker-delta sym-
bol
δμ λ = gμ λ . (A.11)
A scalar product of 2 four-vectors A and B is then defined by

A · B = Aμ B μ = Aμ Bμ = A0 B 0 − A · B (A.12)

if Aμ and B μ are defined by

Aμ = (A0 , A) , B μ = (B 0 , B) . (A.13)

The invariant square of the four-vector is thus given by

A2 = A · A = Aμ Aμ = A20 − A2 . (A.14)
A.2 Metric and Notation 195

Here, as always in the text, the Einstein convention of summing over double
indices, one lower and one upper, is used. The components of four-vectors are
generally labeled by Greek indices, whereas roman indices are used to refer
specifically to the last three (vector) components.
The space–time four-vector is

xμ = (x0 , x) = (t, x) , (A.15)

and the four-momentum is given by

pμ = (E, p) . (A.16)

Finally, the components of the contravariant four-gradient are abbreviated


by  
∂ ∂
∂μ ≡ = , −∇ , (A.17)
∂xμ ∂t
so that the four-momentum operator is given by

p̂μ = i∂ μ , (A.18)

and the Lorentz-invariant d’Alembert operator by

∂2
∂ μ ∂μ = − ∇2 = 2 . (A.19)
∂t2
Other important four-vectors are the current

j μ = (ρ, j) (A.20)

and the electromagnetic potential

Aμ = (φ, A) . (A.21)
B Functionals

B.1 Definition
Very generally speaking a functional is a mapping from a space of functions
into the real or complex numbers. For example, the integral
 +∞
f (x)dx (B.1)
−∞

is a real or complex number that depends on the special function f (x) in the
integrand; other functions in general give other values for the integral. This
dependence of the integral on the function f is called a functional dependence
and the integral itself, correspondingly, a functional of f :
 +∞
F [f ] = f (x)dx . (B.2)
−∞

In order to stress that not a special value of f , but instead the whole function
f is the argument of the functional, square parentheses are used here to show
this functional dependence.
In physics one often deals with such functionals. For example, in classical
mechanics the action 
S = L(q(t), q̇(t))dt , (B.3)

integrated along a trajectory q(t) plays a special role. The action depends
obviously on the trajectory and is thus a functional integral of it: S = S[q(t)].
The Lagrange equations of motion are derived by investigating the
changes of S with the trajectory q(t): the physical trajectory leads to a
stationary value of S. To find this stationary value, and thus the physical
trajectory, requires taking the derivative of S[q(t)] with respect to the entire
function q(t). We thus need to define also the so-called functional derivative.

B.2 Functional Integration


A functional integration can be introduced as an integration over a space of
functions
198 B Functionals

df F [f (x)] , (B.4)

which is defined as a limit n → ∞ of an integration over finite n-dimensional


subspaces. A typical example is given by the path integral for quadratic
Hamiltonians
t
 i
h̄ L(x,ẋ)dt
K (x, t; xi , ti ) = N Dx e ti

  n+1
m 2
= lim
n→∞ i2πh̄η
⎧   ⎫
 
n ⎨i  n 2 ⎬
m xj+1 − xj
× dxk exp η − V (x̄j ) . (B.5)
⎩ h̄ 2 η ⎭
k=1 j=0

B.2.1 Gaussian Integrals

In this book we are often confronted with the evaluation of path integrals
over exponential functions whose exponents are quadratic in the coordinates.
It is thus useful to generalize the Gaussian integral formula


+∞ 
− 12 ax2 2π
e dx = (a > 0) (B.6)
a
−∞

to this case.
From the definition of the path integral it is obvious that we have to
consider products of such integrals
  n
 
1 (2π)n
exp − ak xk dx1 dx2 . . . dxn = n √ .
2
(B.7)
2 k=1 ak
k=1

We next assume that the n numbers ak are all positive and form the elements
of a diagonal matrix A. We thus have
n

det(A) = ak (B.8)
k=1

and
n
 n

ak x2k = xk Akk xk = xT Ax , (B.9)
k=1 k=1

where x is a column vector


B.2 Functional Integration 199
⎛ ⎞
x1
⎜ x2 ⎟
⎜ ⎟
x=⎜ . ⎟ .
⎝ .. ⎠
xn

Thus (B.7) becomes


 n
− 12 xT Ax n (2π) 2
e d x=  . (B.10)
det(A)

So far, we have derived this equation only for a diagonal matrix A. It


is, however, valid for a more general class of matrices. This can be seen by
noting that for each real, symmetric matrix B there exists a real, orthogonal
matrix O that diagonalizes B to A

OT BO = A (O real, orthogonal) (B.11)

or, equivalently

B = OAOT (B real, symmetric) . (B.12)

We thus get
 
T T
OAOT y
e− 2 y By
dny = e− 2 y dny
1 1

 n
T (2π) 2
e− 2 x Ax
dnx = 
1
=
det(A)
n
(2π) 2
=  , (B.13)
det(B)

where we have substituted y = Ox and have used the fact that the Jacobian
of an orthogonal transformation is 1 (because det(O) = 1). The last step is
possible because the determinant of a matrix is invariant under an orthogo-
nal transformation. Equation (B.10) is thus valid also for general symmetric
matrices with positive eigenvalues; it can also be shown to hold for complex
matrices with positive real parts.
This result can also be extended to more general quadratic forms in the
exponent. For a one-dimensional integral of such type we have

+∞ 
−ap2 +bp+c π b2 +c
e dp = e 4a , (B.14)
a
−∞

We now assume an n-dimensional integral over such a form where the inte-
grand is given by
1 T T
e−F (x) = e−( 2 x Ax+B x+C ) (B.15)
200 B Functionals

where A is a real symmetric matrix with positive eigenvalues, B is a vector


and C a constant. We bring F (x) into a quadratic form by writing
1
F (x) = (x − x0 )T A(x − x0 ) + F (x0 ) (B.16)
2
where x0 is given by
x0 = −A−1 B (B.17)
and F0 = F (x0 ) = C − 12 BT A−1 B. Setting now y = x − x0 gives
 
1 T T 1 T
e−( 2 x Ax+B x+C ) dnx = e− 2 y Ay−F0 dny
n
(2π) 2 T −1
e 2 B A B−C .
1
=  (B.18)
det(A)

Similar to the Gaussian integral (B.7) the following integral relation,


which can be proven by induction from n to n + 1 [5], holds also

+∞
:  ;
2 2 2
dx1 . . . dxn exp iλ (x1 − a) + (x2 − x1 ) + . . . + (b − xn )
−∞
  12  
in π n iλ
= exp (b − a)2 . (B.19)
(n + 1)λn n+1
Complex integrals. We can generalize these formulas to complex integration
by noting that (B.6) can be squared and then be written as
  
π
= e−ax dx e−ay dy = e−a(x +y ) dx dy .
2 2 2 2
(B.20)
a
Introducing now the complex variable z = x + iy gives

π 1 ∗
= e−az z dz ∗ dz . (B.21)
a 2i
This can be generalized as before to many coordinates (by replacing orthog-
onal matrices by unitary ones). We obtain
 
−z† Az n ∗ n ∗ (2πi)n
e d z d z = e−zi Aij zj dnz ∗ dnz = , (B.22)
det(A)
where A is a hermitian matrix with positive eigenvalues.
Another convenient, often used relation in this context is

ln det A = tr ln A , (B.23)

most easily proven for diagonal matrices. In this relation ln A is defined by


its power series expansion
B.3 Functional Derivatives 201

(A − 1)2 (A − 1)3
ln A = ln(1 + A − 1) = A − 1 − + − ... . (B.24)
2 3!
With the help of this relation we have


e−z Az dnz ∗ dnz = (2πi)n e−tr ln A
. (B.25)

B.3 Functional Derivatives


Suppose that F [f ] is a functional of the function f (x). The functional deriva-
tive of F is then defined by

δF [f (x)] F [f (x) + εδ(x − y)] − F [f (x)]


= lim . (B.26)
δf (y) ε→0 ε

A simple example is

+∞

F [f ] = f (x)dx . (B.27)
−∞

According to the definition just given we have


  
δF 1
= lim [f (x) + εδ(x − y)] dx − f (x) dx
δf (y) ε→0 ε

= δ(x − y) dx = 1 (B.28)

For an example relevant to the developments in this book let us take



F [f ] = ei f (x)x dx (B.29)

with
 
δF 1  i [f (x)+εδ(x−y)]x dx 
= lim e − ei f (x)x dx
δf (y) ε→0 ε
 
1 i f (x)xdx  iεy  i f (x)x dx
= lim e e − 1 = iy e . (B.30)
ε→0 ε

A second relevant example is given by

F [f ] = 2f (x) , (B.31)

where 2 is the d’Alembert operator (A.19). We get from the definition (B.26)

δF 1
= lim [2 (f (x) + εδ(x − y)) − 2f (x)] = 2δ(x − y) (B.32)
δf (y) ε→0 ε
202 B Functionals

A general formula that we will need quite often involves functionals F [J]
of the form
 
F [J] = dx1 . . . dxn f (x1 , . . . , xn ) J(x1 )J(x2 ) . . . J(xn ) (B.33)

with f symmetric in all variables. Then the functional derivative with respect
to J has the form

δF [J]
= n dx1 dx2 . . . dxn−1 f (x1 , x2 , . . . , xn−1 , x)
δJ(x)
× J(x1 )J(x2 ) . . . J(xn−1 ) . (B.34)

In our later considerations the functional may also be defined by a power


series expansion
∞ 
1
φ[J] = dx1 . . . dxn φn (x1 , . . . , xn )J(x1 )J(x2 ) . . . J(xn ) . (B.35)
n=1
n!

Then the functional derivative is given by


1
δ k φ[J] 1
1 1 
= φk (yp1 , yp2 , . . . , ypk ) , (B.36)
δJ(y1 )δJ(y2 ) . . . δJ(yk ) 1J=0 k! p

where the sum runs over all permutations p1 , . . . , pk of the indices 1, . . . , k.


If we assume that φk is a symmetric function under exchange of any of
the coordinates x1 , . . . , xk , then the functional derivative is given by
1
δ k φ[J] 1
1 = φk (y1 , . . . , yk ) , (B.37)
δJ(y1 )δJ(y2 ) . . . δJ(yk ) 1J=0

just as in a normal Taylor series.


C Renormalization Integrals

In Chap. 9.4 we encountered divergent integrals in the calculation of higher-


order two- and fourpoint functions. To evaluate these integrals analytically
in n dimensional Minkowski space is the purpose of this appendix. We start
with a discussion of the Gamma function whose properties play a role in the
evaluation of the integral and its expansion into n ≈ 4 dimensions. A detailed
collection of these properties can be found in [17].

Properties of the Gamma function. The Gamma function is defined by an


integral representation
 ∞
Γ (z) = e−t tz−1 dt ; (C.1)
0

it is single-valued and analytic everywhere except at the points z = n =


0, −1, −2, . . . where it has a simple pole with residue (−1)n /n!. It can thus
be expanded, e.g., around z = 0
1
Γ (z) = − γ + O(z) , (C.2)
z
where γ is known as the Euler–Mascheroni constant (γ ≈ 0.577 . . .). Since
the Gamma function also obeys the relation

Γ (z + 1) = zΓ (z) = z! (C.3)

we obtain for small z


 
Γ (z) 1 1
Γ (−1+z) = ≈ −(1+z) − γ + O(z) = − −1+γ +O(z) . (C.4)
z−1 z z
Evaluation of integrals over powers of propagators in n dimensions. The
typical integral to be evaluated reads
  
1 1
Il = dnq l
= dq dq0 l
. (C.5)
(q 2 − m2 + iε) (q02 − q 2 − m2 + iε)

with one timelike (q 0 ) and n − 1 spacelike (q k ) coordinates; the vector symbol


denotes a (n − 1)-dimensional vector q = (q 1 , q 2 , . . . , q n−1 ). l is a positive,
integer parameter and m2 a real, positive number.
204 C Renormalization Integrals

We first consider the integration over q0 . The situation here is exactly


as in Sect. 6.1.2. The integrand has its only /
poles in the
 second
0 and fourth
quadrant of the complex q0 plane at q0 = ± q 2 + m2 − iδ (cf. Fig. 6.1).
Since the integral behaves as 1/(q0 )2l−1 for large q0 the integration along
the q0 -axis can be closed in the lower half of the complex q0 plane without
changing the integral’s value.
According to Cauchy’s theorem this integration contour can now be chan-
ged into one that runs along the imaginary q0 axis and closes the contour in
the right half of the complex q0 plane. Because the contour still encloses the
same pole (the one in the fourth quadrant of the complex plane) the value of
the integral does not change.
This then gives the equality
 +∞  +i∞
1 1
dq0 l
= dq0 l
(C.6)
−∞ (q0 − q − m + iε)
2 2 2 −i∞ (q0 − q − m2 + iε)
2 2

since in both cases the contribution of the half-circle that closes the contour
vanishes. In the integral on the rhs the integration runs along the imaginary
axis in the complex q0 plane. On that axis q0 is purely imaginary; the inte-
grand has thus been analytically continued from the originally purely real q0
to a complex one.
Using the transformations (6.21), (6.22)

qn = −iq0 dn qE ≡ dqn dq = −idn q , (C.7)

with real qn we obtain the integral


  +∞ 
l 1 1
Il = (−) i dq dqn l
= (−) i dn qE
l
l
(C.8)
−∞ 2 2
(qE + m ) 2
(qE + m2 )
with
n
  i 2
2
qE = q = q 2 + qn2 . (C.9)
i=1

(C.7) is just the Wick rotation discussed in Sect. 5.1.1.


We now introduce “polar coordinates” by defining an n − 1 dimensional
solid angle element dΩn by the relation
n−1
dnqE = qE dqE dΩn . (C.10)

and get for the integral1


 ∞
l 1
Il = (−) i dΩn q n−1 dq l
. (C.11)
(q 2 + m2 )
0
1
In order to simplify the notation we are no longer denoting the Euclidean vectors
by the subscript E.
C Renormalization Integrals 205

The integral over the solid angle can be analytically performed and yields

2π n/2
dΩn =  n  . (C.12)
Γ 2

This can be seen by considering the n-th power of the Gaussian integral
 ∞ n  -n 2
√ n
dx e−x = dx1 dx2 . . . dxn e− k=1 xk
2
π =
 0  ∞  
1 ∞   n−2
xn−1 e−x dx = dΩn d(x2 ) x2 2 e−x
2 2
= dΩn
2 0
 0
1 n
= dΩn Γ . (C.13)
2 2

Here the integral representation (C.1) of the Gamma function has been used
in the last step.
This gives for the integral
n ∞
l 2π 2 q n−1
Il = (−) i  n  dq l
. (C.14)
Γ 2 (q 2 + m2 )
0

The remaining integral can be evaluated with the help of Euler’s β func-
tion [17]
∞
Γ (x)Γ (y)
B(x, y) = = 2 dt t2x−1 (1 + t2 )−x−y . (C.15)
Γ (x + y)
0

We obtain
 n  
q n−1 1 n n 1 Γ Γ l − n2
l
dq = mn−2l B( , l − ) = mn−2l 2
. (C.16)
2 2
(q + m ) 2 2 2 2 Γ (l)

Thus the complete integral is now given by


     
Γ n2 Γ l − n2 Γ l − n2
n
2π 2 1 n
Il = (−)l i  n  mn−2l = (−)l iπ 2 mn−2l .
Γ 2 2 Γ (l) Γ (l)
(C.17)
In this expression the dependence of the integral on the dimension n is ex-
plicit; it can analytically be continued to the physical case n = 4 where it
has a pole for l ≤ 2.
D Gaussian Grassmann Integration

A general proof of the Gaussian integration formula (10.45) for Grassmann


variables starts by considering the effects of an orthogonal transformation of
the Grassmann variables
 = OT η (D.1)
on the Gaussian integral
 
T T  T
I(n) = dη1 . . . dηn e−η M η = dη1 . . . dηn e−η (OM O )η

T 
= d1 . . . dn J e− M  (D.2)

with J being the Jacobian of the transformation (D.1). Since O is orthogonal


we have [det(O)]−1 = J = 1.
Expanding the exponential in (D.2) gives
n 
(−) 2  n
I(n) =  n  d1 . . . dn T M   2 , (D.3)
2 !

because only that term in the expansion of the exponential can contribute in
a n-dimensional integral that has exactly n factors of i .
Because of the anticommutator properties of the Grassmann variables,
M can be assumed to be antisymmetric without any loss of generality. For
any antisymmetric (n × n) dimensional matrix M with even n an orthogonal
transformation
M −→ M  = OT M O (D.4)
can be found [11] that brings it into the block-diagonal form
⎛ ⎞
0 M1 0 0 ··· 0 0
⎜ −M1 0 0 0 ··· 0 0 ⎟
⎜ ⎟
⎜ .. .. ⎟
⎜ 0 0 0 M2 · · · . . ⎟
⎜ ⎟
 ⎜ . .. ⎟
.
M =⎜ 0 0 −M 0 · · · .. ⎟ ; (D.5)
⎜ 2 ⎟
⎜ . .. .. .. .. .. ⎟
⎜ .. . . . ··· . . ⎟
⎜ ⎟
⎝ 0 0 0 0 · · · 0 Mk ⎠
0 0 0 0 · · · −Mk 0
208 D Gaussian Grassmann Integration

for odd n the matrix M  has one more row and one more column with all
zeros. The form of M  shows clearly that

det(M ) = det(M  ) = M12 M22 · · · Mk2 (D.6)

for even n. For odd n the determinant vanishes.


We now assume that M  in (D.3) has the special, block-diagonal form

(D.5). Thus, we have Mαβ = 0, except for α = 2α , β = 2α − 1 or α =
 
2α − 1, β = 2α . This gives
n 
(−) 2  
I(n) =  n  d1 . . . dn 2α M2α  ,2α −1 2α −1

2 !

 n2
+ 2α −1 M2α  −1,2α 2α
n n 
2 2 (−) 2  
 n2
= n d1 . . . dn 2α −1 M2α  −1,2α 2α , (D.7)
2 !

since M  is antisymmetric. Here the index α runs from 1 to n/2; n is even.


We now perform the exponentiation
n 
n
2 2 (−) 2  
I(n) =  n  d1 . . . dn ... 2α −1 2α · · · 2ν  −1 2ν 
2 ! α β  ν
6 78 9
n sums
  
× M2α −1,2α M2β  −1,2β  · · · M2ν  −1,2ν  . (D.8)

Again, there can be no two equal ’s under the integral, if this integral is to
be nonzero. Since the i appear always pairwise, they can be commuted to
n
normal ordering (n · · · 1 ) with a sign-change  by (−) 2 . The product of the
 
factors M2α  −1,2α · · · M2ν  −1,2ν  just gives det(M  ). The sums then give
(n/2)! times the same result. Thus
 
n
I(n) = 2 2 d1 . . . dn n · · · 1 det(M  )
n
= 2 2 det(M  ) . (D.9)

Using (D.6) we get the desired result



T n
d1 . . . dn e− M  = 2 2 det(M ) . (D.10)

Equation (D.10) corresponds to (B.13) for commuting numbers (appropriate


for boson fields). Note that here – in contrast to the bosonic case – the
determinant appears in the numerator!
References

This is not a complete list of references. Instead I list – and briefly characterize
– here only a number of textbooks that have some discussions of path integral
techniques and that I personally have found most useful as a supplement to
this book. Only a few are actually quoted in the text. These references provide
more information on the topics treated in this manuscript and usually contain
rather complete lists of references to the original literature.
1. J.J. Sakurai, Modern Quantum Mechanics. Addison-Wesley Redwood City
1985; contains a very nice introduction into the non-relativistic path integral
formalism.
2. T.D. Lee, Particle Physics and Introduction to Field Theory. Harwood Chur
1981; very physical treatment of field theory and high-energy phenomenology.
3. C. Grosche, F. Steiner, Handbook of Feynman Path Integrals. Springer Berlin
1998; compact presentation of theory of path integrals and their history. Unique
in its description of evaluation techniques and its tables of analytically calcu-
lable path integrals. Complete list of references.
4. U. Mosel, Fields, Symmetries, and Quarks. 2nd rev. enl. ed., Springer Heidel-
berg 1999; represents an introduction into fundamentals of field theories and
gauge field theories for non-specialists. Companion to the present book.
5. L.H. Ryder, Quantum Field Theory. Cambridge University Press 1985; very
didactical presentation of modern field theories.
6. I. Montvay, G. Mnster, Quantum Fields on a Lattice. Cambridge Univ. Press
Cambridge 1994; rather complete book on theory and methods of Lattice Field
Theory.
7. H. J. Rothe, Lattice Gauge Theories, An Introduction. World Scientific Sin-
gapore 1992; nice introductory text, contains discussion of results at time of
publication and an introduction into thermal field theory.
8. N. Makri (1991), Feynman path integration in quantum dynamics. Comp. Phys.
Comm. 63, 389
9. C. Itzykson, J-B. Zuber, Quantum Field Theory. McGraw-Hill New York 1985;
nearly comprehensive book on field theory, not always easy to follow.
10. M.E. Peskin, D.V. Schroeder, An Introduction to Quantum Field Theory.
Addison-Wesley Reading 1995; didactically very good, rather comprehensive
book on quantum field theory. At many places the new standard book on this
topic.
11. Ta-Pei Cheng, Ling-Fong Li, Gauge Theory of Elementary Particle Physics.
Clarendon Press Oxford 1984; quite comprehensive book, well-structured, easy
to read.
210 References

12. P. Ramond, Field Theory, a Modern Primer. Addison-Wesley Redwood City


(1990); introduction to field theories, also modern aspects, builds entirely on
path integral formalism.
13. S. Pomorski, Gauge Field Theories. Cambridge University Press 1987; very
compact, but quite deep treatment.
14. D. Bailin, A. Love, Introduction to Gauge Field Theory. 2nd ed., Hilger, Bristol
(1994); very comprehensive treatment of gauge field theories based on path
integral methods from the start, didactical, contains modern aspects of field
theory beyond Standard Model.
15. H. Kleinert, Path Integrals in Quantum Mechanics, Statistics and Polymer
Physics, World Scientific Singapore 1995. Application of path integrals to wide
range of physics problems.
16. G. Roepstorff, Path Integral Approach to Quantum Physics. Springer Heidel-
berg 1996. Some emphasis on mathematical foundations.
17. M. Abramowitz, I.A. Stegun, Handbook of Mathematical Functions. Dover 1974;
a comprehensive book on properties of mathematical functions.
Index

S matrix 75 Euclidean
SU (N ) 168 – representation 56
n-point function 81 – propagator 11
1PI graph 105 Euler constant 114, 203

Action 10 Faddeev–Popov determinant 179


– Euclidean 11 – for Abelian gauge field theory 180
Adiabatic switching 78 – for non-Abelian gauge field theory
Asymptotic freedom 190 180, 181
Axial gauge 173, 181 Fermions
– normal mode expansion 138
Bethe–Salpeter equation 21 – propagator 137
Born approximation 20 Feynman
– lowest order 26 – gauge 161
Bosonization 149 – integral 203
Bosons Feynman propagator 60
– normal mode expansion 54
– Euclidean 64
– propagator 60
– for bosons 61
– for fermions 136
Confinement 190
– for photons 160
Correlation function 36, 81
– poles 61
Counter term 120
Covariant derivative 50 Feynman rules
– for Abelian gauge fields 162 – φ4 theory 98
– for non-Abelian theory 169 – fermion loops 143
Covariant gauge 159, 181 – for interacting theory 141
Current conservation 49 – for non-Abelian gauge field theories
184
Divergence – for QED 164
– degree of 110 Field
– momentum 48
Effective potential 70 – – canonical 41
Electromagnetic field – quantization 77
– normal mode expansion 165 – renormalization 120
Electroweak interactions 190 – tensor 42
– Lagrangian 191 Fields
– multiplets 191 – asymptotic 76
Energy shell 17 – Dirac 45
Energy-momentum tensor 47 – electromagnetic 41
212 Index

– Klein–Gordon 44 – – integration 128, 207


– massive vector 43 – fields 134
– normal mode expansion 54, 77, 138, Green’s function
165 – connected 86, 92
Flavor 189 – for bosons 81
Functionals 197 – for electromagnetic field 160
– derivative 201 – for fermions 139
– integration 197 – Schrödinger equation 3
– power series expansion 202 Gribov copy 175
Groundstate–groundstate transition
Gamma function 203 amplitude 30
Gauge
– axial 181 Hamiltonian 41
– condition 173 Higgs field 192
– Feynman 161
– fixing 159 Lagrangian
– Landau 161 – density 39
Gauge condition – equations of motion 41
– generalized 177 – renormalized 123
Gauge field 162 Lie groups 168
Gauge field theory – algebra 168
– Abelian 157 – representations 168
– non-Abelian 168 Loop expansion 70, 150
Gauge invariance Lorentz covariance 40
– integration measure 173
– Lagrangian 158 Mandelstam variables 117
Gauge transformation Mass renormalization 120
– global 49 Maxwell equation 41
– local 162 Metric 194
– non-Abelian 168 Metropolis algorithm 72
Gaussian integration 10, 199–201 Minimal substitution 162, 169
– Euclidean 65
– for fermions 130 Noether’s theorem 46, 48
Gell-Mann matrices 190 Normal mode expansion
Generating functional 31 – electromagnetic field 165
– for Abelian gauge field theories 161 – fermions 138
– for fields 55 – scalar fields 54, 77
– for non-Abelian gauge field theories
176, 180 One-particle-irreducible graph 105
– free scalar fields 61
Ghost field 182 Path integral 8
– couplings 187 – Lagrangian 10
Glueball 190 – normalization 10
Gluon 190 – quadratic 9
Grassmann – semiclassical approximation 14
– algebra 125 Pauli principle 125
– – complex elements 133 Powercounting 110
– – derivative 127 Propagator
– – generators 125 – dressed 102
Index 213

– electromagnetic field 160 – regularized 115


– for bosons 61 Stationary Phase method 67
– for fermions 137 – generating functional 69
– for vector fields 160 Symmetries 46
– Schrödinger equation 3 – conserved current 48
– – free 17 – geometrical 46
– – perturbation theory 18 – internal 49
Symmetry factors 99
Quantum Chromodynamics 171, 189
– fundamental triplet 190 Tadpole diagram 102
Quantum Electrodynamics 49 Time development operator 5
Quarks 189 Time ordering operator 35
Tree graph 118
Reduction theorem 81 Two-point function
– for fermions 138 – for bosons 89
Regularization 113 – for fermions 140
– dimensional 113 – for photons 161
Renormalizable theory 113 – momentum representation 101
Renormalization
– coupling 123 Universality 171
– field 123
– mass 120 Vacuum
– scheme 119 – contributions 101
Representation – process 99
– Heisenberg 5 – state 75
– Schrödinger 5 Vertex function 82, 105
Running coupling constant 123
Weinberg’s Theorem 111
Saddle point method 67 Weyl ordering 12
Scalar field Wick rotation 29, 64, 65
– complex 50 Wick’s theorem 90, 145
Schrödinger equation 3
self-energy 104 Yukawa theory 147

You might also like