Вы находитесь на странице: 1из 160

Quantum Field Theory I

by

Prof. Michael G. Schmidt

24 October 2007

Processed and LATEX-ed by Olivier Tieleman


Supported by
Adisorn Adulpravitchai and Jenny Wagner

This is a first version of the lecture notes for professor Michael Schmidts
course on quantum field theory as taught in the 2006/2007 winter semester.
Comments and error reports are welcome; please send them to:
o.tieleman at students.uu.nl

Contents
1 Introduction
1.1 Particle field duality . . . . . . . . .
1.2 Short repetition of QM . . . . . . . .
1.2.1 Mechanics . . . . . . . . . .
1.2.2 QM States . . . . . . . . . .
1.2.3 Observables . . . . . . . . . .
1.2.4 Position and momentum . . .
1.2.5 Hamilton operator . . . . . .
1.3 The need for QFT . . . . . . . . . .
1.4 History . . . . . . . . . . . . . . . .
1.5 Harmonic oscillator, coherent states
1.5.1 Classical mechanics . . . . . .
1.5.2 Quantization . . . . . . . . .
1.5.3 Coherent states . . . . . . . .
1.6 The closed oscillator chain . . . . . .
1.6.1 The classical system . . . . .
1.6.2 Quantization . . . . . . . . .
1.6.3 Continuum limit . . . . . . .
1.6.4 Ground state energy . . . . .
1.7 Summary . . . . . . . . . . . . . . .
1.8 Literature . . . . . . . . . . . . . . .

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

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

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

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

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

5
5
5
6
6
7
7
8
9
12
12
12
13
14
18
18
19
21
23
24
24

2 Free EM field, Klein-Gordon eq, quantization


2.1 The free electromagnetic field . . . . . . . . . . . . . .
2.1.1 Classical Maxwell theory . . . . . . . . . . . .
2.1.2 Quantization . . . . . . . . . . . . . . . . . . .
2.1.3 Application: emission, absorption . . . . . . . .
2.1.4 Coherent states . . . . . . . . . . . . . . . . . .
2.2 Klein-Gordon equation . . . . . . . . . . . . . . . . . .
2.2.1 Lagrange formalism for field equations . . . . .
2.2.2 Wave equation, Klein-Gordon equation . . . . .
2.2.3 Quantization . . . . . . . . . . . . . . . . . . .
2.2.4 The Maxwell equations in Lagrange formalism

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

27
27
27
30
31
32
32
32
35
36
37

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

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

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

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

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

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

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

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

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

CONTENTS

3 The Schr
odinger equation in the language of QFT
3.1 Second Quantization . . . . . . . . . . . . . . . . . . . .
3.1.1 Schr
odinger equation . . . . . . . . . . . . . . . .
3.1.2 Lagrange formalism for the Schrodinger equation
3.1.3 Canonical quantization . . . . . . . . . . . . . .
3.2 Multiparticle Schrodinger equation . . . . . . . . . . . .
3.2.1 Bosonic multiparticle space . . . . . . . . . . . .
3.2.2 Interactions . . . . . . . . . . . . . . . . . . . . .
3.2.3 Fermions . . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

41
41
41
42
43
45
45
47
48

4 Quantizing covariant field equations


51
4.1 Lorentz transformations . . . . . . . . . . . . . . . . . . . . . 51
4.2 Klein-Gordon equation for spin 0 bosons . . . . . . . . . . . . 52
4.2.1 Klein-Gordon equation . . . . . . . . . . . . . . . . . . 52
4.2.2 Solving the Klein-Gordon equation in 4-momentum
space . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
4.2.3 Quantization: canonical formalism . . . . . . . . . . . 55
4.2.4 Charge conjugation . . . . . . . . . . . . . . . . . . . . 58
4.3 Microcausality . . . . . . . . . . . . . . . . . . . . . . . . . . 59
4.4 Nonrelativistic limit . . . . . . . . . . . . . . . . . . . . . . . 60
5 Interacting fields, S-matrix, LSZ
5.1 Non-linear field equations . . . . . . . . . .
5.1.1 Relation to observation . . . . . . .
5.2 Interaction of particles in QFT . . . . . . .
5.2.1 The Yang-Feldman equations . . . .
5.3 Greens functions . . . . . . . . . . . . . . .
5.4 Spectral representation of commutator vev .
5.5 LSZ reduction formalism . . . . . . . . . . .
5.5.1 Generating functional . . . . . . . .

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

63
63
64
64
65
66
69
69
71

6 Invariant Perturbation Theory


6.1 Dyson expansion, Gell-Mann-Low
6.2 Interaction picture . . . . . . . .
6.2.1 Transformation . . . . . .
6.2.2 Scattering theory . . . . .
6.2.3 Background field . . . . .
6.3 Haags theorem . . . . . . . . . .

formula
. . . . .
. . . . .
. . . . .
. . . . .
. . . . .

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

73
73
76
76
78
80
82

7 Feynman rules, cross section


7.1 Wick theorem . . . . . . . . . . .
7.1.1 4 -theory . . . . . . . . .
7.2 Feynman graphs . . . . . . . . .
7.2.1 Feynman rules in x-space

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

83
83
85
86
86

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

CONTENTS

7.3
7.4

7.5

7.2.2 Feynman rules in momentum space . . . . . . . . . . . 90


7.2.3 Feynman rules for complex scalars . . . . . . . . . . . 91
Functional relations . . . . . . . . . . . . . . . . . . . . . . . 92
Back to S-matrix elements . . . . . . . . . . . . . . . . . . . . 93
7.4.1 2-point functions . . . . . . . . . . . . . . . . . . . . . 93
7.4.2 General Wightman functions . . . . . . . . . . . . . . 95
From S-matrix to cross section . . . . . . . . . . . . . . . . . 96
7.5.1 Derivation . . . . . . . . . . . . . . . . . . . . . . . . . 96
7.5.2 Scattering amplitudes and cross sections: an example 98
7.5.3 The optical theorem . . . . . . . . . . . . . . . . . . . 100

8 Path integral formulation of QFT


8.1 Path integrals in QM . . . . . . . . . .
8.1.1 Vacuum expectation values . .
8.2 Path integrals in QFT . . . . . . . . .
8.2.1 Framework . . . . . . . . . . .
8.2.2 QFT path integral calculations
8.2.3 Perturbation theory . . . . . .

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

103
103
107
108
108
110
112

9 Lorentz group
115
9.1 Classification of Lorentz transformations . . . . . . . . . . . . 115
9.2 Poincare group . . . . . . . . . . . . . . . . . . . . . . . . . . 119
10 Dirac equation
10.1 Spinor representation of the Lorentz group . . . . . .
10.2 Spinor fields, Dirac equation . . . . . . . . . . . . . .
10.3 Representation matrices in spinor space . . . . . . .
10.3.1 Rotations . . . . . . . . . . . . . . . . . . . .
10.3.2 Lorentz transformations . . . . . . . . . . . .
10.3.3 Boosts . . . . . . . . . . . . . . . . . . . . . .
10.4 Field equation . . . . . . . . . . . . . . . . . . . . . .
10.4.1 Derivation . . . . . . . . . . . . . . . . . . . .
10.4.2 Choosing the -matrices . . . . . . . . . . . .
10.4.3 Relativistic covariance of the Dirac equation .
10.5 Complete solution of the Dirac equation . . . . . . .
10.6 Lagrangian formalism . . . . . . . . . . . . . . . . .
10.6.1 Quantization . . . . . . . . . . . . . . . . . .
10.6.2 Charge conjugation . . . . . . . . . . . . . . .

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

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

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

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

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

123
123
125
126
126
127
128
130
130
131
132
133
135
136
137

11 Fermion Feynman Rules


11.1 Bilinear Covariants . . . . . .
11.2 LSZ reduction . . . . . . . . .
11.3 Feynman rules . . . . . . . .
11.3.1 The Dirac propagator

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

139
139
140
141
141

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

CONTENTS
11.4 Simple example in Yukawa theory . . . . . . . . . . . . . . . . 144

12 QM interpretation of Dirac equation


12.1 Interpretation as Schrodinger equation . .
12.2 Non-relativistic limit . . . . . . . . . . . .
12.2.1 Pauli equation . . . . . . . . . . .
12.2.2 Foldy-Wouthousen transformation
12.2.3 F.-W. representation with constant
12.3 The hydrogen atom . . . . . . . . . . . . .

. . .
. . .
. . .
. . .
field
. . .

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

149
149
152
152
154
155
156

Chapter 1

Introduction
1.1

Particle field duality

From classical physics we know both particles and fields/waves. These are
two different concepts with different characteristics. But some experiments
show, that there exists a duality between both.
1. electromagnetic waves
waves photons (photoelectric effect, E = h = ~)
fields are quantized, consisting of particles called photons.
2. particles (e.g. electrons) may exhibit interference phenomena, like
waves. Thus, particles must be described by a wavefunction . However, this has a probabilistic interpretation, it is not like an electromagnetic field.
The latter leads to quantum mechanics (QM), the former to quantum
field theory (QFT). QM is nonrelativistic, and describes systems with fixed
particle number. The quantization of the electromagnetic field requires
quantum field theory, but is based on the same principles as quantum mechanics.

1.2

Short repetition of QM

QM cannot be derived from mechanics; rather, mechanics should follow from


QM. But in obtaining appropriate Hamiltonians in QM, the correspondence
principle, which substitutes quantities from mechanics by quantum mechanical operators, plays a key role.
7

1.2.1

CHAPTER 1. INTRODUCTION

Mechanics

In the Lagrangian formulation of mechanics, we substitute the equations of


motion by an extremal postulate for an action functional
Z t
S[L] =
dtL
(1.1)
t0

of the Lagrange function L(qi , qi ), where the qi are the (finitely many) generalized coordinates in the specific problem. Postulating S = 0 for variations
in the qi (t) and keeping the endpoints fixed, we obtain the Lagrange equations
d L
L

=0
(1.2)
qi dt qi
Here, L/ qi = pi are the generalized canonical momenta. The Hamiltonian
H(pi , qi ) is the Legendre transform of L:
X
H=
qi pi L
(1.3)
i

and the Hamiltonian equations in phase space follow:


pi =

H
H
, qi =
qi
pi

(1.4)

The Poisson bracket of two functions f (pi , qi ), g(pi , qi ) in phase space is


defined as

X  f g
f g
{f (pi , qi ), g(pi , qi )}P oisson =

(1.5)
pi qi qi pi
i

We have
{pi , qj }P oisson = ij

(1.6)

and the Hamiltonian equations can be generalized to


d
f (pi , qi ) = {H, f }P oisson
dt

(1.7)

In case of explicitly time-dependent f we have f(pi , qi ) = {H, f }P oisson +


f /t. For these formal aspects see e.g. F. Scheck, Mechanik.
Continuum mechanics can be obtained by taking the number of coordinates N to infinity, as will be seen in a specific example in QM.

1.2.2

QM States

States are described by Hilbert space (ket) vectors |i H (or by the density
matrix ; see below) with the following properties:

1.2. SHORT REPETITION OF QM

1. representation space: (~x) = h~x|i are functions in the L2 Hilbert


space H. These are coordinates in the h~x|-basis.
2. probabilistic interpretation:
|(~x)|2 d3 x is the probability to find the particle in the state |i
in volume element d3 x.
R
R
h|i = d3 x (~x) (~x) = d3 p (~
p) (~
p) is called the inner
product of and
| h|i |2 is the probability to find the state |i in |i, and vice
versa.

1.2.3

Observables

Observables are described by self-adjoint linear operators A in H: A = A


and def(A) = def(A ).
The eigenstates of A are orthogonal and form a complete basis, and the
eigenvalues are real. The expectation value of an observable A in a state
|i is given by
hAi = h|A|i
More generally, one can introduce a density matrix , and obtain
hAi = Tr(A)
with
:

+ = ,

0,

Tr() = 1

for general mixed states, and


= P = |i h|
for pure states.

1.2.4

Position and momentum

and momentum P
fulfill the canonical (Heisenberg) commutation
Position X
relations
h
i

Xk , Pl = i~kl
(1.8)
in accordance with the general quantization rule

10

CHAPTER 1. INTRODUCTION
i~ {A, B}P oisson [A, B]

(1.9)

where the Poisson bracket for f (x, p), g(x, p) is defined in eq. (1.5).

1.2.5

Hamilton operator

The correspondence principle relates mechanics to QM:

~
p~ ~i

mechanics
QM in x-space
E i~ t

~x ~x
So:

~2
p~2
~2
+ V (~x) H =
+ V (~x),
2m
2m
and similarly for other operators corresponding to observables.
H=

Time development (without measurement!)




H
|(t)i = exp i t |(0)i
~
|
{z
}

(1.10)

=:U

This is the representation of time development in the Schrodinger picture. The exponential function is a unitary operator (U = U1 ).
In the Schr
odinger picture, the states are time-dependent, and operators
are time-independent (except when explicitly time-dependent). In contrast,
in the Heisenberg picture the states are time-independent, and the operators
are time-dependent. The expectation values are the same:
D
E
h(t)|AS |(t)i = (0)|U AS U|(0) = h(0)|AH |(0)i = hH |AH |H i
where the Hamilton operator H is the same as above. The (Heisenberg)
equation for time development is given by
d
i
AH (t) = [H, AH (t)]
dt
~
(with an additional term A/t in case of explicit time dependence in
A).
In the Heisenberg picture,
D
E D
E D
E
S |S (t) = X
S |U(0) =! X
H |H (0)
(~x, t) = X
S and X
H are as follows:
The actions of the operators X

1.3. THE NEED FOR QFT

11

S |xS i = ~xS |xS i


X

|xiH

|xH i

(1.11)

U X U U |x i = ~x U |xS i
| {zS } | {z S}
| {z }
H
X

Remark
In QM, multiparticle states can be represented, in spaces like
H1 H2 HN
which describe the whole space as a tensor product of the individual
spaces of each of the N particles. This representation is used to describe
e.g. atomic structure, nuclear shells, or solid state physics, but the particle
number N is always fixed!

1.3

The need for QFT

QFT is the quantum theory of fields, the main difference to QM being the
huge number of degrees of freedom ( ). The principles, however, are the
same as those of QM. There is a (multiparticle) Hilbert space, called Fock
space, and a probability interpretation, all as we know them from QM. So
dont worry!
The electromagnetic wave equation is the prototype of a relativistic field
equation:


2
~ 1

c t2
2


A(~x, t) = 0

(1.12)

It can be solved by a wave ansatz, which leads to:




2
k 2
c
2

Ak ei(tk~x) = 0

(1.13)

With p~ = ~~k and E = ~ we get the dispersion relation for photons in


the particle language:


k2

2
c2

p2

E2
=0
c2

(1.14)

Note that the wave equation is not a kind of Schrodinger equation for the
probability amplitude of photons. In this case it would be an equation for the
probability amplitude of a single photon, which would lead to contradictions.
Consider the physics:

12

CHAPTER 1. INTRODUCTION
It is very easy to produce soft or collinear quanta
soft: E p small
p
~

collinear:

p
~

p
~

with p~ = p~1 + p~2 and p~ || p~1 || p~2

We want to measure x with precision x. Assume x DeBroglie =


~/p. Multiplying with p and using the uncertainity relation we obtain x p ~ p p. This means that new particles may be
produced, because E 2 = p2 c2 for massless (highly relativistic) particles.
More formally: the wave equation contains second derivatives with respect to time, which means thatp
a probability interpretation like the one for
the Schr
odinger equation fails ( m2 c2 + p2 is nonlocal).
Remarks
The same problems arise for massive relativistic particles (e.g. pion
~
,0 , e): we want to measure x with precision x mc
= Compton
Again using x p ~, we now obtain
p mc
With the relativistic relation
E 2 = p2 c2 + m2 c4

2EE = 2ppc2

we see that
E = vp & mcv
This allows particle production for v c. Thus, a particle can not be
localized without allowing for the production of further particles.
(~x, t) assumes that one can measure ~x arbitrarily exactly, but we have
just seen that then the particle number is not conserved. QM emerges
for non-relativistic massive particles in the limit of neglegible particle
production (i.e., it is a special case of QFT). However, also for very
slow massive particles there are quantum field theoretical corrections:
an example is the uncertainity relation Et ~, which leads to
tunneling and particle production. For small times t a very high
energy E is possible, i.e. at small time scales we cannot exclude
particle production.

1.3. THE NEED FOR QFT

13

Figure 1.1: Particle production

In short

Relativistic field equations require to be treated in the framework of QFT,


where particles can be produced and annihilated. Their interpretation is
different from that of the Schr
odinger equation. The electromagnetic field
has nothing to do with the localization probability of a single photon.

Still: The principles of QM remain true!

Other relativistic field equations are the Klein-Gordon equation and the
Dirac equation. The Dirac equation, which contains only first order time
derivatives, allows for a one particle interpretation in the nonrelativistic
limit, although not without further ado, as will be seen later.

We will first discuss free particles, and later their interactions, almost
exclusively in the context of perturbation theory. Particularly interesting are
gauge theories (electrodynamics, chromodynamics, flavordynamics).

This might give the impression that QFT was made exclusively for fundamental theories, for elementary particle physics. However, it is also very
important in statistical mechanics and solid state physics (see e.g. the role
of path integrals and their relation to the partition function); and of course
historically, the connection between QFT and relativity was very important!

14

1.4

CHAPTER 1. INTRODUCTION

History

1925/1926:
1926:
1927:
1928:
1929/1930:
1930:
1948:
1949:
1954:
1955:
1957:
1966:
1969:

1.5
1.5.1

Heisenberg and Schrodinger develop their QM formalisms,


which are proven to be equivalent
Bohm, Heisenberg and Jordan begin developing QFT
Dirac postulates his equation and explains spontaneous emission
Jordan and Wigner (anti-commutation relations, Pauli principle)
Heisenberg and Pauli develop the canonical formalism
Dirac works on hole theory and antiparticles; the positron is
detected
Schwinger, Feynman and Tomonaga publish on the Lamb shift;
the anomalous magnetic moment e is explained
Dysons work leads to a better understanding of the Feynman
graph rules
Yang and Mills publish their (at the time mostly unnoticed)
article on non-Abelian gauge theories
Lehmann, Szymanzik and Zimmermann (S-matrix in QFT)
Bogoliubov and Parasiuk (renormalisation)
Hepp and ...
... Zimmermann work out renormalisation to all orders

Harmonic oscillator, coherent states


Classical mechanics

In classical mechanics, we know the Hamiltonian of the harmonic oscillator:


H=

1 2 m 2 2
p +
x
2m
2

with the equations of motion


x =

H
p
=
p
m

p =

H
= m 2 x
x

(see Hamilton equations, sec. 1.2.1). By a canonical transformation, we


obtain the holomorphic representation:

p
a
= ( mx + i
)/ 2
m

p
a
= ( mx + i
)/ 2
(1.15)
m
x
, p

a
, i
a
In terms of these new variables, the Hamiltonian becomes:
H=


(
a a
+a
a
)
2

1.5. HARMONIC OSCILLATOR, COHERENT STATES

15

The equations of motion are:


H
=
a

a
H
= i
a
(i
a )

ia
=
a
=

(1.16)

These are first order differential equations, allowing us to solve for a


, a
:
a
(t) = eit a
(0)

a
(t) = e

1.5.2

(1.17)

it

a
(0)

Quantization

We use the quantization rule known from QM:


i
commutator
~

Poisson-Bracket
For example:

i
[P, X] = 1
~
f
i
f
x +
p = {H, f }P f(X, P) = [H, f ]
f(x, p) =
x
p
~
{p, x}P = 1

When we quantize the harmonic oscillator, we promote a


and a
to operators:
a
, a

a,
a

which obey the usual commutation relation:


ih i
i
a ,
a =1
~
In QM we will often find ~ included in a
. So we can set ~ = 1, or define a
new a, to make the equations look nicer:

a
a :=
~


H = ~ a a + aa
|
{z
}

i
a, a = 1


1

H = ~ a a +
2

a a+1

The occupation number operator N is defined as


N := a a

16

CHAPTER 1. INTRODUCTION

and has eigenvalues


N |ni = n |ni

n = 0, 1, ...

|0i is the ground state. In this setting a and a are the annihilation and
creation operators, and have the following actions:

a |ni =
n |n 1i
(1.18)

a |ni =
n + 1 |n + 1i
(1.19)

1.5.3

Coherent states

The states |ni do not correspond to quasiclassical states. The closest approximations to classical states are called coherent states, which, in the
Heisenberg picture, which we will use most frequently, are defined as follows:
|i = N e(t)a
From
i~

(t)

|0i

(1.20)

h
i

a = H, a = a ~
t

we obtain

a (t) = eit a (0) (t) = eit (0)

which, if plugged into the definition of a coherent state, gives:


|i = N

X
n a+n

n!

n=0

X
n
|ni
|0i = N
n!
n=1

To determine the normalization N , we take the inner product of |i with


itself:
2

h|i = |N |

X
||2n
n=0

n!

= |N |2 e|| = 1

N = e||

2 /2

(1.21)

If we now apply the annihilation operator a to such a coherent state, we


obtain:

X
n (0) n

a |i = N
|n 1i = |i
(1.22)
n!
n=1
We can also use this as a definition of coherent states. From the above,
it follows that the expectation values of a and a are given by
D
E
h|a|i =
|a | =

1.5. HARMONIC OSCILLATOR, COHERENT STATES

17

and
D

E
|a a| = ||2

Analogously, we can calculate a |i, which gives


d
a |i =
|i
d

(1.23)

d
i.e., a and ia act on coherent states as and i d
.
For coherent states, the sum of the variances of X and P is minimal, i.e.
|i comes closest to classical motion.
For the variance in the occupation number N , we have:


(N )2 = |(N hNi)2 |


= |N2 | h|N|i2
D
E
|a aa a|
= ||4 + ||2
D
E
|a a|
= ||2

(N )2 = ||2 = hNi

N
1
=p
hN i
hN i

(1.24)

The relative variance goes to 0 for large N.


Remarks
Introducing coherent states:
In the Heisenberg picture, we have
a |i = |i
|iH = e||

2 /2

(t) a
e|{z}

(t)

|0i

eit

In the Schr
odinger picture, this becomes:
|iS = eiHt/~ |iH
= N e(t)a

(0)

|0i

Coherent states are not orthogonal:

0
0 2
0
| = e| | /2 ei=( )

(1.25)

18

CHAPTER 1. INTRODUCTION
The completeness relation holds for coherent states:
dd
|i h| = 1
2i

(1.26)

We can obtain this relation by expanding an inner product hn|mi in


the -basis and integrating:
Z
Z
dd
n m
dd
hn|i h|mi =
NN
= nm = hn|mi
2i
2i
n! m!
where in the second step we have used
m
h|mi = m ( ) = N
m!
The following normalization is often used in the literature:
h|i = e||

In this case, the completeness relation changes:


Z

dd
e
|i h| = 1
2i

The representation of an arbitrary operator A in the -basis is as


follows:

h|Af i

1
2i

|A|0 0 |f d0 d0
| {z }

A( ,0 )

A( , 0 )

=
=

with A


h|ni hn|A|mi m|0
( )n 0 m
2
0 2
Amn e|| /2 e| | /2 (with summing convention)
n! m!
n m
Knm a a
(normal ordered)

Knm ( , ) := Knm n 0

A( , 0 )

e(||

2 +|0 |2 )/2

K( , 0 )

Applying this to a and a , we plug in Knm = 0 except for n = 1,


m = 0 for a , and n = 0, m = 1 for a, in which cases K = 1. This
2
0 2
2
0 2
gives a ( , 0 ) = e(|| +| | )/2 and a( , 0 ) = e(|| +| | )/2 0

1.5. HARMONIC OSCILLATOR, COHERENT STATES

19

Canonical transformations in phase space are symplectic (see e.g. F.


Scheck, Mechanik). A general phase space vector

~zP h =

~x
p~

x1 , . . . , xN = z1 , . . . , zN
p1 , . . . , pN = zN +1 , . . . , z2N

has the following Hamilton equation:


d~zP h
H
=J
dt
~zP h
with

J=

0 1
1 0

a metric in phase space, J 1 = J

A canonical transformation ~z ~z0 preserves the Hamilton equation:

z0 =

M :=

z
z0

1
M
:=

z0
z

, = 1, . . . 2N

z0
z 0 z
z 0
H z0 !
H
H
1
= 0
= J
= M
J 0
= J 0
t
z t
z
z
z z
z
|{z}
|{z}
1
M

1
M

(1.27)

1
M
J M T = J

J = M JM T

M is a symplectic matrix (the M form a group)!


The Poisson bracket
{f, g}P oisson (z) =

f
g
J
z
z

is invariant under canonical (symplectic) transformations.

(1.28)

20

CHAPTER 1. INTRODUCTION

1.6
1.6.1

The closed oscillator chain


The classical system

Figure 1.2: The closed oscillator chain

Let us consider a circular chain of radius R, with N masses m, connected


by springs with spring constant D. Let the ith mass be at a displacement
qi = Ri from some rest position, and let us impose the periodic boundary
condition q0 qN for convenience of writing. The distance between the
masses when they are at rest is a = 2R/N .
The Lagrangian for this system is:
L(q, q)
=

N 
X
m

j=1

qj2

D
(qj+1 qj )2
2

Varying ql , we get the equation of motion


m
ql + D (ql ql1 (ql+1 ql )) = 0
which is solved by the ansatz
ql (t) = ei(lkt)

(+vt)

leading to
2

ik

+ik

m + |{z}
D 2e
e
{z
m
2 |
4 sin2

k
2

=0

Because of the periodic boundary condition qN = q0 , we get N k = 2n


with n = 0, . . . , N 1 or n = N/2, . . . , +N/2 1 (for even N ), and we
obtain for the oscillator modes

1.6. THE CLOSED OSCILLATOR CHAIN


n2 =

21

4D
n
sin2
m
N
|{z}

(1.29)

4
2

Superposition of modes gives the general solution for real qj (t):

1
qj (t) =
N

N/21

X h

cn ei(2nj/N n t) + cn ei(2nj/N n t)

(1.30)

n=N/2

(note that this is a Fourier expansion).


The Hamiltonian is then given by:
#
"
N
X
p2j
D
2
+ (qj qj1 ) ,
H(q, p) =
2m
2
j=1

where pj = mqj , and the corresponding equations of motion are


pj =

H
,
qj

qj =

H
.
pj

To ensure that the the transformation (qj , pj ) (cn , cn ) is canonical,

we normalize c
n and cn by defining the new canonical variables an and ian
through an = 2mn cn ein t (cf. harmonic oscillator, sec. 1.5.1). These
variables are canonical:
{an , iam }P oisson = nm ,

{a, a} = {a , a } = 0

The new Hamiltonian is


H=

X1
n

n (an an + an an )

Note that this transformation is still classical. It is preferable, however,


to discuss this in the already quantized version.

1.6.2

Quantization

The procedure is exactly the same as it was for the harmonic oscillator:

pl , ql pl , ql
i
{f (p, q), g(p, q)}P oisson
[f , g]
~
an , an
an ,
an
Then the
an , i
an fulfill the usual commutation relation

22

CHAPTER 1. INTRODUCTION

h
i

an ,
am = i~nm
and the Hamiltonian looks like this:

X 
1

H=
n
an
an + ~
2
n
a ,
a ] = i~nm and
Exercise: obtain the commutator [pj , qj 0 ] from [
PN 1 2in(jj 0 )/N n m
= N jj 0 ).
derive the Hamiltonian above (use n=0 e
In the following we will often use units where ~ = 1. We will sometimes
reinsert the factors of ~, in cases with experimental relevance; the same goes
for setting c = 1.
()
()
Defining an = ~
an , the Hamiltonian becomes:


X
1

(1.31)
H=
~n an an +
2
n
After quantization each mode n has an excitation number (the occupation number of some quasiparticle-mode), as we have already seen in the
case of the harmonic oscillator. This begs the question whether, as a physicist, one can distiguish these modes (called phonons if they live in crystals)
from real particles.
If one can see the lattice (e.g. a solid state crystal), the answer is yes,
because a real particle should exist independently of the medium it lives
in. One can not always see this medium, however, in particular when the
continuum limit has been taken. Then, the lattice has become some sort
of ether, which is not necessarily visible to the physicist, and might not
be needed anymore.
An interesting case is fundamental (super)string theory, where particles
are modes of strings.
The states on which the operators defined above act live in a Fock space
F = H1 H2 HN , where Hi the Hilbert space of i-th mode. The
vacuum state is denoted by |0i, and has am |0i = 0 for all m.
General states
(aN )nN
(a1 )n1 (a2 )n2
...
|0i
|n1 , n2 , . . . , nN i =

n1
n2
nN
are normalized eigenstates of H and have energy


X
1
E=
~m nm +
2
m
Note: Elastic binding to lattice, mimicking a crystal, at xj = j2R/N
gives a term (m/2)2 qj2 in L which shifts the n2 in eq. (1.29) by +2 ;
there is no zero mode.

1.6. THE CLOSED OSCILLATOR CHAIN

23

One can rewrite eq. (1.30) by substituting n for n in the second term:
N/21

qj =

"
2inj/N

n=N/2

an + an

2mn

N/21

e2inj/N Qn

n=N/2

Similarly, one can introduce


r
Pn =

mn
i(an + an )
2

Check the following relations:


[Pn , Ql ] = inl
and
H=

X 1
1
Pn Pn + mn2 Qn Qn
2m
2
n

with Pn = Pn and Qn = Qn

1.6.3

Continuum limit

In the continuum limit, we take N to infinity and a to zero while keeping


2R = N a fixed. The index j becomes a continuous variable, leading to:
qj (t) q(x, t),

x=aj

(1.32)

Note that in the closed string case discussed above, one does not have
elastic binding to the lattice like one has for a crystal, and x is defined up
to an overall translation.
Now, we make the following translations:

(qj qj1 )

X
j

1
a

q(x, t)
x

2R
Z

dx
0

Further, we postulate

m = a

D=
a

with finite mass density


with finite string density

24

CHAPTER 1. INTRODUCTION
and obtain the following Lagrangian:
Lcont

1
=
2

2R
Z

" 


 #
q(x, t) 2
q(x, t) 2
dx

t
x

(1.33)

The equations of motion


2 q(x, t)
2 q(x, t)

=0
t2
x2
can be read off from the discontinous case. Later, we will obtain them
directly from eq. (1.33).
p
Now, with c = / we can rewrite the above equation as:

2
2q
2 q

c
=0
t2
x2

(1.34)

Note: introducing elastic binding (cf. note in section 1.6.2) gives an


R 2R
additional mass term 12 0 dx2 q 2 (x, t) in eq. (1.33), and 2 q in eq.
(1.34), which then has the form of a Klein-Gordon equation to be discussed
in the next chapter.
Now, we can just translate our discrete solution to the continuum case,
or we can make the following ansatz (from electrodynamics):
q(x, t) = Aei(kxt)
Plugging this into eq. (1.34) gives
2 = c2 k 2
Imposing periodic boundary conditions q(0, t) = q(2R, t) leads to
kn =

n
R

(n = 0, 1, 2, . . . )

which, in combination with 2 = c2 k 2 , gives


|n|c
n =
R

(1.35)

Note that in the finite N case, the continuum n is only approximately


valid. We have to use the formula from the discrete case:
r


4D
|n|
n =
sin
(1.36)
m
N
We expand this for small n, using the linear approximation for the sine:
r
4D |n|
2c|n|
c|n|
n
=
=
m N
aN
R

1.6. THE CLOSED OSCILLATOR CHAIN

25

(remember D = c2 m/a2 and a = 2R/N ). This approximation is only

allowed when |n|


N < 4 ; then, we have:
cN
(1.37)
= c
R 4
Note that in the above discussion, c does not always have to be the speed
of light; it may also be a characteristic speed for the medium in which our
phonons are living.
n

1.6.4

Ground state energy


E0 =

X ~n
n

+
X

n=

|n|c
2R

(1.38)

In the continuum limit the ground state energy is divergent! We will observe such divergences more often in QFT; they are related to the continuum
limit and to interaction at a point, and require physical discussion.
In the finite N case:
N/21

E0 =

~
2
~
2

X
n=N/2

Z1/2
dx

cN
sin
R

|n|
N

c N2
sin(|x|)
R

(use x

n
)
N

1/2

E0
2R

~
1
cN 2
2
2
R 2R
2
~c N
2~c
=
with
2 3 R2
a2

N2
=
R2

2
a

2

Figure 1.3: Validity range for the linear approximation

26

CHAPTER 1. INTRODUCTION

The absolute value of E0 does not have a physical meaning (except in


general relativity, where it appears on the right hand side of the Einstein
equation), but we can compare ground state energy in two different physical
situations; in doing so, we can observe the Casimir effect.

1.7

Summary

1.8

Literature

Literature list for QFT I and II; books marked with * are particularly recommended for studies accompanying the lectures

1.8. LITERATURE
*L. H. Ryder
*M. E. Peskin,
D. E. Schroeder
*M. Maggiore
*C. Itzykson,
J. B. Zuber

*P. Ramond

*S. Weinberg
*K. Huang

*M. Stone
T. Kugo
E. Harris

F. Mandl, G. Shaw
V. P. Nair

A. Zee

L. S. Brown
S. J. Chang
J. D. Bj
orken, S. Drell

N. N. Bogoliubov,
D. V. Shirkov

27
Quantum Field Theory (Cambridge Univ. Press)
rather elementary, clear, wide variety of subjects
An introduction to QFT (Addison-Wesley)
A modern Introduction to QFT (Oxford Univ. Press)
Quantum Field Theory (Mc Graw Hill, also paperback)
comprehensive standard book, not always up to date,
hard to read in a row (gauge theories)
Field Theory, a modern Primer (Frontiers of Physics,
Benjamin)
early book on path integral formulation, non abelian
gauge theory, unconventional
The Quantum Theory of Fields I, II
very extensive, sometimes non-standard conventions
Quantum Field Theory: from operators to Path integrals (Wiley Interscience)
looking to QFT outside elementary particle physics
see also Quarks... (World Scientific Publish.)
The Physics of Quantum Fields (Springer)
Elementary, wide range
Eichtheorie (Springer)
Thorough discussion of gauge theory quantization
A Pedestrian Approach to QFT (Wiley, also in German)
rather elementary, short
Quantum Field Theory (Wiley Interscience)
Elementary introduction
Quantum Field Theory, A Modern Perspective
(Springer)
Interesting advanced chapters
Quantum Field Theory in a Nutshell (Princeton Univ.
Press)
unconventional presentation of concepts, diverse
range of topics
Quantum Field Theory (Cambridge University Press)
Introduction to QFT (World Scientific, Lecture Notes
in Physics, Vol. 29)
*I. Rel. Quantenmechanik II. Rel. Quantentheorie
(BI 98 and 101) (II. outdated, no QCD)
a standard book some decades ago
Quantum Fields (Benjamini, Cummings;..)
related to the famous monography of the same authors, which today is partly outdated

28
L. D. Fadeev,
A. A. Starnov
R. J. Rivers
J. Zinn-Justin

I. J. R. Aitchison

C. Nash
R. P. Feynman

J. M. Jauch,
F. Rohrlich
A. L. Fetter,
J. D. Walecka
A. Smilga
M. Le Bellac

CHAPTER 1. INTRODUCTION
Gauge Fields, Introduction to QFT (Benjamini,
Cummings)
advanced studies
Path Int. Methods in QFT (Cambdridge Mon. on
Math. Phys.)
Quantum Field Theory and Critical Phenomena (Oxford Science Publications)
the connection to statistical mechanics
Relativistic Quantum Mechanics (Mac Millan)
strong relation to elementary particle physics, elementary
Relativistic Quantum Fields (Academic Press)
renormalization theory for 4
Quantum Electrodynamic (Frontiers in physics, Benjamin. inl.)
unconventional, somewhat outdated
The Theory of Photons and Electrons (Springer)
old QED-monography
Quantum theory of Many Particle Systems (McGraw
Hill)
Feynman diagrams in solid-state physics
Lectures on Quantum Chromodynamics (World Scientific)
Quantum and Statistical Theory (Clarence Press, Oxford)
Thermal QFT

Chapter 2

The free electromagnetic


field; Klein-Gordon equation;
quantization
We quantize the free electromagnetic field in the Coulomb/radiation gauge,
analogously to the oscillator string. Then, we discuss electromagnetic emission and absorption in this language. We introduce the Lagrange- and
Hamilton-formalisms for fields and present a more thorough discussion of
the canonical quantization of the Klein-Gordon equation. To conclude, we
come back to the Maxwell equations in the canonical formalism.

2.1
2.1.1

The free electromagnetic field


Classical Maxwell theory

The classical source-free Maxwell equations are:

~ B
~ = 0,

~ E
~ = 0,

~ E
~ = 1 B,
~

c t

~ B
~ =1E
~

c t

In the Coulomb/radiation gauge, we have


~ A
~ = 0,

= 0,

so

~ = 1 A
~
E
c t

~ =
~ A
~ and Amp`eres law, we obtain
Using B




~
~
~ A
~ =
~
~ A
~
~
~A
~ = 1 A

c t
| {z }
=0

29

(2.1)

30CHAPTER 2. FREE EM FIELD, KLEIN-GORDON EQ, QUANTIZATION


resulting in


1 2
2 2
c t

~ x, t) = 0
A(~

(2.2)

for the wave equation.


In order to solve the wave equation in a finite volume L3 with periodic
boundary conditions, we make the following ansatz, based on a Fourier
transform (later on called Fourier ansatz):


X
~
~
~ x, t) = 1
A(~
N~k ~a~k (t) eik~x + ~a~k (t) eik~x
L3 ~

(2.3)

with normalization N~k and discrete ki L = 2ni .


~ A
~ = 0, we have ~k ~a~ = 0, i.e. the solution has only transversal
From
k
modes; this is the origin of the name transversal gauge. The wave equation
gives ~a~k eit , with dispersion relation c|~k| = . Now, let us calculate the
energy of the electromagnetic field:
Z


1
3
2
2
~
~
d x E +B
E =
8
!


Z
 

1
1 ~ 2 ~
3
~
~ A
~
d x
=
A + A
8
c2 t
Through partial integration, the second term becomes

 




~
~
~ A
~
~
~
~ A
~
(1)2 A
= A

Coulomb gauge
~
~
= A A

partial integration
~
~ (
~ A)
~
= (A)
{z
}
|
i Ak i Ak

leading to
1
E=
8

d x

3
X
i=1

1
c2

Ai
t

2

~ i
+ A

2

Substituting the Fourier ansatz gives


E=



2L3 1 X 2 2

N
(
)
~
a
~
a
+
~
a
~
a

~k
~k ~k
~k
~k ~k
8L3 c2
~k

(2.4)

2.1. THE FREE ELECTROMAGNETIC FIELD

31

Exercise: provide the intermediate steps. Now, using = c|~k| and orthogonality, we have
Z
~ ~0
d3 x ei(kk )~x = L3 ~k,~k0
V

where a Kronecker delta appears because the k-indices are discrete (remember that we had periodic boundary conditions).
2 /c2
~
The cross terms (~a~k ~a~k +~a~ ~a ~ ) in the sum come from both ( A/t)
k k
~ A),
~ but with opposite sign, so they cancel. Finally, choosing N =
and (
p
2
2c /k , we obtain
E=


1X 
~k ~a~k ~a~k + ~a~k ~a~k
2

with ~k ~a~k = 0

(2.5)

~k

which is a sum over harmonic oscillator terms.


The Poynting vector is calculated similarly:
1
P~ =
4c
~ B
~ = 1
which, with E
c

1
P~ =
c

~ B
~ =
d3 x E

~k~a ~a~
~k
k

~k

~
t A



~ A
~ , becomes:





~
~
~
~
~
A A
A A
t
t

The second term becomes zero after partial integration.


Note
In the continuum
where L , the sum is replaced
an inteR 3 limit,
R by
1
3
3
3

gral: d k/(2) L . With N


, it follows A d k/(2)3 (...),
L3
R 3
and H = 1/2 d k/(2)3 (...)
The expression (2.5) above can be interpreted intuitively as a collection of an infinite number of harmonic oscillators, with ~a~k and i~a~ as
k
generalized coordinates and momenta (compare to the oscillator chain
from chapter one). We will have to substantiate this heuristic interpretation when introducing the canonical formalism for fields in the
next section.
The corresponding position (and momentum) operators cannot navely
~ x, t).
be identified with the argument ~x of A(~

32CHAPTER 2. FREE EM FIELD, KLEIN-GORDON EQ, QUANTIZATION

Figure 2.1: Electromagnetic wave propagation

2.1.2

Quantization

The quantization is procedure is similar to the one for harmonic oscillator


i h i
a , ia
= ~k,~k0 , ,
~ ~k ~k0

, = 1, 2

(2.6)

for transversal polarization degrees of freedom : ~a~k = ~e 1 a~1 + ~e 2 a~2 with


k
k
~e 1 ~e 2 ~k.
Promoting the ~a~k and i~a~ from eq. (2.5) to operators, we have
k

H=


1 X 
~k a~k a~ + a~ a~k
k
k
2
~k,

leading to
H=


~k

a~
k

a~k

~k,

1
+ ~
2


(2.7)

Note that a~ a~ = ~n~ is the occupation number of the harmonic oscilk


k
k
lator n and the zero-point energy of the ~k-mode is ~ ~ /2.
~k

Again, the space of states is the direct product space of the Hilbertspaces corresponding to the modes, called Fock-space:

E
Y

H=
H~k, ; states: n~1 , n~2 , . . . n~N
k

~k,

E=

X
~k,



1
;
~~k n~k +
2

P~ =

~ ~k n~k

~k,

Here we have an infinite number of degrees of freedom (modes), each


with its own occupation number as for the harmonic oscillator. Because

2.1. THE FREE ELECTROMAGNETIC FIELD

33

Figure 2.2: Absorption and Emission process

of the sum there is an infinite ground state energy, the so-called vacuum
energy. This can be removed it by normal ordering, : aa : = a a (moving
all annihilation operators to the right of all creation operators), but this is
ad hoc. Infinite energy shifts are also a problem in general relativity, where
energy density influences the curvature of spacetime.

2.1.3

Application: emission, absorption

Consider induced emission or absorption, or spontaneous emission due to an


interaction (in QM):
e
~ x, t).
p~ A(~
(2.8)
HI =
mc
With a quantized electromagnetic field, we have

E

initial state |ii = |aiAtom . . . n~k

E

final state |f i = |biAtom . . . n~k 1
With


1 X  i~k~x
~
~ x,
A(
t) =
N a~k ~e~k e + a~ ~e~k eik~x
k
L3 ~
k,

we obtain

hf |HI |ii =

e
mc

2~c2
L3 k

1/2 D

q
n~ + 1
E
k
i~k~
x
b|~p ~e~k e
|a
q

n~
k

n~ = 0 means spontaneous emission ( natural line width, considering


k
perturbation theory).
Finally, we have
transition probability / time =

2
|hf |HI |ii|2 (Ef Ei ).
~

34CHAPTER 2. FREE EM FIELD, KLEIN-GORDON EQ, QUANTIZATION

2.1.4

Coherent states

To approximate an electromagnetic wave one introduces coherent states like


~ E,
~ and B
~ through exfor the single harmonic oscillator, and expresses A,
pectation values in these coherent states. First observe that with

X  2~k 1/2 

i~k~
x
i~k~
x
= 1 A
= i
E
~

a
(t)
e

a
(t)
e
~k
~k
~k
c t
L3
~k,

|ni = 0 for E
between states with fixed photon number n. Also note
hn| E
h

i
6= 0
N~k , E

(2.9)

i.e. photon number and electric field can not be measured simultaneously.
However, we also have
*
+


2
1
~
E
6= 0.
n| |n = 3 n +
4
L
2
Now consider a monochromatic coherent state

E
~
(t)a
,
k,

= Ne ~k, ~k |0i~k, .

(2.10)

and calculate
D



E

2~ 1/2 
i~k~
x

i~k~
x
~
~

, k, |E|, k, = i

(t)
e

~
e
(t)
e
~
~k
~k,
k,
L3

Observe
which is the Fourier-component of a monochromatic, polarized E.
that determines the amplitude. The conclusion is that for induced absorption/emission, one can get back the normal QM description
by using
Q
coherent states. (Note: the Fock-space vacuum |0i = ~k, |0i~k, allows
for product states of (2.10) in the more general case.)

2.2
2.2.1

Klein-Gordon equation
Lagrange formalism for field equations

Consider an action
Z

t1

S[L] =

Z
dt

t0

d x L (~x, t),
|
{z




~
,
t
}

L((,t),... )

and a field (~x, t). Discretizing the field, using a lattice, turns it into a
vector, with position serving as an index: (~xi , t) i (t). Note, however,

2.2. KLEIN-GORDON EQUATION

35

Figure 2.3: Field Lattice

that we retain infinitely many degrees of freedom even in the discretized


case.
The Lagrangian L is a functional of (, t), the action S of (, ). We
postulate S[L] = 0 with a variational derivative, i.e. S is extremal (changing at position ~x and time t).
Functional derivatives
For a functional F (()), we have
Z
F (() + ()) F (()) = 

dx

F
(x) + O(2 )
(x)

F
. An elegant, if a little heuristic,
which defines the functional derivative (x)
choice for the perturbation function is (x) = (x x
).

F ((x) + (x x
)) F ((x))
F
lim
=
0

(x) x=x

Remarks:
Discretization (x xi , (xi ) i ) turns the derivative into a gradient and (x x
) into ij .
R
If the functional is defined as F (()) = dx F ((x)), then
F (())
F ((x))
=
(x)
(x)
i.e. we also can get along without functional derivative notation.
(xa ) is a functional itself:
() (xa )

36CHAPTER 2. FREE EM FIELD, KLEIN-GORDON EQ, QUANTIZATION



(xa ) + (xa x
) (xa )
(xa )
= lim
= (xa x
)

(x) x=x 0

This will be important later on when calculating the Poisson bracket
of , .
Applying the above to the action, we get:
Z

t1

L
(t)
L
(t)(~x0 ~x) +
(~x0 ~x)

(/t)
t
t0

L

0
+
(t) 0 (~x ~x)
|i
xi


Z
Z
t
1
=
d
L
L
L
3 0
0

part.int.
dt (t)
d x (~x ~x)
dt (/t) x0i |i
t0
S =

dt

3 0

d x

where |i := x
. There are no boundary terms in this formulation with
i
-functions and (t0 ) = (t1 ) = 0. In the original, and more orthodox,
formulation, without deltas, + (~x, t), and for all , = 0 at the
endpoints t0 and t1 and |~x| .
Regardless of the formulation, one has

S = 0 (t), ~x.
resulting in
L

L
L

=0
t (/t) xi |i

(2.11)

for the field equation.


Remark: in case of a finite volume, one takes periodic boundary conditions in x-space.
The momentum field is defined as
(~x, t) =

L
L
=
(/t)
(/t)

Note: the (canonical) momentum field (~x, t) is not the momentum P of


the field.
From the Lagrangian density L and associated momentum field we
obtain the Hamiltonian by means of a Legendre transform.


Z

3
~
H = d x (~x, t) (~x, t) L(, , )
t
t
|
{z
}
Hamiltonian density H

2.2. KLEIN-GORDON EQUATION

37

The Hamilton equations read

H
H
=
=
t

H
=
=
t

H
H
|i

|i


(2.12)

The Poisson bracket is defined analogously to the discrete case:



Z 
B
A
B
A

d3 x = (~x~x0 )
{A(, ), B(, )}P oisson :=
(~x) (~x) (~x) (~x)
Exercise: show that this indeed gives a delta function. Note that , (~x)
are particular functionals, see remark page before.

2.2.2

Wave equation, Klein-Gordon equation

The wave equation



1 2
A(~x, t) = 0
c2 t2
{z
}
|

=

is solved by the Fourier ansatz


~

A = ei(tk~x)

with

2
+ k2 = 0 .
c2

With E = ~ and p = ~k, this translates into the photon energy relation:

E2
+ p2 = 0
c2

Similarly, for massive particles we have

E2
+ p2 + m2 c2 = 0
c2

or, dividing by ~

2 (~k)
m2 c2
+ k2 + 2 = 0
2
c
~
This translates to the differential equation

1 2
m2 c2

+
c2 t2
~


(~x, t) = 0

(2.13)

which is the promised Klein-Gordon equation for a scalar field. The equation
follows from the Lagrangian
" 
#

1 1 2  ~ 2 m2 c2 2
2 .
L=
2 c2 t
~

38CHAPTER 2. FREE EM FIELD, KLEIN-GORDON EQ, QUANTIZATION


The corresponding momentum field is
L
=
(~x, t) =

resulting in the following Hamiltonian density:


" 
#
2 
2 m2 c2
1 1
~
+
H=
+ 2 2
2 c2 t
~
which is positive definite. It can be simplified by using the Fourier decomposition of (~x, t) in a volume V = L3


1 X
~
~
(~x, t) =
N~k a~k (t)eik~x + a~k (t)eik~x .
L3 ~
|
{z
}
k
~k (t) real!
It solves the Klein-Gordon equation if


1 2
m2 c2
2
~
~k (t) = 0
+k + 2
c2 t2
~
~
i.e. if a~k (t) = ei(k)t a~k (0), where ~k 2 + (mc/~)2 2 (~k)/c2 = ~k 2 +
~
eik~x

m2 c2
~2

Since

do not matter
the differentiation is only with respect to time, factors of
in this equation. Plugging in this definition of in H gives
H=


1X ~ 
(k) a~k a~k + a~k a~k
2

(2.14)

~k

 2 1/2
c
. As we would expect, a~k and ia (~k) are
where we have set N~k = 2
~
k
canonically conjugate variables:
a~k
H
=
= i(~k) a~k
t
(iak )

(ia~k ) =

H
= ~k a~k
a~k

which can be checked using the form of a~k derived above.

2.2.3

Quantization

This is done most easily in momentum space. The commutator of the canonically conjugate variables is
i
h
i
ih
a~k0 , a~ = ~~k,~k0
(2.15)
ia~ 0 , a~k = ~k,~k0 leading to
k
k
~
We redefine this
a~
a~k k :
L3

i
a~k0 , a~ = ~L3 ~k,~k0
k

2.2. KLEIN-GORDON EQUATION

39

(again, see the oscillator chain). Since we have for L


Z
X
d3 k 3 3 (~k ~k 0 )(2)3
~k,~k0 = 1
L
=1
(2)3
L3
~k~
n

we see that in the continuous case, the commutator must be


[a~k0 , a~ ] = ~(2)3 3 (~k ~k 0 )
k

(2.16)

We could have derived this directly from the canonical formalism, promoting the Poisson bracket to a commutator:


(~x, t), (~x0 , t) P oisson = 3 (~x ~x0 )
~/i
quantization


0
(~x, t), (~x , t) = i~3 (~x ~x0 )
from which the above result for a~k follows. Exercise: invert the relation


between , and a, ia and calculate a, a . Alternatively, which is easier,
plug in , in terms of a, a and obtain the relation above.
Of course we have now
Z


1
d3 k

i~k~
x
i~k~
x
q
a
(t)
e
+
a
(t)
e
(2.17)
(~x, t) =
~k
~k
(2)3
~
2(k)
Z

3
d k 1 ~ 

H=
(
k)
a
(t)a
(t)
+
a
(t)a
(t)
(2.18)
~k
~k
~k
~k
(2)3 2
Remarks
Since we know the wave equation from electromagnetics, and since we used a
relativistic dispersion relation between and ~k to begin with, it is clear that
our equation is relativistically invariant, as well as the Lagrangian related
to it. We will learn more about this later on. The remarkable aspect of
this whole episode is that canonical quantization, where time plays a special
role, can be brought into consistency with relativity, where time and space
are on the same footing.
L=

2.2.4

1
m2 2

...
2
2

The Maxwell equations in Lagrange formalism

Here, we will shortly discuss gauge theories (again, more details will follow).
For now, we will restrict ourselves to the free case, i.e. without sources. We
will use units where x0 = t, c = 1. Consider the electromagnetic tensor:
F = A A

(2.19)

40CHAPTER 2. FREE EM FIELD, KLEIN-GORDON EQ, QUANTIZATION


~ and B
~ in relativistic notation.
and note that it contains E
The electric and magnetic fields appear in it as follows:
F 0i = E i ;

F ik = ikl B l
~ =
~ A;
~
E
t

A0 = ;

~ =
~ A
~
B

~ = i = i = i . So, two Maxwell equations are automatically


with
x
fulfilled. What remains to be shown is
~ E
~ =0

or, in other symbols,

~
~ B
~ = E
and
t
F
=0
x

With eq. (2.5), this becomes


2 A A = 0

(2.20)

In the Coulomb gauge, i Ai = 0 and = A0 = 0 (hence the term radiation


gauge). Hence, eq. (2.20) reduces to 2 A = 0, which is just the wave
2
~ 2 . The equation for A follows from L =
equation, as 2 = /x0
1/4F F if one varies the A and A independently (Exercise: work this
out.)
~ and B,
~ L = 1/2(E
~2 B
~ 2 ), but we do not vary with
Note: in terms of E
respect to these variables.
Calculating the canonical momenta (see Bjorken and Drell, vol. II) gives:
0 =

L
= 0;
(A0 /t)

k =

L
Ak
A0
=

= Ek
(Ak /t)
t
xk

Inspecting the canonical formalism, we get the following result, which looks
reasonable:

H = k Ak L =
t
Z
Z
1
3
H =
d xH=
2

1 ~2 ~2
~
~
(E + B ) + E
2
~2 + B
~ 2)
d3 x (E

~E
~ = 0.
The last term can be disposed of through partial integration, since
~
~
This is not without problems, however, since E = 0 is violated by both
the fact that 0 = 0, and the canonical commutator of Ak and k , which is
given by
h
i
?
i (~x, t), Ak (~x0 , t) = i(~x ~x0 ) ik
Indeed it turns out that in this case, we have to quantize with constraints in
a more rigorous approach (see the literature on the Dirac bracket, which is

2.2. KLEIN-GORDON EQUATION

41

a special line of research). Note that working with constraints already pops
up in classical field equations.
The simplest implementation of this idea is to do like is done in QED:
take the Coulomb gauge.

Z
Z 
1 A A 1 A A
3
+
d3 x

L =
d xL=
2 x x
2 x x

Last term: partial integration, use i Ai = 0, A0 = 0 = 0


Z
1
~ A
~ d3 x.
=
A
2
We get 3 times the scalar case. Varying this separately for the two transver~ long ) components of A
~ with the constraint that
sal and one longitudinal (A
~ A
~ = 0, leads us to the canonical picture for E
~ trans (
~ E
~ = 0) and for

~
~
~
Atrans ( A = 0), as was used in heuristic expose.

42CHAPTER 2. FREE EM FIELD, KLEIN-GORDON EQ, QUANTIZATION

Chapter 3

The Schr
odinger equation in
the language of QFT
3.1

Second Quantization

For radiative transitions, the electromagnetic field is described by an operator in Fock space, whereas the atom retains its old QM description, i.e.
~~
~
with an eikX , where the position is the operator X.
Now, the following question comes up: can we treat the Schrodinger
equation, written down according to the rules of the correspondence principle, in its x-space form, like an ordinary field equation and quantize it?
This sounds odd, since it is already quantized; why do it again? The point
is that we want QM to be a special case of QFT, as mentioned in chapter
1, and therefore, it should be possible to express it in the language of QFT.
This is called second quantization.
Note that we are not looking for new physics: QM has been tested
thoroughly, and looks far too nice to throw away.

3.1.1

Schr
odinger equation

In classical mechanics, we have E = p~2 /2m + V (~x); following the correspondence principle, this gives
"
!
#
~2
~
~2
i~ ((x), t) =
+ V (~x) (~x, t)
(3.1)
t
2m
{z
}
|
HSchr

This has stationary solutions n = n (~x)eiEn t/~ ; we will also use


En /~ = n . General solutions, obtained from a generalized Fourier transform, are
X
(~x, t) =
an (t)n (~x)
(3.2)
n

43


44CHAPTER 3. THE SCHRODINGER
EQUATION IN THE LANGUAGE OF QFT
where an (t) ein t and the {n } form a complete set of orthonormal
functions.
In the QFT quantization, the functions an (t) are promoted to operators
an (t); as usual, we are in the Heisenberg picture. The final result of this
second quantization will be:
Z

H = d3 xop
(~x, t)HSchr op (~x, t)
(3.3)
which is an operator in Fock space. In terms of an , we have
X
H=
En an an

(3.4)

where the an and ian are canonically conjugate operators in the Hamiltonian
formalism. Let us now consider in detail how we get this result.

3.1.2

Lagrange formalism for the Schr


odinger equation

In order to quantize the Schrodinger equation, we will treat it as an ordinary


field equation, for which we will first develop the Lagrange formalism. The
Lagrangian density is:
~
~
~2
L = i~
V (~x)
2m

(3.5)

Following the usual procedure, we obtain the momentum fields:


=

L
= i~ ;

= 0

(3.6)

We could obtain our field equation by varying L with respect to the real
and imaginary parts of , but we can also vary with respect to and ,
since these are complex conjugates. This gives:
~
~

i~ +
V (~x) = 0
2m

(3.7)

which is the Schr


odinger equation, as we had hoped. Now, our Hamiltonian
density (as used in eq. (3.3)) is given by
~
~

H = i~ L = ~2
+ V (~x)
2m

(3.8)

Here, and i~ are canonical conjugates. In this equation, we could also


introduce electromagnetic coupling.
Note that we could also have taken a more symmetric Lagrangian density,
i~
) to obtain this result.
(
2

3.1. SECOND QUANTIZATION

3.1.3

45

Canonical quantization

The canonical quantization relation, where the and are promoted to

op and op
, is given by

[op (~x, t), i~op


(~x0 , t)] = i~ 3 (~x ~x0 )

(3.9)

Note that the factor i~ appears on both sides of the equation, and hence
()
cancels. In terms of the an , we have
X
(3.10)
[an , an0 ]n (~x)n0 (~x0 ) = 3 (~x ~x0 )
n,n0

where the generation and annihilation operators fulfill their usual commutation rule
[an , an0 ] = n,n0
(3.11)
This gives the completeness relation: Rthe n (~x) form a complete basis.
The Hamiltonian operator H = d3 xH is, as discussed in chapter 1,
given by
X
H=
En an an
(3.12)
n

Using eq. (3.11), we find the Heisenberg equations for the an and an :
d
i
an (t) = [H, an ] = in an (t)
dt
~
d
i
an (t) =
[H, an ] = in an (t)
dt
~

(3.13)
(3.14)

For the op and op


, we have:

i~

d
op (~x, t) = [H, op (~x, t)]
dt
d

i~ op
(~x, t) = [H, op
(~x, t)]
dt

(3.15)
(3.16)

The energy states in Fock space are obtained as usually, by applying an to


the vacuum. The action of an and an on a state is given by
p
an |. . . , Nn , . . . i =
Nn |. . . , Nn 1, . . . i
p

an |. . . , Nn , . . . i =
Nn + 1 |. . . , Nn + 1, . . . i
()

Letting the op (~x, t) act on the vacuum gives:


op (~x, t) |0i = 0,

(3.17)


46CHAPTER 3. THE SCHRODINGER
EQUATION IN THE LANGUAGE OF QFT
since op contains only annihilators, and

~
op
(~x, t) |0i = |Xi
H

(3.18)

~ is fixed, but described


which is a state with sharp position. The position X
by time-varying Hilbert space vectors, which is where the time dependence
comes into the picture.
Exercise: prove eq. (3.18). Hint: use the Schrodinger equation for
R

~ H (t) = d3 x0 op
(~x, t) = hX|H i, or write X
(~x0 , t)~x0 op (~x, t) and show
~
~ H |Xi
~
that X
H = X |XiH .
For the particle number density, we have:

n(~x, t) = op
(~x, t)op (~x, t)
0

n(~x

, t)op
(~x0 , t) |0i

= (~x ~x

(3.19)

) op
(~x, t) |0i

{z

localized

(3.20)

The total particle number N should be conserved, since we are dealing with
normal QM systems, where particle production is forbidden. Indeed, using
eq. (3.19) and the Heisenberg equation, we find that
Z
d3 xn(~x, t) = N
is conserved.
Quite generally, for any Schrodinger operator AS (XS , PS ), we have an
associated (Heisenberg) operator AF (ock) acting on the Fock space of the
one-particle state:
Z

AF (t) = d3 xop
(~x, t)AS op (~x, t)
(3.21)

Using the commutator [op , op


] = 3 , the fact that double annihilation on a
one-particle state gives zero (op op |1 part.i = 0), and the definition above,
we see that commutation relations of the form

[AS , BS ] = CS
translate to
[AF , BF ] = CF
(the notation AH is also used
An example of the above
which generates translations.
~ F acts on states in the Fock
P

for AF ).
correspondence is the momentum operator,
~
It is defined by U~a = eiPF ~a/~ , where now
space. Applying it to a state |~xi gives


3.2. MULTIPARTICLE SCHRODINGER
EQUATION

47

(~x, t)U~a U~a |0i


U~a op
(~x, t) |0i = U~a op
| {z }
|
{z
}
=|0i

=|~
xi

Using the fact that |0i is invariant under translations, this becomes:

|~x + ~ai = op
(~x + ~a, t) |0i

For infinitesimal translations, we have


h

i
()
~ F , () = ~
~ op
P
op
i

For rotations, generated by the angular momentum operator ~JF , we have


~
the same procedure: U~ = eiJF ~/~ , where ~JF is given by


Z
~JF = d3 x (~x, t) X
~ S ~
~ op (~x, t)
op
i
And again, for infinitesimal rotations:

h
i 
()
~JF , () = X
~ S ~
~
op
op
i

3.2
3.2.1

Multiparticle Schr
odinger equation
Bosonic multiparticle space

The description of systems consisting of more than one identical particles


is a nice application of the new QFT-inspired formalism obtained from the
second quantization. Note again that we are still talking about QM as we
know it; only the language has changed.
Consider a system of N identical particles, which can be in states like

E
E E E




n~k1 n~k2 . . . n~k = n~k1 n~k2 n~k
l

where the possibility


of
polarization has been left out to avoid drowning
in indices. The states n~k are given by

(a~k )n~k
n~ = p
|0i
k
n~k !

(3.22)

General states (still without polarization) look like this:

X
X


E

f (n~k1 ~k1 , . . . , n~k ~kl ) n~k1 . . . n~k
l

l=1 {~k1 ...~kl }

(3.23)


48CHAPTER 3. THE SCHRODINGER
EQUATION IN THE LANGUAGE OF QFT
Here, the order of the ~ki is arbitrary, since the a~ commute. The total
ki
P
particle number is given by N = li=1 n~ki . These states live in a Fock space
obtained by a direct product of harmonic oscillator Hilbert spaces:
F=

H~osc
k
l

l=1

They obviously inherit the linear structure of their constituents. Similarly, the inner product between two such states is obtained by taking the
inner product in each H~k separately.
Another way of obtaining F is summing spaces with fixed particle numbers:
F = H(0) H(1) H(l) =

H(n)

where
H(i) =

H(1)

symmetrized

is the Hilbert space for i identical particles.

|~xi . . . ~xl iH = Ns op
(~x1 , t) . . . op
(~xl , t) |0i

(3.24)

The op
generate identical particles, and hence commute. They are defined
as follows:
P
~

N~k a~ (t)eik~x for free particles

k
~k

(3.25)
op
(~xj , t) =
P

an (t)n (~xj ) j = 1, . . . , l for bound particles

~k

Because of the fact that the op


commute, we obtain l! terms with the same
~
set of ki in the free particles-case. This gives rise to a factor of l! in the
inner product of a state with itself. This factor is always the same (exercise:
show this), even if some of the ~k are identical. Examples are the case where
all ~k are different and the case of m identical ~k: in the first case, the norm
squared of the state gets a factor l!, as just mentioned, and in the second,
it is a factor


l
m!(l m)! = l!
m

So, we set the normalization coefficient in eq. (3.24) Ns = 1/ l!.


Exercise: show, using the normalization coefficient derived above, that
the following identity holds for a function fsym that is symmetric in the xl :
Z


d3 x01 . . . d3 x0l ~x1 . . . ~xl |~x01 . . . ~x0l fsym (~x01 . . . ~x0l ) = fsym (~x1 . . . ~xl )


3.2. MULTIPARTICLE SCHRODINGER
EQUATION

3.2.2

49

Interactions

The total Hamiltonian can be split up into the parts concerning the individual particles, parts resulting from interactions between two particles, etc.:
H = H (1) + H (2) + . . .

(3.26)

The parts H (i) are defined as follows:


H

(1)

Z
=

H (2) =

d
Z

xop
(~x, t)


~2
+ V (~x op (~x, t)
2m

(3.27)

~0
d3 xd3 x0 op
(~x, t)op
(x , t)V (~x, x~0 )op (~x, t)op (x~0 , t)(3.28)

The interaction potential V (~x, ~x0 ) = V (~x, ~x0 ) is self-adjoint, e.g. e2 /|~x ~x0 |
for the interaction between two electrons.

Exercise: using [op (~x, t), op


(~x0 , t)] = 3 (~x ~x0 ), show that
i~

op (~x1 , t) . . . op (~xl , 1)

(~x1 , . . . , ~xl , t) = i~ h0|


|iH
{z
}
t |
t
l!
=H h~
x1 ...~
xl |iH

op (~x1 , t) . . . op (~xl , t)

= h0| [H (1) + H (2) ,


|i
l!

l
l
X
X
!
~ i) + 1
H (1) (~xi , ~
= h0|
V (~xi , ~xj )
S
i
2
i=1

i,j=1,i6=j

op (~x1 , t) . . . op (~x1 , t)

|i
l!
(1)

(2)

= (HS + HS )(~x1 , . . . , ~xl , t)


Given that we are dealing with two-body interactions here, instead of the
single particles discussed so far, one could think that the commutation rela
tion between op and op
might not be the same as before. However, since
the canonical momentum remains unchanged, we know that the old relation
must still hold.
We started with the linear Schrodinger equation, but we could also have
started with a nonlinear equation, such as the Gross-Pitaevsky equation
for Bose-gasses. In this case, we would have to give up the superposition
principle for probability amplitudes. Since we are doing nonrelativistic QM
here, we prefer taking into account nonlinear parts only in H (2) and higher
orders. Of course, one can also apply ordinary perturbative QFT, as will be
discussed later on.


50CHAPTER 3. THE SCHRODINGER
EQUATION IN THE LANGUAGE OF QFT
Remark
With the results obtained above, we can write any interaction completely in
~
quantized fields. The electromagnetic interaction, with A(x),
is, as we will
see later on, obtained by minimal gauge coupling:
~~
~~
e~

A
i
i
c

Z
H0 =

xop
(~x, t)

HI = H

(2)

1
2m

Z
+

~~

2

xop
(~x, t)

1
+ V (~x) op (~x, t) +
8

~2 + B
~ 2)
d3 x(E



e~ ~ ~
e2 ~ 2
A op (~x, t)

A+
imc
2mc2

~ are now quantum fields.


where , and A

3.2.3

Fermions

We know from QM that we need the Pauli exclusion principle for electrons,
protons, and other particles with half-integer spin. Later on, this will appear
as a necessary consequence of QFT; we will see this in our discussion of the
Dirac equation. Here, only the results will be given: fermionic fields are
quantized by substituting anti-commutators, rather than commutators, for
the Poisson bracket. We postulate:
[cn , cn0 ]+ = cn cn0 + cn0 cn = nn0

(3.29)

[cn , cn0 ]+ = [cn , cn0 ]+ = 0

(3.30)

(the notation {, } is often used as an alternative for [, ]+ .)


Let us investigate the consequences of postulating anticommutating creation and annihilation operators. A striking result is that is immediately
obvious is that c2n = c2
n = 0. Now, consider the action of the occupation
number operator N = c c, N |ni = n |ni. Let it act on a state c |ni:
Nc |ni = c |{z}
cc |ni = 0 =
=0

= (cc + 1)c |ni = (cN + c) |ni = (n + 1)c |ni = 0


So, either n = 1 or c |ni = 0. Now, let it act on a state c :
Nc |ni = c cc |ni = c (c c + 1) |ni = c |ni = (n + 1)c |ni
| {z }
=0


3.2. MULTIPARTICLE SCHRODINGER
EQUATION

51

So, either n = 0 or c |ni = 0, giving the following relations:


c |1i = 0,

c |0i = 0

c |0i = |1i ,

c |1i = |0i

(3.31)
(3.32)

In other words, the only occupation numbers are 0 and 1, as we wanted for
fermions. The normalization is as follows:
h0|0i = 1

(3.33)

h1|1i = h0| c c |0i = h0| cc + 1 |0i = h0|0i = 1

(3.34)

A general fermion state is:


l
X Y

f (~k1 . . . ~kl )c~ |0i


ki

{~k1 ...~kl } i=1

(3.35)

Q
with n~ki = 1. Since the c~ anticommute,
c~ is totally antisymmetric,
ki
ki
which means that only the totally antisymmetric part of f is interesting.
In x-space, states look like this:
1

(~x1 , t) . . . op
(~x1 , t) |0i
|~x1 (t) . . . ~xl (t)i = op
l!

(3.36)

where the op
are, as usual, general solutions of the field equations:

op
(~x, t) =

cn n (~x)

(3.37)

A state with l different oscillators occupied is given by:


X
1

(1)P kp (~x1 ) . . . kp (~xl )ck1 . . . ckl |0i =


1
l
l! {1,...,l}
P


k (~x1 ) . . .
1
1 k1 (~x2 ) . . .

..
l!
.
(~xl ) . . .
k1

kl (~x1 )
kl (~x2 )
..
.

ck1 . . . ckl |0i




kl (~xl )

(3.38)

The determinant is called Slater-determinant, and gives rise to antisymmetric wave functions. It arises from
which are summed over
P the permutations
P
with changing sign (the part {1,...,l}P (1) in the left hand side).


52CHAPTER 3. THE SCHRODINGER
EQUATION IN THE LANGUAGE OF QFT

Chapter 4

Quantizing covariant field


equations
So far, our discussion of linear field equations has been quite generic, with
the exception of the electromagnetic wave equation. The importance of this
special case is indicated by the fact that it led Einstein to special relativity,
and was the first triumph of QFT, which was absolutely essential for the
acceptance of the theory. Therefore, it seems to make sense for us to inspect the relativistic aspects of the QFT quantization. Later on, a thorough
classification of Lorentz covariant fields will be given; here we start with the
simplest case, the Klein-Gordon equation. (Note that the wave equation is
a special case of the Klein-Gordon equation: the one where m = 0.)

4.1

Lorentz transformations

We define contravariant and covariant spacetime vectors


x = (x0 , ~x)

x = g x

contravariant
covariant

(4.1)
(4.2)

with the metric tensor (in Cartesian coordinates)

g =

1
1

(4.3)

1
and x0 = ct. Hereafter, we will use units where c = 1, so x0 = t. Now, we
set
x2 = x x = (x0 )2 ~x2 = x g x
53

54

CHAPTER 4. QUANTIZING COVARIANT FIELD EQUATIONS

This is invariant under Lorentz transformations:


x0 = x

(4.4)

for which the shorthand x0 = x is frequently used. This invariance means


that
(4.5)
x2 = x02
which requires
x g x = x g x
g = T g

(4.6)

where we have relabelled and


on the left hand side. In
shorthand, we write
g = T g

(4.7)

The gradient, /x , is a covariant vector if x is a contravariant vector:

=
= (1 ) = (1 )T
0
0

x
x x
x
x
where we have used that x0 = x, x = 1 x0 , and hence x /x0 =
So /x = transforms with (1 )T , i.e. like a covariant vector:

(1 ) .

x0 = g x0

= g x
= g g x
|{z}
=g

which, using gg = (T )1 = (1 )T , gives


x0 = (1 )T x

4.2
4.2.1

Klein-Gordon equation for spin 0 bosons


Klein-Gordon equation

Let us begin by studying the simplest case of the Lorentz group: a scalar
field, denoted by (x). It can be transformed by a Lorentz transformation
or a translation, giving the following general transformation:
(x) 0 (x) = (1 (x a))
This is the active view of symmetry transformations. Above, we have used
that
0 (x0 ) = (x) with x0 = x + a
(4.8)

4.2. KLEIN-GORDON EQUATION FOR SPIN 0 BOSONS

55

Figure 4.1: Transformation of (x)


We already know the Klein-Gordon equation:
( + m2 )(x) = 0

(4.9)

(here written in units where ~ = c = 1). In components, it looks like this:


"
#



2
2

+ m2 (x0 , ~x) = 0
x0
xi
(note that due to the squaring of the second term, the Einstein summing
convention applies to the index i). It is form invariant under Lorentz transformations, since both m2 and transform like scalars, which are invariant under Lorentz transformations:
(0 0 + m2 )0 (x0 ) = 0

(rename x x0 )

( + m2 )0 (x) = 0
Hence 0 (x) is a solution of eq. (4.9) if (x) is one.

4.2.2

Solving the Klein-Gordon equation in 4-momentum


space

The general solution is obtained from a Fourier transform:


Z
d4 k

(x) =
(k)eik x
(2)4
Plugging this into eq. (4.9) gives

(k 2 + m2 )(k)
=0
so

(4.10)

56

CHAPTER 4. QUANTIZING COVARIANT FIELD EQUATIONS

(k)
= a(k)2(k 2 m2 )(k0 ) + a(k)2(k 2 m2 )(k0 )
which we insert into eq. (4.10) after substituting k k in the second term
(a(k) b (k)):
Z
(x) =

n
o
d4 k
2
2
ikx

ikx
(k

m
)(k
)
a(k)e
+
b
(k)e
0
(2)3

(4.11)

Here, b (k)eikx represents the negative energy solutions.


For real fields, this means b (k) = a (k), i.e. a(k) = a (k); see chapter
2. We can rewrite a complex field (x) in terms of two real fields 1 and
2 :

= (1 + i2 )/ 2

= (1 i2 )/ 2
yielding for a(k) and b(k):

a(k) = (a1 (k) + ia2 (k))/ 2

b(k) = (a1 (k) ia2 (k))/ 2

(4.12)
(4.13)

Lorentz invariants
(k 2 m2 ) is Lorentz invariant: it makes sure the solution lies on the mass
shell.R (k0 ) is Lorentz invariant on this mass shell (exercise: show this),
4
and
invariant as well, since k 0 = k, and | det | = 1, so
R 4 d Rk is4 Lorentz
0
d k = d k . This implies that
Z
Z
Z
d4 k
d3 k
2
2
dk0 (k02 ~k 2 m2 )(k0 )
(k m )(k0 ) =
(2)3
(2)3
Z

d3 k
1
=
k0
where
k
=
+
0,+
(2)3 2k0,+
is Lorentz invariant, and, finally, that (x) itself is a Lorentz scalar. To see
this, rewrite eq. (4.11) as
Z
o
1 n
d3 k
ikx

ikx
(x) =
a(k)e
+
b
(k)e
(2)3 2k0,+
and observe that the part between braces consists only of scalars. Later on,
we will sometimes write a(~k) instead of a(k), since k lies on the mass shell,
and hence has only the degrees of freedom of the three-momentum.
The normalization of the a(k) obtained above is somewhat different from
the one used in chapter 2, since it has been adapted to the relativistic character of the equation. The as are not exactly the ones from the harmonic
oscillator case.

4.2. KLEIN-GORDON EQUATION FOR SPIN 0 BOSONS

57

Reversing the equation, we can obtain out of it:


Z

~
a(k) = i d3 xeikx 0 (x)


Z
eikx
3
ikx (x)
:= i d x
(x) + e
x0
x0
Z

b (~k) = i d3 xeikx 0 (x)


Note that the inner product
Z
(1 , 2 ) = i

1 (x) 0 2 (x)d3 x

is only positive definite for positive energy solutions, as the differentiation


with respect to x0 brings down a factor of k0,+ .
Exercise: show that (k0 ) is Lorentz invariant. We also have k00 =
+ 0i ki . If k lies on the mass shell and k0 > 0, we have the following
inequalities:

a0 k0

~
P i|k|2 < k0 0 2
(0 ) < (0 )
i

T g

writing
= g in orthogonal coordinates (i.e. using the Minkowski
0
metric). If 0 > 1, we are dealing with the orthochronous Lorentz group,
which we will encounter later on.

4.2.3

Quantization: canonical formalism

The Lagrange density corresponding to a real field (x) with eq. (4.9) as
equation of motion is:
L=


1
(x) (x) m2 2 (x)
2

(4.14)

For a complex field = (1 + i2 )/ 2, we have


Lcom =
=

2

1 X
j (x) j (x) m2 2j (x)
2

j=1
(x) (x)

m2 (x)(x)

Note that here, we use x = (x0 , ~x) and = x 0 x 0 x


i xi .
Now, in the real case, we vary the Lagrangians with respect to . In the
complex case, we can choose between 1,2 and , , of which the latter

58

CHAPTER 4. QUANTIZING COVARIANT FIELD EQUATIONS

is the more practical one. This gives us the Euler-Lagrange equations of


motion:
!
L
L

+
=0

(4.15)

( x )
Note that the derivative x transforms like a covariant vector, whereas
L
is contravariant. This makes the first term a scalar, as it should be,
( )
x

since the second one is a scalar as well.


In the complex case, we obtain eq. (4.15) with instead of from
varying with respect to , and with from varying with respect to .
To obtain the Hamiltonian density, we need the momentum fields:
(x) =

=
0

x
( x0 )

This gives the Hamiltonian density


"
#



2

1
2
2 2
H(x) = 0 L =
+
+m
x
2
x0
xi

(4.16)

(4.17)

In the complex case, we have


(x) =

x0

(x) =

x0

and
Hcom (x) =

L
=

+ m2 (4.18)

x0
x0
xi xi

Both the complex and the real Hamiltonian densities are positive definite.
From these, we obtain the Hamiltonian
Z
H = d3 xH(, )
(4.19)
Later on, we will see that this is the zero component of a four-vector.
The Poisson bracket of these fields is the usual delta function:



(~x, t), (~x0 , t) P = 3 (~x ~x0 )

(4.20)

Quantizing this according to the usual rule gives the commutator




(~x, t), (~x0 , t) = i 3 (~x ~x0 )

(4.21)

4.2. KLEIN-GORDON EQUATION FOR SPIN 0 BOSONS

59

where ~ = 1. This does not look Lorentz covariant, since time and space
are treated differently here. Later on, however, we will see that it is Lorentz
covariant. In the complex case, we have the same commutator as above, or
its Hermitian conjugate:
i
h
(4.22)
(~x, t), (~x0 , t) = i 3 (~x ~x0 )
This implies:





(~x, t) 0
(~x, t)
0
3
0
, (~x , t) = i (~x ~x ) =
, (~x , t)
t
t

Rewriting the field (x) as a Fourier transform, we get


Z
o
d3 k
1 n ~ ikx
~ ikx
a(
k)e
+
a
(
k)e
(x) =
(2)3 2k0,+
or, for the complex case,
Z
com (x) =

o
d3 k
1 n ~ ikx
~ ikx
a(
k)e
+
b
(
k)e
(2)3 2k0,+

(4.23)

(4.24)

Inserting this in the definition of H gives the following Hamiltonian for the
real field:
Z
Z
o
d3 k 1 n ~ ~
3
~k)a (~k)
a
(
k)a(
k)
+
a(
(4.25)
H = d xH =
(2)3 4
Exercise: provide the intermediate steps. In the complex case, we have
Z
o
d3 k 1 n ~ ~
~k)a (~k) +
Hcom =
a
(
k)a(
k)
+
a(
(2)3 4
Z
o
d3 k 1 n ~ ~
~k)b (~k)
b
(
k)b(
k)
+
b(
(4.26)
(2)3 4
which features a second type of oscillators, the b-particles.
folNote that this integration measure is really the Lorentz covariant dk,
lowed by a factor of k0,+ /2, which we know from the traditional Hamiltonian
(it was called /2 before):
Z
Z
d3 k
1 k0,+
d3 k 1
H=
(.
.
.
)
=
(. . . )
(2)3 2k0,+ 2
(2)3 4
When quantizing this field, we promoted the a and a to operators a and
and similarly for the b in the complex case. The a, a have the following
commutation relation:

a ,

[a(~k), a (~k 0 )] = 2k0,+ (2)3 3 (~k ~k 0 )

(4.27)

60

CHAPTER 4. QUANTIZING COVARIANT FIELD EQUATIONS

The same holds for the bs. Also, any a commutes with any b, i.e. they are
completely independent and on the same level.
Note: in the literature, the factor of 2k0,+ is sometimes omitted; the
notation employed here, which does include it, along with some -factors,
is called the Cambridge notation.
Both the a, a and the b, b build up a symmetric Fock space. Mixed a,
b combinations therefore act in a product space. A state with n a-particles
and n0 b-particles looks like this:
0
n+n
Y

i=1

d3 ki c(k1 . . . kn ) c(kn+1 . . . kn0 )

(2)3 2ki,0
n!
n0 !

a (~k1 ) . . . a (~kn )b (~kn+1 ) . . . b (~kn+n0 ) |0i

(4.28)

Since these states have to be normalized, the normalization functions c and


c
have to obey:
Z
d3 kn+n0
d3 k1
.
.
.
|
c|2 |c|2 = 1
(4.29)
h|i = 1 =
(2)3 2k1,0
(2)3 2kn+n0 ,0

4.2.4

Charge conjugation

Consider the Hermitian conjugate of a complex quantized field (x):


Z n
o
(x) =
a (~k)eikx + b(~k)eikx
The roles of a and b have clearly been exchanged. Hence, we can conclude
that the b-particles are antiparticles of the a-particles. This is illustrated
more clearly by the coupling of a complex (-meson) field (x) to an
electromagnetic field:
+ ieA = D
Coupling a charged field to a vector potential, one replaces the momentum p i by the kinetic momentum in the Hamiltonian p eA ; this is
called minimal coupling and leads to the modified Klein-Gordon equation:
(D D + m2 )(x) = 0

(4.30)

which is gauge invariant under the following transformations:


0 (x) = eie(x) (x)
A (x) A0 (x) = A +
since the covariant derivative transforms as
D0 0 (x) = ( + ieA + ie )eie(x) (x)
= eie(x) D (x)

(4.31)

4.3. MICROCAUSALITY

61

( acting on eie(x) gives a term ie in the factor between brackets).


fulfills the same equation with e e , which is the same as replacing D with D . Since changing into means exchanging a and b, this
exchange really achieves charge conjugation. Now, we would like to have
a charge conjugation operator, in order to make this operation mathematically more fitting to the model. This operator should be unitary: reversing
a charge conjugation should give the original situation. Its action should of
course be:
(x) = UC (x)U1
C

(U1
C = UC )

(4.32)

(this is equivalent to b(~k) = UC a(~k)UC ). Exercise: show that this operator


is given by

Z
o
1n

a a+b ba bb a
UC = exp i dk
2
Hint: use

2 =
=
i 2

dk

a b i~k~x a b i~k~x
e
e
i 2
i 2

Then, rewrite the integrand in the formula for UC as a2 (~k)a2 (~k), where
.
a2 := ab
i 2

4.3

Microcausality

In its quantized form, the Klein-Gordon equation still should fulfill the requirements of special relativity. In evaluating whether it does, we will use
complex fields; we could also have used real ones.
The commutator of a field (x) at position x with its Hermitian conjugate at position y is given by
[(x), (y)] =: i(x y)
Z

d3 k
d3 k 0  ~ ~ 0 ikx+ik0 y
~k), b (~k 0 )]eikxik0 y
[a(
k),
a
(
k
)]e
+
[b(
=
2k0 (2)3 2k0 (2)3
Z

d3 k  ik(xy)
ik(xy)
=
e

e
2k0 (2)3
(4.33)
Renaming ~k ~k in the second term, the integrand can be rewritten as
~
2ieik(~x~y) sin(k0 (x0 y 0 )). This implies that if x0 = y 0 , = 0. Now, if
the difference between x and y is spacelike, one can always transform to a
frame of reference where (x y)0 = 0. Since the commutator is invariant
under transformations, this means that for a spacelike pair of points, is
always zero. Thus, microcausality is conserved. Commutators of or
alone vanish identically.

62

CHAPTER 4. QUANTIZING COVARIANT FIELD EQUATIONS

4.4

Nonrelativistic limit

We have the Klein-Gordon equation:


(02 ~ 2 + m2 )(~x, t) = 0

(4.34)

Extracting the time derivative and expanding around m gives:


i0 (~x, t) = (m2 ~ 2 )1/2 (~x, t)


1 ~2
+ . . . (~x, t)
= m
2m
Retaining only the first two terms, and choosing the positive sign convention,
we obtain the non-relativistic Schrodinger equation with a constant potential
(resulting from the first term, m). Note that this is only valid in the nonrelativistic limit, where |~k| << m: the full operator (m2 ~ 2 )1/2 is nonlocal.
The inner product of a field with itself is given by:

d3 k  ~ ~
~
~k)
a(
k)a
(
k)

b
(
k)b(
(2)3 2k0
(4.35)
d3 k
which is not positive definite. We can interpret |a(~k)|2 (2)
as
a
proba3 2k
0
bility density if we consider only solutions with positive energy, i.e. where
b 0. Negative energy solutions would have disastrous results in interaction
theory.
In QM, the normalization factor of 1/(2)3 2k0,+ is absent, and we just
~k)|2 d3 k. Fourier transforming this gives the wave function (~x, t),
have |(
which has a positive definite inner product for positive energy solutions.
Taking a localised ~x0 (~x, 0) = 3 (~x ~x0 ), Fourier transforming back to
5/4
H5/4 (imr)
momentum space ( a(~k)) we obtain ~x0 (~x, 0), ~x0 (~x, 0) m
r
with r = |~x ~x0 |. Although this drops exponentially for r >> m1 , it is
not really localized.
Z

(, ) =

d x (x) 0 (x) =
3

Figure 4.2: Unsharp position

4.4. NONRELATIVISTIC LIMIT

63

The position operator


~ becomes ~x0 :
In position space, the position operator X
~ ~x (~x, 0) = ~x0 ~x (~x, 0)
X
0
0
In momentum space, we get something different from QM:
p
~ i
~ p~ i~
X
2p20
The extra term of i~
p/2p20 is responsible for the zitterbewegung, which obscures the exact position.

64

CHAPTER 4. QUANTIZING COVARIANT FIELD EQUATIONS

Chapter 5

Interacting fields, S-matrix,


LSZ reduction
5.1

Non-linear field equations

Last chapter, we studied the Klein-Gordon equation, ( 2 + m2 )(x) = 0.


Now, let us consider non-linear equations, such as:
( 2 + m2 )(x) =

3!

(5.1)

where we have a (local) self interaction of the field, which is obtained from a
4
term Lint in L:
4! . Sometimes it is convenient to separate the nonlinear
effects, introducing a source field j(x) in the above equation::
( 2 + m2 )(x) = j(x)

(5.2)

and obtain solutions to eq. (5.1) in this way.


In eq. (5.1), the canonical momentum field is unchanged, since Lint
does not contain any derivatives, and therefore the canonical commutation
relations are still valid. However, the Fourier decomposition used before is
not available here, because the equation is, as said, non-linear.
Remark
It is possible to construct the Noether currents and charges (the generators
of transformations) and check whether Lorentz covariance is consistent with
the canonical commutation relations. This turns out to be the case, as we
will see later on.
In this course, we will mostly treat Lint as a small perturbation of the
free theory, which brings us to perturbation theory. First, however, we
will present a more general framework: the LSZ formalism, named after its
inventors Lehmann, Szymanzik and Zimmermann.
65

66

5.1.1

CHAPTER 5. INTERACTING FIELDS, S-MATRIX, LSZ

Relation to observation

Elementary particle physics


In elementary particle physics, we study the scattering of particles. As we
will see, this problem can be reduced to calculating vacuum expectation
values (v.e.v.s) in the ground states of the interacting theory:
h0| T ((x1 , t1 ) . . . (xn , tn )) |0i
Statistical mechanics
Although QFT is most directly connected to elementary particle physics,
there also are links to other areas, such as statistical mechanics. There,
one works with phase space quantities, whose dynamics are described by a
Hamiltonian. In the QM version, one has a Hilbert space, in which states,
including multiparticle states, live; for quantum gases we have a Fock space.
In both the Hilbert and the Fock space, observables correspond to operators,
as is known from QM. Quantities that are averaged over space, such as
magnetisation M (~x, t), correspond to field operators (~x, t). Then, it is
possible to investigate correlations in some ensemble
h(x1 , t1 ) . . . (xn , tn )i
which, in the canonical ensemble, are weighed by a factor of eH/kB T for a
temperature T . We will come back to this case later, in the context of path
integrals.

5.2

Interaction of particles in QFT

In the QFT description of interacting particles, the in and out fields


are free fields approached asymptotically by the interacting fields. The in
and out Fock spaces with state vectors |ini and |outi, and the quantum
fields in (x) and out (x), are unitarily equivalent: physically equal states
|ini, |outi are described by different state vectors. In the Heisenberg picture
(as usual in QFT), we have
out (x) = S 1 in (x)S
|outi = S 1 |ini = S |ini

and

(5.3)
(5.4)

5.2. INTERACTION OF PARTICLES IN QFT

67

with a unitary matrix S. Note that states containing less than two particles
do not interact, as there is nothing to interact with:
|0iin = |0iout

and

|1iin = |1iout

Now consider the limit where (x) approaches in (x). The equation
( 2 + m2 )(x) = j(x)
is solved by
Z
(x) = hom (x) +

d4 y Gret (x y, m)j(y)

(5.5)

where Gret is the retarded Greens function, for which we have


( 2 + m2 )Gret (x y, m) = 4 (x y)
and Gret (x y, m) = 0

for x0 < y0

since the current j(x) influences only the future. From this equation, one
might navely assume that (x) approaches the homogeneous solution as
x0 , as mentioned at the beginning of this section. However, before
being able to assume this, one has to consider the covergence type. The
problem is that the normalizations of (x) and hom (x) are not generally
the same:
6 h1~
p| (x) |0i
h1~
p| in (x) |0i =
in (x) = eiPx in (0)eiPx
(x) = eiPx (0)eiPx
Both matrix elements are proportional to eiPx , but the norms are not equal,
since (x) |0i is in general not a pure one-particle state. States in Hilbert or
Fock space are normalized by their inner product with themselves, whereas
operators are normalized through commutators. So, what we have to consider is the convergence of (x) towards in (x) multiplied by some scalar:
lim (x) = Z 1/2 in (x)
{z
}
|
hom. solution

x0

5.2.1

(5.6)

The Yang-Feldman equations

This idea has been implemented in the Yang-Feldman equations:


Z

Zin (x) = (x) d4 xGret (x y, m)j(y)

Z
Zout (x) = (x)

d4 xGav (x y, m)j(y)

(5.7)
(5.8)

68

CHAPTER 5. INTERACTING FIELDS, S-MATRIX, LSZ

These can be used as definitions for in,out (x), since Z is fixed by the
canonical commutation relation for in,out . It also is the same in both equations. To see this, consider h1-part.| (x) |0i, and note that in this case, the
second term does not contribute.
We do not have convergence in the operator sense (or strong convergence)
for (x): it is not true that (x) Z 1/2 in (x), since this is not compatible
with the canonical commutation relations. Assuming (x) Z 1/2 in (x),
the equal time commutator would give:
[0 (x), (y)]x0 =y0 = i 3 (~x ~y )
3

Z[0 in (x), (y)]x0 =y0 = i (~x ~y )

so
for x0

which cannot be true, since Z is not in general equal to 1. Therefore, we


only have convergence in the following sense:
Only individual matrix elements converge. Since not all matrix elements necessarily converge at the same rate, the whole matrix, or the
operator, does not converge.
Wave packets built out of matrix elements, such as
Z

(f, t) = i d3 xf (x) 0 (~x, t)


with ( 2 + m2 )f = 0, also converge.
This type of convergence is also called weak convergence.
Remark
j(~x, t) 0 for t in the operator sense is not possible, unless one goes
to the Bogoliubov-Shirkov construction, where couplings are made to go to
zero as t (see their textbook on QFT).

5.3

Greens functions

We have already seen the causal Greens function,


DC (x0 x) = i(x0 x) = [f ree (x0 ), f ree (x)]
= h0| [f ree (x0 ), f ree (x)] |0i
which is given by
0

DC (x x) =


d3 k  ik(x0 x)
ik(x0 x)
e

e
2k0 (2)3

(5.9)

5.3. GREENS FUNCTIONS

69

and is zero for (x0 x)2 < 0, i.e. for spacelike displacements. Exercise:
show that DC can be written as follows with the help of the zeroth-order
Bessel function J0 :

i
J0 (m((x0 )2 r2 )1/2 )
(x0 )2 > r2
0
DC (x) = DC (x , ~r) =
0
(x0 )2 < r2
4r r
The canonical commutation relation ([t (x0 ), (x0 )]ET = i 3 (~x0 ~x))
gives

DC (x0 x) = i 3 (~x0 ~x)


(5.10)
x00
Now, we can see that the retarded and advanced Greens functions from
electrodynamics are given by:
DR (x0 x) = (x00 x0 )DC (x0 x)

(5.11)

DA (x0 x) = (x0 x00 )DC (x0 x)

(5.12)

Exercise: check that (x20 +m2 )DR,A (x0 x) = i 4 (x0 x). Use

(x00
x00

x0 ) = (x00 x0 ) and the fact that 0 is defined by partial integration.


The usual way to obtain DR,A is solving the differential equation via a
Fourier transform and realizing the boundary conditions by a choice at the
poles of the integrand:

Figure 5.1: Integration path for DR


d4 p
i
eipx
4
2
(2) p m2
Z
Z
d3 p
dp0
1
0
=
eip0 x ei~p~x
2
3
2
2
(2)
2i p0 p~ m
Z

DR (x) =

(5.13)

For x0 > 0, we close the integration loop in the lower half plane, and for
x0 < 0 in the upper half plane. For DA , we pass by the poles in the lower
half plane and close the loop in the upper half plane. The denominator in
the integrand comes from the two poles:
p0 +

(~
p2

1
1
1

2
2
1/2
1/2
2
2
2
p0 p~ m2 + i
+ m ) + i p0 (~
p + m ) + i

70

CHAPTER 5. INTERACTING FIELDS, S-MATRIX, LSZ

Figure 5.2: Integration path for DA


Later on, the Feynman propagator will be important; it is obtained by
the following integration path, and is given by
Z
DF (x) =

d4 p
i
eipx
(2)4 p2 m2 + i

(5.14)

Figure 5.3: Integration path for DF


For x0 > 0, the integral loop is closed in the lower half plane, so this
pole contributes

Z
1 ipx
d3 p
e

(2)3 2p0,+
p0 =p0,+
which is the first half of the commutator from DC ; for x0 < 0, the loop is
closed in the upper half plane, which gives the second half:
Z


d3 p
1 ipx
e

(2)3 2p0,+
p0 =p0,+

Altogether, this gives


DF (x x0 ) = (x00 x0 ) h0| (x0 )(x) |0i + (x0 x00 ) h0| (x)(x0 ) |0i
or
DF (x x0 ) = h0| T ((x0 )(x)) |0i

(5.15)

5.4. SPECTRAL REPRESENTATION OF COMMUTATOR VEV

5.4

71

Spectral representation of commutator vev

Consider the vacuum expectation value of the commutator of a field at


two positions x and y, h0| [(x), (y)] |0i. With a complete set of states |i,
it can be rewritten:
X
h0| (0) |i eip (xy) h| (0) |0i
h0| [(x), (y)] |0i =


h0| (0) |i eip (yx) h| (0) |0i
(remember (x) = eiP x (0)eiP x and P |0i = 0). Defining
X
(k) = (2)4
4 (k p )| h0| (x) |i |2 = (k 2 )(k0 )

(5.16)

and using 1 =

d4 k 4 (k p ), we can rewrite this as


Z

h0| [(x), (y)] |0i =


Z

d4 k
(k)(eik(xy) eik(xy) )
(2)4

dm2 (m2 )DC (x y, m2 )

(5.17)

which is called the spectral representation. Taking out physical 1-particle


contributions with the weak asymptotic condition allows us to rewrite this
as
Z
dm02 (m02 )DC (x y, m02 )

ZDC (x y, m2 ) +

>m2

Now applying the operator i0 to both sides and setting x0 = y 0 , we get


a factor of 3 (~x ~y ) on both sides. Setting the residues equal to each other
gives the K
allen-Lehmann sum rule:
Z

1=Z+

dm02 (m02 )

(5.18)

>m2

From this we can conclude that 0 Z 1. This will not be obvious in


perturbation theory, where Z 1 = 1 + g 2 . . . is divergent.

5.5

LSZ reduction formalism

The LSZ reduction formalism will be particularly useful for abstract Smatrix theory; in the Lagrangian formulation and in perturbation theory
one finally returns to the interaction picture.
The idea is to take a one-particle state out of the incoming state, calculate its contribution, and thus reduce the problem at hand to a n1-particle

72

CHAPTER 5. INTERACTING FIELDS, S-MATRIX, LSZ

Figure 5.4: Disconnected state running through


state.
S,p = hout |(p)in i = hout | ain (p) |in i
= hout | aout (p) |in i
Z

i hout | d3 xeipx 0 (in (x) out (x)) |in i

(I)
(II)

where in the second term, we have inverted the formula for to rewrite
ain aout in terms of . In term (I), hp| is being annihilated from the
outgoing state, so it only contributes in case of a state |pi running through
but not interacting. Since we are interested in the interaction, we will neglect
this term in the end.
For term (II), we now use the (weak) convergence of in/out (x0 , ~x) to

1/ Z(x0 , ~x) for x0 ; plus for outgoing and minus for incoming.
This allows us to rewrite it as
Z

i
(II) = lim
d3 xeipx 0 hout | (x0 , ~x) |in i
0
x
Z Z

i
d3 xeipx 0 hout | (x0 , ~x) |in i
lim
0
x
Z
Z
Z

i
0
=
dx
d3 xeipx 0 hout | (x0 , ~x) |in i
0
x
Z
With (02 ~ 2 + m2 )eipx = 0, the mixed term resulting from 02 cancels,
so we get
Z
i
(II) =
d4 xeipx ( 2 + m2 ) hout | (x) |in i
Z
where the Klein-Gordon operator acting on (x) gives j(x).
For outging particles (hout | = h(p0 )out |), the story is more or less the
same:

0

(p )out = hout | (x) ( \ p0 )in
(I)
+ hout | aout (p0 )(x) (x)ain (p0 ) |in i

(II)

5.5. LSZ REDUCTION FORMALISM

73

and again
Z

0
i
(II) = lim
d3 y hout | T ((y)(x)) |in i 0 eip y
0
y
Z
Z

0
i
lim
d3 y hout | T ((y)(x)) |in i 0 eip y
y 0
Z
Z

i
0
=
d4 y hout | T ((y)(x)) |in i ( 2y + m2 )eip y
Z
and we ignore term (I) again, since it does not pertain to the interaction.
Applying this reduction algorithm until the in- and out-states have been
reduced to the vacuum, and every time ignoring disconnected parts, gives
the following general result:

hp1 . . . pn out|q1 . . . qm ini =
iqi xi

n+m Z Y
m

d4 xi

i=1

Z Y
m

d 4 yj

(5.19)

j=1

(~x2i + m2 ) h0| T ((y1 ), . . . , (yn ), (x1 ), . . . , (xm )) |0i ( 2yj + m2 )eipj yj

which, although rather long, does look nicely symmetric.

5.5.1

Generating functional

In elementary particle physics, one is mostly interested in S-matrix elements,


to calculate scattering amplitudes and eventually cross sections. These can
be related to time-ordered vacuum expectation values, as we have seen:
h0| T ((x1 ), . . . , (xn ) |0i
Vacuum expectation values can be obtained by differentiating the generating
functional Z(j):

Z(j) = h0| T
since

Z
exp(i


d x(x)j(x) |0i
4

(5.20)


 n

1
Z

= h0| T [(x1 ), . . . , (xn )] |0i
i
j(x1 ) . . . j(xn ) j=0

Remarks
The S-matrix, which really is a unitary operator, can be expressed in
terms of Z(j) as well:

Z




4
2
2 1

S =: exp
d y in (y) (y + m )
: Z(j)
Z j(y)
j=0

74

CHAPTER 5. INTERACTING FIELDS, S-MATRIX, LSZ


(again with normal ordering: annihilation operators to the right of
creation operators). Exercise: show this.
The T-product vacuum expectation values are the objectives of axiomatic QFT, which postulates spectral conditions, Lorentz/Poincare
invariance and microcausality, and seeks to build up its theories from
these postulates.
In perturbation theory, the Fourier transform of h0| T ((x1 ), . . . , (xn )) |0i
produces poles in the outer momenta, which have to be truncated by
the Klein-Gordon differential operators.

Chapter 6

Invariant Perturbation
Theory
6.1

Dyson expansion, Gell-Mann-Low formula

To obtain the Gell-Mann-Low formula, one postulates a unitary operator


U(t) which transforms in and in into and :
(~x, t) = U1 (t)in (~x, t)U(t)

(6.1)

(~x, t) = U1 (t)in (~x, t)U(t)

(6.2)

This would allow for canonical commutation relations for and in . However, the representation of the commutators is unique up to unitary equivalence only in QM, where the number of degrees of freedom is finite.
The Hamiltonian H can be split up into two terms, one for the interaction
and one for the free part:
H = H0 + Hint
These are all operator-valued
R 3 4 functions of field operators; for example, in

4
-theory, Hint = 4! d x (~x, t). The Heisenberg equations give the time
developments of in and as the equal-time commutators with the relevant
Hamiltonians:
in (x)
= i[H0 (in , in ), in ]ET
t
(x)
= i[H(, ), ]ET
t

(6.3)
(6.4)

Plugging eq. (6.1) allows us to rewrite the left-hand side of eq. (6.4) as
U1
in
U

=
in U(t) + U1 (t)
U(t) + U1 in
t
t
t
t
75

(6.5)

76

CHAPTER 6. INVARIANT PERTURBATION THEORY

We know the middle term of eq. (6.5) from eq. (6.3). Sandwiching eq. (6.5)
between U(t) and U1 (t) gives:
U(t)

1
U1 (t)
U(t) 1
U (t) = U(t)
in + i[H0 (in , in ), in ] + in
U (t)
t
t
t

Since H is a polynomial in and , for which we have eqs. (6.1) and (6.2),
respectively, we know
U(t)

1
U (t) = iU(t)[H(, ), ]U1 = i[H(in , in ), in ]
t

and so we can conclude that


U(t)

1
U (t) = i[H(in , in ), in ]
t

Since U is unitary, we know that


UU1 = 1

and hence

Ut U1 + (t U)U1 = 0

Using H H0 = Hint and the above gives


i[Hint (in , in ), in ] = [(t U(t))U1 (t), in ]

(6.6)

where we could in many cases also Rwrite Hint (in ), since Hint usually does

not depend on in (e.g. Hint = 4!


d3 x4 (~x, t)). So, up to constants that
are of no interest to us, we have
(t U(t))U1 (t) = iHint (in )
Defining
U(t, t0 ) := U(t)U1 (t0 )

(6.7)

U(t, t0 )
= iHin (in (t))U(t, t0 )
t

(6.8)


 Z t
00
00
U(t, t ) = T exp i
dt Hint (in (t ))

(6.9)

we have

which is solved by
0

t0

where the time-ordering applies to the terms in the exponential series.


Exercise: show that this is a solution. Hint: transfer into an integral
equation:
Z
U(t, t0 ) = 1 i

t0

dt00 Hint (in (t00 ))U(t00 , t0 )

Solve this integral equation by iteration and doing the nested integrations
over a generalized box, instead of a tetrahedron, using the time ordering

6.1. DYSON EXPANSION, GELL-MANN-LOW FORMULA

77

prescription (of any two fields, the one at the prior time is to the right of
the one to the left). Alternatively, check that U(t1 , t2 )U(t2 , t3 ) = U(t1 , t3 );
then, with t1 t2 = t, the differential equation (eq. (6.8)) is fulfilled.
we avoid, unlike in the case presented by Itzykson & Zuber, U(t) =
RNote:
t
T( . . . ), which, with U() = t would give () = in .
Now, assume x01 x02 . . . (if this is not the case, rename your x so
that it is the case). Then,
h0| T((x1 ) . . . (xn )) |0i = h0| U1 (t)U(t, t1 )in (x1 )U(t1 , t2 )
. . . in (xn )U(tn , t)U(t) |0i
Now taking t gives
1

lim h0| U

(t)T in (x1 ) . . . in (xn ) exp i


dt Hint (t )
U(t) |0i
0

where again the exponential is to be split up into time-ordered parts. Now


consider
!

lim U(t) |0i = |0i

and

lim U(t) |0i = + |0i

where the s are phase factors resulting from Us unitarity. The proof of
these two limits is as follows (compare Bjorken & Drell, vol. II, p. 186):
assume
lim h| U(t) |0i =
6 0
t

where h| =
6 h0|. Then take a particle with momentum p out of h|:
hp in | U(t) |0i = hin | ain (p)U(t) |0i =

Z
3 ip0 t0 i~
p~
x
hin | in (t0 , ~x)U(t) |0i
i d xe
x00
and note that in (t0 , ~x)U(t) = U(t0 )(t0 , ~x)U1 (t0 )U(t). Now let
t = t0 ; this gives
=0

z }| { Z

= Z hin | U(t) ain (p) |0i +i d3 xeip0 ti~p~x


hin | (t U(t))(t, ~x) + U(t)(t, ~x)(t U1 (t))U(t) |0i
which gives zero. To see this (exercise), use the unitary equivalence and
the equation for t U derived above. Now, since we had assumed it to be
nonzero, this is a contradiction.

78

CHAPTER 6. INVARIANT PERTURBATION THEORY


Now, we obtain + as follows:
1 = h0|0i = lim h0| U1 (t) U(t)U1 (t) U(t) |0i =
t
|
{z
}
U(t,t)

+ h0| U(, ) |0i =


 Z


+ h0| Texp i
dtHint (in (t)) |0i

which leads us to our final result: the Gell-Mann-Low formula:


h0| T((x1 ) . . . (xn )) |0i =

h R
i

h0| T in (x1 ) . . . in (xn )exp i dtHint (in (t)) |0i


h R
i

h0| exp i dtHint (in (t)) |0i

(6.10)

This will be our basis when doing perturbation theory. Note that we have
one vacuum |0i, and that everything is written in terms of free in -fields,
obtained via the Yang-Feldman equations, acting in Fock space.

6.2

Interaction picture

For completeness sake, we will present the interaction picture (also known
as Dirac picture), since many textbooks use it to derive the Gell-Mann-Low
formula. It also offers more insight in the differences between the different
pictures.

6.2.1

Transformation

In the interaction picture, the time dependence is divided over the operators
and states, unlike in the Heisenberg and Schrodinger pictures. The time developments of those parts of operators that are not explicitly time-dependent
are given by H0 ; those of states by Hint . In the Schrodinger picture,
H = H0 + Hint
At t = 0, all three pictures (Heisenberg, Schrodinger and Dirac) are assumed
to coincide. From there, we obtain field operators in the interaction picture
as follows:
0 (~x, t) = eiH0 t (~x, 0)eiH0 t
(6.11)
The Heisenberg analogue is
(~x, t) = eiHt (~x, 0)eiHt
Note that H is time independent in all pictures.

(6.12)

6.2. INTERACTION PICTURE

79

The transformation from the Dirac to the Heisenberg picture,


t iHt
(~x, t) = e|iHt e{ziH0}t 0 (~x, t) e|iH0 {z
e }
e (t)
U

(6.13)

e
U(t)

e with the following property:


defines a unitary operator U
e
dU(t)
= ieiH0 t (H H0 )eiHt = ieiH0 t Hint eiHt
dt

(6.14)

Defining H0int := eiH0 t Hint eiH0 t , we can rewrite this as


e
dU(t)
e
= ieiH0 t Hint eiH0 t eiH0 t eiHt = iH0int eiH0 t eiHt = iH0int U(t)
dt
e
This gives for U(t),
like for U(t, t0 ) before, the following iterative formula:
Z

e
U(t)
=1i

e 0)
dt0 H0int (t0 )U(t

(6.15)

From this, one can again obtain the Gell-Mann-Low


formula, this time with

0 of the interaction picture, for which
expectation
values
in
the
vacuum

H0 0 = 0. Note that H0int is a polynomial of field operators of the interaction picture.
H0 now is the free Hamiltonian with the mass term 12 m20 20 from the
original (also called bare) Lagrangian. This is generally different from the
H0 we had before, which gives the time dependence of in (x) and has a
(m)
physical mass m2 . The notation H0 is also used.
Starting again from the T-product vacuum expectation value (vev)
h0| T((x1 ) . . . (xn )) |0i
and transforming
to 0 we now want to know the limit (in the weak sense)

of
e (T )
U
0 as T goes to . Shifting the energy such that both H0 0 = 0
and H |0i = 0, one has


h| eiHT eiH0 T
0 = h| eiHT 0 =
Z

h|0i 0|0 + dE h|Ei E|0 eiET


R
where dE h|Ei E|
0 eiET 0 for T by the Riemann-Lebesgue
lemma. From this, we read off


e (T )
U
0 * |0i 0|0
as T
(6.16)
which is the Gell-Mann-Low theorem.

80

CHAPTER 6. INVARIANT PERTURBATION THEORY


Using


e )U
e (T ) 0
0 U(T


h0|0i = 1 = lim
T
| 0|0 |2

2
we can solve for | 0|
0 | and obtain
h0| T((x1 ) . . . (xn )) |0i =
h R
i


T

0 T 0 (x1 ) . . . 0 (xn )exp i T H0int (t)dt 0


h R
i
lim


T
0 T exp i T H0 (t)dt 0
int
T

(6.17)

if we substitute |0i according to eq. (6.16) and plug in eq. (6.13) for .

6.2.2

Scattering theory

If one wants to start the discussion of the interaction picture from the beginning, one has to introduce scattering states, like in QM. States in the
interaction picture have a time development:
i

d
|iI = VI (t) |iI
dt

(6.18)

Note that VI was called H0int in the previous section, so


H = H0 + VSch

(6.19)

VI = eiH0 t VSch eiH0 t

(6.20)

and
For large positive or negative values of t the interaction is switched off.
|()i are time independent and correspond to states in the Schrodinger
picture changing in time with H0 .
Navely, H and H0 are assumed to have the same spectrum and to act
in the same Hilbert space. However, this has to be checked carefully; for
example, the bare masses in H0 are not the physical masses appearing
in the full H unless one shifts terms in the Hamiltonian H from V to H0
(adding and subtracting mass terms 21 (m20 m2 )2 ).
Consider stationary states
H0 | i = E | i


H = E
where the scattering states | i serve to construct wave packets which approach | i for t (see figure 6.1); i.e.
Z


d eiE t g()

d eiE t g() | i

(6.21)

6.2. INTERACTION PICTURE

81

Figure 6.1: | i as t
Formally, this means

= () | i or

| i = ()
with Hermitian (t) = eiHt eiH0 t , also called the Mller operator. Now, it
is easy to check that
d
i (t) = VI (t)
dt
which is solved by
 Z t

0

(t) = Texp i
dtVI (t )
0

(where (0) = 1). Using this , a solution to eq. (6.18) can be written
|()iI = ()() |()iI
| {z }
=| i

with the S-matrix:

S = ()() = Texp i


dt VI (t )
0

(6.22)

S-matrix elements are given by


D
E
|+ = h | ()() | i = h | S | i

(6.23)

Evaluating the S-matrix operator with the Wick theorem will also lead to
invariant pertubation theory and the Feynman rules.

82

CHAPTER 6. INVARIANT PERTURBATION THEORY

Note
If one wants to stay close to QM, one considers the Lippmann - Schwinger
equations for | i:


= | i + (E H0 i)1 V

(6.24)

which on the face of it looks like a simple identity, but contains the asymptotes from eq. (6.21).
The evaluation of the S-matrix elements could also be done in time
ordered perturbation theory, with on-shell intermediate states. To get the
right i prescriptions, one goes to the Lippmann-Schwinger equations and
introduces

T+


h| V + = T+
fulfilling

Z
= V + dV T+
(E E + i)

Then, one has


S = ( ) 2i(E E )T+

The equation for T+


can be solved by a Neumann series. This results in
a set of rules equivalent to the Feynman rules; later on, we will obtain the
denominators by integrating k0 in the Feynman graphs.

6.2.3

Background field

If one has just an outer field j (also called background field), one has to
distinguish |0iin and |0iout , but apart from this, the reduction formula works
like the LSZ formalism. One then has to consider
h0|out T((x1 ) . . . (xn )) |0iin
with
and

(6.25)

(x) = U

(t)in (x)U(t)

 Z t
0
0
U(t) = Texp i
dt Hint (t )

where Hint =

d3 x(x)j(x). Now, h0|out and h0|in are related by


h0|out = h0|in S

with


S = Texp i


dt Hint (t )
0

(6.26)

6.2. INTERACTION PICTURE

83

Then, (6.25) reads:




lim h0|in Texp i


dt Hint (t ) U1 (T )U(T )U1 (t)
0

in (x1 )U(t1 , t2 ) . . . in (xn )U(tn )U1 (T )U(T ) |0i =



 Z

0
0
dt Hint (t ) |0iin
h0|in T in (x1 ) . . . in (xn )exp i

(6.27)

This allows us to discuss the production of n particles out of the vacuum


by an outer source (like a current) j(x) switched on at some point in time,
and switched off at another. The calculation of the S-matrix transition
amplitudes and of transition probabilities leads to a Poisson distribution.
For more on this topic, see Itzykson and Zuber, p. 176.
Since this is a tree-level calculation, without loops, Z 6= 1 is not necessary
here, so we can caculate the amplitude directly:
h(p1 . . . pn )out |0iin =

in hp1 . . . pn | S |0iin

with


S = Texp i

dt


d x(j(~x, t )in (~x, t ))
3

Splitting into time intervals and using eA eB = eA+B+[A,B]/2 , this becomes




S = exp i




Z Z
1
4
4
0
0
d xin (x) j(x) exp
d xd y(x y )[in (x), in (y)]j(x)j(y)
2
4

Only the a -part of in (x) contributes to the first part; the second
part is a common factor


1
exp
2

Z Z




Z
1 f
2
d xd yDR (x y)j(x)j(y) = exp
dk|j(k)|
2
4

with |j(k)|2 = j(k)j (k) = j(k)j(k).


S produces a coherent state out of |0iin .
The probabilities for producing n particles (see Itzykson & Zuber, p.
169)
n
 Z

Z
1
2
2
f
f
pn =
dq|j(q)|
exp dq|j(q)|
n!
form a Poisson distribution.

84

6.3

CHAPTER 6. INVARIANT PERTURBATION THEORY

Haags theorem

Assume
(i) translational invariance of the physical vacuum (i.e. the one from the
Heisenberg picture)

(ii) vacuum polarization, i.e. 0 is not an eigenstate of H
e
Then, one finds that U(t)
does not exist. In other words, the interaction
picture does not exist in QFT.
This is a typical problem with the limit V . It also appears in
the thermodynamical limit of statistical mechanics (see also R. Haag, Local
Quantum Physics, Springer monographs in physics). The essential point is
that H does not exist in the Fock space HF of the H0 -states. H and H0
act in non-equivalent Fock spaces; they exist as different irreducible representations of the commutation relations in the case of an infinite number of
degrees of freedom.
Sketch of a proof (see Hegerfeld, Gottingen): define
Z


0 = lim
d3 x(H0 (x) + Hint (x)) 0 =: |i
H
V


By assumtion (ii), |i =
6 0 . Now, from the Gell-Mann-Low theorem

e (t) 0
U
|0i = lim

e (t) 0
t h0| U

we see that
0 has to be translationally invariant, because |0i is (assumption
(i)). Hence


0 H(x) |i = C (const)


since both
0 and |i are translationally invariant. Therefore,
Z


h|i = lim
d3 x 0 H(x) |i =
V

for C 6= 0.
Alternatively, one might use that h|i
is finite, so C must be zero, so

h|i = 0, which means that |i = H 0 = 0, which contradicts assumption
(i).

Chapter 7

Feynman rules, cross section


7.1

Wick theorem

Last chapter, we ended with the Gell-Mann-Low formula, which contained


time ordered products of in (x) (Hint is also an expression in in ). We
want to rewrite these in terms of c number valued functions, which will
later on turn out to be distributions, and normal ordered products, where
the annihilation operators are placed to the right of the creation operators.
The latter do not contribute to vacuum expectation values, but they are still
important in time ordered perturbation theory, where they act on 0 -states.
In this chapter, we will omit the subscript in from in , and simply
write . Now, let us start by considering
T((x1 )(x2 ))
Recall that (x) = (x) + + (x), where (x) contains the generation
operator, and + (x) the annihilation operator. Further, assume that x01
x02 . Then,
T((x1 )(x2 )) = (x1 ) (x2 ) + + (x1 )+ (x2 )+
(x1 )+ (x2 ) + + (x1 ) (x2 )
In this expression, the first three terms on the right hand side are already
written in their normal ordered form. The fourth term, + (x1 ) (x2 ), still
has the annihilation operator to the left of the creation operator. Introducing
the normal ordering, this term has to be removed:
T((x1 )(x2 )) =
: (x1 )(x2 ) : + + (x1 ) (x2 ) (x2 )+ (x1 ) =
: (x1 )(x2 ) : + D+ (x1 x2 )(x01 x02 )
85

(7.1)

86

CHAPTER 7. FEYNMAN RULES, CROSS SECTION

D+ (x1 x2 ) = h0| (x1 )(x2 ) |0i gives the first part of DC . In the case
where x02 x01 , i.e. where 1 and 2 are exchanged, gives
T((x1 )(x2 )) = : (x1 )(x2 ) : +D+ (x2 x1 )(x02 x01 ) =
: (x1 )(x2 ) : D (x1 x2 )(x02 x01 )

(7.2)

which is the second part of DC . To see the connection to D , recall the


definition of (x):
Z


a(k)eikx + a eikx
(x) = dk
R
ikx ). Then, for x0 x0 , the commutator of + (x )
f
(where + = dka(k)e
1
1
2
and (x2 ) gives
Z
+

1 dk
2 eik1 x1 +ik2 x2 [a(k1 ), a (k2 )] =
[ (x1 ), (x2 )] = dk
Z
1 eik1 (x1 x2 ) = D+ (x1 x2 ) =
dk
h0| (x1 )(x2 ) |0i

(7.3)

This is a c number. Taking the vacuum expectation value of T((x1 )(x2 )),
the normal ordered part drops out, and the second part has exactly the
structure of the Feynman propagator we have seen before. Therefore, the
vacuum expectation value h0| T((x1 )(x2 )) |0i can be written as
h0| T((x1 )(x2 )) |0i =
+

D (x1

x2 )(x01

x02 )

(7.4)

D (x1
Z
DF (x1 x2 ) = iGF (x1 x2 ) = i

x2 )(x01
d4 k

x02 ) =
eikx

(2)4 k 2 m2 + i

Figure 7.1: Integration path for DF

(7.5)
(7.6)

7.1. WICK THEOREM

87

Now, Wicks theorem states that the general T-product can be reduced to
normal ordered products and vacuum expectation values of T-products of
pairs of :
T((x1 ) . . . (xn )) =

(7.7)

: (x1 ) . . . (xn ) : +
h0| T((x1 )(x2 )) |0i : (x3 ) . . . (xn ) : + +
h0| T((x1 )(x2 )) |0i h0| T((x3 )(xn )) |0i : (x5 ) . . . (xn ) : + +
...

going over all possible contractions

Here, h0| T((x1 )(xn )) |0i is called the contraction of (x1 ) and (x2 ),
which is often denoted by (x1 )(x2 ).
The theorem is proven by induction: first, we see that it works for n = 2.
Then, we prove that if the theorem holds for n, it also holds for n + 1. To
begin the proof, let us note that without limitation of generality, we can
assume that x0n+1 is smaller than all other x0 , so
T((x1 ) . . . (xn )(xn+1 )) = T((x1 ) . . . (xn ))(xn+1 )
Now, by assumption, T((x1 ) . . . (xn )) can be written in terms of normal
ordered products. Rewriting (xn+1 ) as + (xn+1 ) + (xn+1 ), we see that
the term with + is already normal ordered. The other term still has to be
ordered, so (xn+1 ) has to be commuted with all the + .
[+ (xk ), (xn+1 )] =c-number = h0| [+ (xk ), (xn+1 )] |0i =
h0| + (xk ) (xn+1 ) |0i = h0| (xk )(xn+1 ) |0i =
h0| T((xk )(xn+1 )) |0i
Commuting with all the + gives all possible contractions, as Wicks theorem says.
Hint is often assumed to be already normal ordered, e.g. : 4 (x) :. Otherwise, one would get self-contractions, as we will see later when treating
the path integral formulation.

7.1.1

4 -theory

Evaluating interactions in 4 -theory by Wicks theorem highlights the central problem of local QFT: it involves polynomials in GF . If GF is a distribution (like a delta function), taking it to any power other than one is a

88

CHAPTER 7. FEYNMAN RULES, CROSS SECTION

very tricky affair. Consider T(: 4 (x) :: 4 (y) :):


T(: 4 (x) :: 4 (y) :) = : 4 (x)4 (y) : +
 2
4
1!
(x)(y) : 3 (x)3 (y) : +
1
 2 
2
4
2!
(x)(y) : 2 (x)2 (y) : +
2
 2 
3
4
(x)(y) : (x)(y) : +
3!
3

4
4! (x)(y)
(7.8)


and note that the last three terms have (x)(y) , which was shown to
be DF , to some power higher than 1.

7.2
7.2.1

Feynman graphs
Feynman rules in x-space

We want to evaluate the Gell-Mann-Low formula for the T-product vacuum


expectation value of interacting fields from last chapter:

R
h0| T[exp i d4 xHint ((x)) (x1 ) . . . (xn )] |0i


R
h0| T exp i d4 xHint ((x)) |0i
For the sake of simplicity, let us restrict ourselves to real 4 -theory, where

: 4 :. Each polynomial T-product vacuum expectation value is


Hint = 4!
evaluated with Wicks theorem - we can forget about the normal ordered
parts, as they do not contribute. The best way of bookkeeping is by means
of graphs: the pionts x1 , . . . , xn are to be connected by propagators and
vertices:

The first term here is zeroth order in ; not in the figure are the other
possible disconnected combinations. The second term is first order in and
comes with a factor of 4!, since there are 4! different possibilities to connect
the lines to the vertex. The third term has a factor of (4! 4!/2) 2!: two

7.2. FEYNMAN GRAPHS

89

vertices, hence (4!)2 , but exchanging the two inner lines is an automorphism
and should hence not be counted; the 2! is for the possibility of exchanging
y1 and y2 .

These are further graphs of order 2 , except for the last one, which is 3 .
Considering a general graph, we see from the above that when calculating
the statistical factor for exchanging vertices, we have to reckon with the
following:
Exchanging vertices (figures 7.2.1 and 7.2.1): this yields topologically
different graphs in the first case (fig. 7.2.1), but not in the second case
(fig. 7.2.1). So, we have 2 2-automorphisms, yielding a factor 21 12 .
Exchanging lines (fig. 7.2.1): this yields topologically identical graphs.
Note that the vertex exchange in fig. 7.2.1 can also be viewed as an
exchange of the lines connected to the vertices y3 and y4 .

Figure 7.2: Vertex exchange (1)

Figure 7.3: Vertex exchange (2)

90

CHAPTER 7. FEYNMAN RULES, CROSS SECTION

Figure 7.4: Line exchange


In case of doubt, counting the number of possible contractions is a way
out. This is more cumbersome than drawing the graphs, but also more
reliable.

drops out, since one has to sum over all combinations


The factor 4!1 in 4!
of the four lines connecting to the rest of the graph. Similarly, in the Gell1
in the exponential is cancelled by the
Mann-Low formula, the factor n!
n! ways to arrange the vertices. In loops, there are combinatorial factors
correcting the fact that situations are counted several times: one has to
divide by the number of automorphisms ()graph of inner lines and vertices
which leave the graph unchanged.

Figure 7.5: Automorphisms

The free propagator = h0| T(in (xi )in (xk )) |0i has the form
Z
= DF (xi xk ) =
CF

d4 k
i
eik(xi xk )
4
2
(2) k m2 + i

(7.9)

where CF is the Feynman path in the imaginary k0 -plane (see fig. 7.1).
Remarks
for the interacting field ,
Z

h0| T((x)(y)) |0i = ZDF (x y, m) +

dm02 (m02 )DF (x y, m0 )

>m2

(7.10)

7.2. FEYNMAN GRAPHS

91

one has a prefactor Z (derivation like in chapter 5).

We will soon see that in the reduction formula for S-matrix elements,
the first part contributes.

The normal ordering in Hint ensures there are no tadpole graphs:

Figure 7.6: Tadpole graphs

There are both connected and vacuum graphs in the numerator, but
only vacuum graphs in the denominator.
Let us say there are p vertices in the numerator expansion: s of these
are connected to outer fields, and the other p s are in vacuum graphs.
Decomposing the graph into its connected and vacuum parts,
(i)p
h0| T(in (x1 ) . . . in (xn )Hint (y1 ) . . . Hint (ys )) |0i
p!
p!
h0| T(Hint (ys+1 . . . Hint (yp )) |0i
s!(p s)!
we see that the factors of p! cancel each other. This leaves us with
X
p,s

X
s,ps

X
X1
1
h. . . i
h. . . i
s!
(p s)!
s
ps
| {z } |
{z
}
no vac. graphs vac. graphs

The denominator cancels the vacuum graphs in the numerator.

We thus obtain the Feynman rules in x-space for Wightman functions


h0| T((x1 ) . . . (xn )) |0i:

92

CHAPTER 7. FEYNMAN RULES, CROSS SECTION

(i) draw all graphs without vacuum parts to some order in the vertices
s
(ii) a factor DF (x) for each propagator connecting outer points x1 , . . . , xn
and inner points (vertices) y1 , . . . , yn ; x is the difference between
endpoints (here the sign does not matter, but later it will represent
the direction of the propagator for charged particles)
(iii) a factor i for each vertex (coupling
(iv) combinatorial factors

1
()

(v) integrate over the inner points yi :

7.2.2

d 4 yi

Feynman rules in momentum space

It is also very useful to Fourier transform the function in the outer xj , i.e.
to perform the integrations
Z
d4 xj eikj xj . . .
with the outer momenta kj . Writing down a Fourier representation of DF ,
where outer points and inner vertices are on the same footing,
Z
eikx
d4 x
DF (x) = i
(7.11)
(2)4 k 2 m2 + i
one can perform the xj and yi integrations and obtain (2)4 4 (ki , . . . ). This
gives 4-momentum conservation at the vertices and overall momentum conservation. We can perform the integrations at the vertices, which leaves us
with one integration per loop and overall momentum conservation. This
gives the Feynman rules in momentum space:
(i) draw all graphs without vacuum parts to some order in the vertices
s
(ii) a factor

i
k2 m2 +i

for each propagator

(iii) a factor i for each vertex


(iv) combinatorial factors

1
()

(v) integrate over the loop momenta:


for outer momentum conservation

d4 ki
,
(2)4

and add a factor (2)4 4

7.2. FEYNMAN GRAPHS

93

Remarks
We still have disconnected graphs, and in particular lines running
through without interacting. This is where the S-matrix comes in,
where we have already taken out these cases while applying the reduction formula.
The vacuum graphs cancel in our case, so here they
very inR 4areixnot
0
4

, i.e. an
teresting. They contain an overall factor (0 ) = d xe
infinite volume element. This is related to the fact that these graphs
can be shifted in space independently of each other. If there are several disconnected vacuum pieces, they will appear as a sum in the
exponential:
h0| . . . |0ivac = exp

X


h0| connected vacuum pieces |0i

Exercise: prove this. Hint: use (a + b)n /n! =

7.2.3

bnk
k! (nk)! .

P ak

Feynman rules for complex scalars

For complex scalars, we have


Hint =


( (x)(x))2
4

(7.12)

The only propagators we have are of the form


h0| T((x) (y)) |0i = DF (x y)

(7.13)

with DF as in eq. (7.9). The other two possibilities give zero:


h0| T() |0i = h0| T( ) |0i = 0
Since we have complex scalar particles here, we need to change the rules:
particles propagate from y to x, whereas antiparticles propagate from x to
y. Therefore, an arrow is needed to indicate the direction of propagation.
Note that this is not related to momentum. See fig. 7.2.3 for a graphical
explanation. Also, the factor 14 cancels the number of possibilities to take
two and two out of Hint . Each vertex has two incoming lines, corresponding to incoming particles or outgoing antiparticles, and two outgoing
ones, corresponding to outgoing particles or incoming antiparticles. In momentum space, p0 has to be positive for particles and antiparticles, meaning
the momentum arrows are different from the particle/antiparticle arrows.

94

CHAPTER 7. FEYNMAN RULES, CROSS SECTION

Figure 7.7: Particles and antiparticles with momenta

7.3

Functional relations

As already indicated before, Wightman functions h0| T((x1 ), . . . (xn )) |0i


with interacting fields can be expanded via the generating functional
Z
Z(j) = h0| T{exp(i

d4 x(x)j(x))} |0i

(7.14)

Z
by a series of partial differentiations ( 1i )n j(x1 )...j(x
|j=0 . In the Gell-Mannn)
Low formula, the T-products of interacting fields
are
R 4 expressed by those of
in -fields, including the S-matrix T{exp[i d yHint (in (y))]}. Thus,
we obtain
 n
1
h0| T((x1 ) . . . (xn )) |0i =

(7.15)
i



R 4
R 4
1

exp[i
d
yH
]
h0|
T
exp
i
d
x
(x)j(x)
|0i
int
in
i j(y)
j(x1 )...j(xn )


R

h0| T exp i d4 xin (x)j(x) |0i

j=0

(j)

This gives vacuum graphs with interaction Hint (x) = in (x)j(x), which
exponentiate to


Z
1
4
4
exp
d x1 d x2 j(x1 )DF (x1 x2 )j(x2 )
2
which, with

Hint

1
i j(y)

=
4!

j(y)

4

7.4. BACK TO S-MATRIX ELEMENTS

95

reproduces the Feynman rules of 4 -theory. Exercise: show that from the
term (j)2n /(2n)!, one obtains


DFn (2n 1)(2n 2) . . .
DF n
=
/n!
(2n)!
2
Graphs with through-going lines are special cases of disconnected diagrams. The latter can be decomposed into connected diagrams by definition; connected diagrams have all outer lines connected and no separable
vacuum graph pieces present (so unseparable vacuum graphs are also called
connected). An example of a disconnected graph:

+
The generator W (j) of such diagrams is related to Z(j) by
Z(j) = eiW (j)

or

(7.16)

W (j) = ilnZ(j)
for normalized Z(j)

Z(j)
Z(0) .

Now,



W

= i (x1 , . . . , xn )
j(x1 ) . . . j(xn ) j=0

(7.17)

corresponds to connected diagrams with n outer lines.


Note
For a graph without any outer lines, use W (0). Sum of connected
vacuum diagrams: parts cannot be taken apart. This is the case
considered before.

7.4
7.4.1

Back to S-matrix elements


2-point functions

Now, let us consider the Wightman function, also called 2-point function or
dressed propagator, for interacting fields . The vacuum expectation value
h0| T((x)(y)) |0i, in 4 -theory, can be expanded in the language of the
Feynman-graphs as depicted in fig. 7.4.1.
The first row gives DF , for the first graph, and DF (i)DF for the other
graphs, where is the self-energy or vacuum polarization. These are the

96

CHAPTER 7. FEYNMAN RULES, CROSS SECTION

Figure 7.8: Feynman graph representation of h0| T((x)(y)) |0i


so-called one-particle irreducible graphs, which cannot be further reduced
to two-point graphs by cutting just one inner line. The other rows give a
geometrical series:
DF + DF (i)DF + DF (i)DF (i)DF + =
DF (1 + iDF )1 = (DF1 + iDF DF1 )1 = (DF1 + i)1

(7.18)

For Fourier transformed 2-point functions, with relative and overall momentum conservation, this reads
"
#1
1
i
i
+ i(p2 )
= 2
(7.19)
2
2
2
p m + i
p m (p2 ) + i
The dressed propagator does not contain a particle pole at p2 = m2 anymore:
the denominator now vanishes at p2 = m2 + (p2 ), moving the pole to some
other m2 . We rename the original m2 m20 , called Lagrange mass or
bare mass, and consider the physical mass m2 = m20 + (m2 ). This means
we have to assume a real , which will be discussed later. Expanding (p2 )
around p2 = m2 , i.e. around the physical mass, we get
p2

m20

((m2 )

(p2

i
=
m2 )0 (m2 ) + rest ) + i

+
i
=
2
2
2
2
p m (p m )0 (m2 ) + rest + i
i(1 0 (m2 ))1
p2 m2 rest /(p2 m2 ) + i
As p2 m2 , rest /(p2 m2 ) goes to zero, so
lim

p2 m2

i(1 0 (m2 ))1


iZ
= 2
2
2
2
2
p m rest /(p m ) + i
p m2 + i

(7.20)

7.4. BACK TO S-MATRIX ELEMENTS

97

with Z = (1 + 0 (m2 ))1 . This Z is the same as in the LSZ-formalism from


chapter 5.

7.4.2

General Wightman functions

In general Wightman functions


h0| T((x1 ) . . . (xn )) |0i
the outer lines of the corresponding Feynman graphs also contain these selfenergy subgraphs, each with their own inner lines depending on the order
of perturbation theory to which one goes. So, in the Fourier transformed
iZ
2
2
expression we have p2 m
2 +i for p m .

Figure 7.9: Self-energy of outer lines

Now, let us substitute the Fourier transformed vacuum expectation values into the LSZ reduction formula. With only one incoming q1 , we have
Z
Z 4
d q1 iq1 x1
i
4
iq1 x1
2

d x1 e
(x1 + m )
e
F (h0| T(. . . ) |0i) =
(2)4
Z
Z 4
d q1
iZ
i

(
q12 + m2 )(2)4 4 (q1 q1 ) 2
4
2
(2)
q1 m + i + rest /(
q12 m2 )
Z
(7.21)
Since the outer particle q1 is on the mass shell,
the only pole is at q12 =

q12 = m2 . Therefore, eq. (7.21) condenses to Z. We will discuss ways to


get rid of this remaining factor in the context of renormalization theory, by
renormalizing the field (this is known as wave function renormalization).

98

CHAPTER 7. FEYNMAN RULES, CROSS SECTION

Note
i(p2 ) is a divergent quantity in continuum QFT. Consider e.g. the tadpole contribution:
Z
Z
d4 kEucl

4
=
i
itadpole = (i)
d
k
2
(2)4
(2)4
k 2 m20 + i
(kEucl
+ m20 )
The Wick rotation
x0 = ix4
k 0 = ik 4
~k = ~kEucl
~x = ~xEucl

(7.22)

kx = k 0 x0 ~k~x = k 4 x4 + ~kEucl ~x = (kx)Eucl

(7.23)

gives

Z
=

1
dk 2 k 2
d4 2
2
k + m20

(It is not outer p2 -dependent in this case). Introducing a cut off 2 in the
k 2 -integration, we get
(Z 2
)
Z 2
Z 2

2 1
k2

dk 2
2
2
2
2
= 2
=
dk 2
dk m
2
(2)4 0
k + m2
16 2
k 2 + m2
0
0
|
{z
}
2 m2 ln

2 +m2
m2

In the general case, there also is a p2 -dependence, which we will discuss


later on in detail. Note that if we want to fix a physical mass (m2 ), m20
becomes cut off dependent by m20 + = m2 .

7.5
7.5.1

From S-matrix to cross section


Derivation

The S-matrix gives the transition amplitude from in- to out-states:


Sf,p1 p1 = hout |p1 , p2in i

(7.24)

for scattering two incoming particles p1 and p2 into a final state |i, where
the in-states have fixed momentum, i.e. are plane waves. Now, we take out
the trivial part of S, where particles are only running through:
S = 1 + iT

(7.25)

7.5. FROM S-MATRIX TO CROSS SECTION

99

Making energy-momentum conservation explicit, we have


h| T |p1 , p2 i = (2)4 4 (pf p1 p2 ) h| M |p1 , p2 i

(7.26)

In the end, we want to calculate the transition probability. In order to handle


the -functions properly, let us, before squaring, introduce wave packets for
incoming states:
Z
g1 f1 (p1 ) |p1 i
|p1 i dp
where |f1 (p1 )|2 is peaked arbitrarily strongly around p1 . Now, we obtain for
the transition probability
Z
Z
f0 (2)4 4 (p0 + p0 p1 p2 )
f0 dp
f
f
Wp1 ,p2 = dp1 dp2 dp
2
1
1
2
(2)4 4 (pf p1 p2 )f1 (p1 )f1 (p01 )f2 (p2 )f2 (p02 )


h| M |p1 , p2 i h| M p01 , p02
R
With (2)4 4 (p01 + p02 p1 p2 ) = d4 x exp(ix(p01 + p02 p1 p2 )) and
R
f f1 (p1 )eixp1 = f1 (x) (and similar for f , f2 and f ), this becomes
dp
1
1
2
Z
Wp1 ,p2 = d4 x|f1 (x)|2 |f2 (x)|2 (2)4 4 (pf p1 p2 )| h| M |
p1 , p2 i |2
Neglecting the variation of the transition amplitude over the wave packet
and assuming fi (x) = eipi x Fi (x) with slowly varying Fi (x) (i.e. fi (pi ) is
strongly peaked around pi ), the transition probability per volume per time
interval is
dW
= |f1 (x)|2 |f2 (x)|2 (2)4 4 (pf p1 p2 )| h| M |
p1 , p2 i |2
dV dt

(7.27)

Now, we still need to normalize the probability current for the positive
energy solution: it should satisfy the continuity equation

if (x) f(x) 2
p |f(x)|2
(7.28)
where the f fulfill the Klein-Gordon equation. The -symbol is used here,
because f (p) is strongly peaked, so f (x) is almost a plane wave.
Now, when we go into actual experiment, we have a target, at rest in the
laboratorys frame, which is being bombarded with particles. In the target,
the number of particles per volume element is:
dn2
= 2
p02 |f2 (x)|2 = 2m2 |f2 (x)|2
dV
The incident flux of bombarding particles is given by
|~
p1 |
p1 ||f1 (x)|2
2p01 |f1 (x)|2 = 2|~
| {z }
p01
|{z}
density
velocity

(7.29)

(7.30)

100

7.5.2

CHAPTER 7. FEYNMAN RULES, CROSS SECTION

Scattering amplitudes and cross sections: an example

Consider two scalar fields, one complex () and one real (), with the following Lagrangian:
1
L = + V (, , )
2

(7.31)

where

1
V (, , ) = m2 2 + m2 + m + ( )2
2
4
Remember the formulas for
Z
(x) =
Z
(x) =

(7.32)

and :
n
o
f a (k)eikx + b (k)eikx
dk

n
o
f a(k)eikx + a (k)eikx
dk

(7.33)
(7.34)

Exercise: inserting this into the LSZ-reduction formula, the


(i) outgoing (a ) particles produce a factor
Z
eipx (~x2 + m2 ) h0| T(. . . (x) . . . ) |0i
(ii) incoming (a ) particles produce a factor
Z
eipx (~x2 + m2 ) h0| T(. . . (x) . . . ) |0i
(iii) outgoing (b ) antiparticles produce a factor
Z
eipx (~ 2 + m2 ) h0| T(. . . (x) . . . ) |0i
x

(iv) incoming (b ) antiparticles produce a factor


Z
eipx (~x2 + m2 ) h0| T(. . . (x) . . . ) |0i
We want to calculate the scattering amplitude for particles/antiparticles of
-type to order 2 (this is our choice of parameters). Fig. 7.5.2 represents particle-particle scattering; for antiparticle-antiparticle scattering, one
has the same graphs with reversed charge transport directions. Fig. 7.5.2
represents particle-antiparticle scattering.

7.5. FROM S-MATRIX TO CROSS SECTION

101

Figure 7.10: Particle-particle 4 -interaction

Figure 7.11: Particle-antiparticle 4 -interaction


To make the calculations a bit easier, let us introduce the so-called Mandelstam variables s, u and t:
(p1 + p2 )2 =s = (p3 + p4 )2
(p3 p1 )2 =t = (p4 p2 )2
(p4 p1 )2 =u = (p3 p2 )2

(7.35)

The sum of these three is always 4m2 :


s + t + u = 4m2

(7.36)

For cases (i) and (ii), the Feynman rules give:


i + (im)2

(p3 p1

i
i
+ (im)2
=
2
2
m + /i/
(p4 p1 ) m2 + /i/


2 m2
2 m2
i +
+
= iM(i)
t m2 u m2
(7.37)

)2

For case (iii), we have




2 m2
2 m2
i +
+
= iM(iii)
t m2 s m2

(7.38)

102

CHAPTER 7. FEYNMAN RULES, CROSS SECTION

So, exchanging (i) and (iii) comes down to exchanging s and u. Note that
we do not need the i-prescription here, since we do not reach the poles.
Note
Amplitudes are real analytic in two of the three Mandelstam variables; this is
called Mandelstam analyticity and has been studied extensively in the 1960s.
At that time perturbation theory in the case of the strong interactions, what
today is the field of quantum chromodynamics (QCD), was considered to be
just a guideline to analytic properties of amplitudes, whereas the unitarity
of the S-matrix, especially in the 2-2 channel, was considered to be most
important. See The analytic S-matrix by Eden, Landshoff, Olive and
Polkinghorne (and bootstrap ideas by Chew) for more details.
The amplitudes M can be used to calculate cross sections with the formulas explained before; in our case, this is particularly simple, since we have
only tree-level graphs, i.e. without loops. See the literature on elementary
particle physics for more on this topic.
We finally obtain our cross section by calculating
(transition probability) / (incident flux target density):
1
| hf | M |
p1 , p2 i |2
d = (2)4 4 (pf p1 p2 )
4m2 p1
For an n-particle final state in Lorentz covariant notation, this is:
Z
1
f . . . dp
f
dp
dn2 =
n+1
3
4[(p1 p2 )2 m21 m22 ]1/2

(7.39)

(2)4 4 (p1 + p2 p3 + p4 pn+2 )| hp3 , . . . pn+2 | M |p1 , p2 i |2


For the fermion case, we also need to sum over, or fix, the spin, but apart
from that, the result is the same (this is different from Itzykson and Zubers
book, where the 2m factor in u
u is not introduced; this will be discussed
later). For the case with n identical outgoing particles, there will also be a
combinatorial factor of 1/n!.

7.5.3

The optical theorem

Because of probability conservation, S = 1 + iT is unitary:


S S = 1

(7.40)

so, i(T T ) = T T. Consider

hk1 , k2 | T T |k1 , k2 i =

n Z
XY
n i=1

f hk1 , k2 | T |{qi }i h{qi }| T |k1 , k2 i


dq
i

7.5. FROM S-MATRIX TO CROSS SECTION

103

with {qi } a complete set of states. Taking out (2)4 4 (...), we obtain a
relation for M-matrix elements:
n Z
X
XY
f (2)4 4 (k1 + k2
dq
qi )
2=(M(k1 , k2 k1 , k2 )) =
i
n i=1

i
2

|M(k1 , k2 {qi })|

(7.41)

The second factor of (2)4 4 (. . . ) has been taken out here on both sides as
(2)4 4 (k1 + k2 (k1 + k2 )). This relates the imaginary part of the forward
scattering amplitude to the total cross section. This can be seen explicitly
for Feynman diagrams (Cutkosky cutting rules): [picture] Again, see Eden,
Landshoff, Olive and Polkinghorne, The analytic S-matrix for more details.
Remark
We have done scattering theory rather superficially - of course this can
be done better. See e.g. Weinberg (literature list) for a more elaborate
treatment.
Cutkosky rules in short
1
=P
x i

 
1
i(x)
x

(7.42)

R
Under the integral dx, P is the principal value, obtained by averaging over
both possibilities to go around x = 0. The Feynman-propagator is given by


1
1
=P
2i(p2 m2 )
(7.43)
p2 m2 + i
p 2 m2

M=

1 (i)2
2 i

d4 k1
i
i
2
4
2
(2) k1 m + i (p1 + p2 k1 )2 m2 + i

(7.44)

Z 4
()2
d k1
=(M) =
(k12 m2 )((p1 + p2 k1 )2 m2 )(2)2 =
2
(2)4
Z
Z
()2
d3 k1
d3 k2
(2)4 4 (k1 + k20 p1 p2 )
2
2k10,+ (2)3
2k20,+ (2)3
Now, the question is whether k1 and k2 are positive. Let us check:
p2

1
1
1
=

2
1/2
2
2
2
m + i
p0 (~
p + m ) + i p0 + (~
p + m2 )1/2 i

104

CHAPTER 7. FEYNMAN RULES, CROSS SECTION

Figure 7.12: Cutting graphs

as discussed before. Integrating around the poles as in the figure [picture],


we get from the part around p0 = +(~
p2 + m2 )1/2
 


1
1
2
2 1/2
(i) = P

i(p

(~
p
+
m
)
)
=
0
p2 + m2 )1/2
p2 + m2 )1/2
p0 (~
p0 + (~


1
P
2i(p2 m2 )
p 2 m2
and from p0 = (~
p2 + m2 )1/2 , we get
 


1
1
2
2 1/2
(ii) = P
+ i(p0 + (~
p +m ) )
=
1/2
2
2
2
p +m )
p + m2 )1/2
p0 + (~
p0 (~


1
2i(p2 m2 )
P
p 2 m2
Closing around the poles at k10 = (~k12 + m2 )1/2 ,
1/2

k10 = p10 + p20 = + (~k1 p~1 p~2 )2 + m2
in the lower half plane in the above integral (eq. (7.44)), we see that k10
and k20 are positive. This enables us to write down the Feynman rules in
momentum space for S-matrix elements; they differ from the previous ones
only
in that the outer propagator lines are truncated and that there is a
Z instead.

Chapter 8

Path integral formulation of


QFT
So far, we have studied QFT in the canonical formalism with operator-valued
fields, in the Heisenberg representation. Richard Feynman has formulated
another representation of both QM and QFT, which is very intuitive and
does not use operators. Its mathematical status, however, is still in development. This method is particularly powerful if one wants to quantize gauge
theories - this is why it is necessary to discuss it - and it also allows one
to derive the Feynman rules very easily, and to discuss problems beyond
perturbation theory, although we will not go into the latter here.

8.1

Path integrals in QM

The path integral formulation of QM centers around the transition amplitude for a QM particle from a position x(t) at a time t to a position x0 (t0 )
at time t0 . It starts from the Heisenberg picture, where the time dependent
operators X(t) and P(t) have their respective eigenvectors |x(t)i and |p(t)i,
with time developments
X(t) = eiH(tt0 )/~ X(t0 )eiH(tt0 )/~
iH(tt0 )/~

|x(t)i = e

|x(t0 )i

(8.1)
(8.2)

and similarly for P and p. The factors of ~ have been reinserted here for
clarity. Note that the sign in eq. (8.2) is opposite to that of the Schrodinger
equation. Note also that this equation describes a transition in time, while
the position does not change. Finally, in eq. (8.1), X(t0 ) = XS ; in the
Schrodinger picture, t0 is usually taken to be 0.
Let us start at the end of our discussion of the path integral formulation,
105

106

CHAPTER 8. PATH INTEGRAL FORMULATION OF QFT

with the result:


x (t )|x(t) =

" Z
DxDp exp i

t0


 #
dx
H(p, x) /~
d p( )
d

(8.3)

R
Here, Dx is a path integral, in mathematical circles know as functional
integral, an integral over all possible paths x( ) connecting x and x0 , with
x(t) = x and x(t0 ) = x0 .

Figure 8.1: Integrate over all possible paths


R

Dp does not have boundary conditions, since the problem asks for the
transition amplitude between positions, but not between momenta. (As a
small aside: note that the exponent is just the classical action times i/~.)
After discretization of the time integral, it becomes a product of integrals
at 1 , 2 , . . . over x(1 ), x(2 ), . . . :

Figure 8.2: Discretization of the time integral

Z

Z
Dx exp


d . . .

YZ
i

dx(i )

Defining := i i1 , one eventually has to take the limit 0, which is


mathematically demanding. Obviously, the integral over Dx does not need
to be evaluated at the endpoints, since these are fixed.

8.1. PATH INTEGRALS IN QM

107

Now, let us derive this result. From the canonical formalism, we have
the following formula:

0 0


0
x (t )|x(t) = x0 eiH(t t)/~ |xi
(8.4)
where |xi is a Schr
odinger state. Decomposing the interval into N bits,
0
t t = N , gives
hxN | eiH/~ |xN 1 i hxN 1 | eiH/~ |xN 2 i hxN 2 | . . .
. . . |x1 i hx1 | eiH/~ |x0 i
where xN = x0 and x0 = x, and H =
matrix elements, to order :

P2
2m

+ V (X). Consider one of these

iH
 + . . . |xk i =
hxk+1 | eiH/~ |xk i = hxk+1 | 1
~



Z
dpk
xk+1 + xk
hxk+1 |xk i i
V
hxk+1 |pk i hpk |xk i +
2~
2
 2 
P
xk
hxk+1 |pk i pk
+ =
(8.5)
2m






Z
dpk
i p2k
xk+1 + xk
ipk
1
exp
+V
(xk+1 xk ) + . . .
2~
~ 2m
2
~
Note: for more complicated X/P-mixed operators one needs Weyl ordering, a symmetrization of the operator sequence in X/P; see Peskin
& Schr
oder, p. 281 for more on this topic. Note that the argument in
V ( xn+12+xn ) is written like this for cosmetic reasons; we could just as well
have written xn , since in the end, the limit N will be taken. Continuing our derivation, let us define





Z
dpk
i p2k
xk+1 xk
xk + xk+1
() =
exp
+V
pk
(8.6)
2~
~ 2m
2

which is the right hand side of eq. (8.5) to order O(). Multiplying all s
and taking the limit  0, we have
" Z 0 
#
Z
t

0 0

px H(p, q)
x (t )|x(t) = DxDp exp i
dt
(8.7)
~
t
with x(t) = x and x(t0 ) = x0 . This limit is of course accompanied by some
higher-level mathematics. The nave expression, however, has to be based
on the discretized version we started from. Concretely, for physicists, this
means that in QFT, numerical lattice calculations are an adequate way to
approach this integral. Note that the continuous and differentiable functions
are a dense set of measure zero in the functional integral.

108

CHAPTER 8. PATH INTEGRAL FORMULATION OF QFT

The dpk -integration in eq. (8.6) can be performed: it is just a Gaussian


integral, here restricted to one dimension for simplicity. It is solved by
completing the square:
p2k 
+ pk (xk xk+1 ) =
2m
 2

(xk xk1 )2
1 (xk xk1 )2
1 pk
 2pk (xk xk+1 ) +
m + m
=

2 m

2

p0 2 
1 (xk xk1 )2
k
+ m
2 m 2


with p0k :=

pk
m

xk xk+1
.


Using the standard Gaussian integral,


Z
 1/2
2
dxex =

(8.8)

we can perform the p0k -integral. The first part becomes


r


Z
dp0k
1 i 0 2
1
1
2~m
y 2 /2
exp
pk =
dye
=
2 ~m
2~
2~
i
2~
p

where y = p0k i/~m. Note that due to the presence of i in the conversion from p0k to y, this substitution constitutes a 45-degree rotation of the
integration path:
Z

So, the final result, the product of all the separate integrals, is
N
1 Z
Y

r
dxj

j=1

m
i~2

N


 !
N
X
m (xi xi1 )2

exp i
V (xi ) /~
2
2
i=1

which, after taking  to zero and N to infinity, becomes:


x(t)|x (t ) =

t0

d L(x, x)/~

Dx exp i
t

(8.9)

8.1. PATH INTEGRALS IN QM

109

with singular integration measure ( N/2 ) and Lagrangian density L =


mx 2
2 V (x).
The x-space path integral is less general (namely only for H quadratic
in p) than the first version, but for our purpose, we can settle for this one.
Remarks
For hx0 (t0 )| T(O1 (t1 )O2 (t2 ) . . . ) |x(t)i, the operators O1 , O2 , . . . act
in the time slices around t1 , t2 , . . . . Then, hx(t1 )| O1 (t1 ) |x(t01 )i =
O1 (x(t1 ))(t1 t01 ) where O1 is a function. Thus, we obtain timeordering in the path integral, which is decomposed into time slices.
This remark is also important if the potential has the shape of a matrix in more complicated settings.
The oscillating behaviour of the Feynman exponential, which makes
the convergence of the integral a more subtle affair, can be avoided if
we go to imaginary, or Euclidean, time (t = x0 = ix4 = itE ). This
so-called Wick rotation helps us to define certain expressions properly.
Of course, one has to rotate back at the end of the calculation .

8.1.1

Vacuum expectation values

When going to QFT, we will be interested in vacuum expectation values (cf.


correlation functions in statistical physics). Let us briefly investigate them
here:
h0| T(x(t1 ) . . . x(tn )) |0i = ?
Let us start from
Z

xT

hxT (T )| T(x(t1 ) . . . x(tn )) |xT (T )i =

Dx x(t1 ) . . . x(tn )

(8.10)

xT


 Z

mx 2
+ V (x)
exp de
2
where e stands for Euclidean (Wick-rotated) time. Now,
X
|xT (T )i = e(T )H |xT (0)i =
|ni hn|xT i eEn T
n

where |xT (0)i is also written |xT i, and is a Schrodinger vector, which at
t = 0 coincides with the corresponding Heisenberg vector. When T is large,
only the ground state contributes. Then, eq. (8.10) becomes
= hxT (0)|0i h0| T(x(t1 ) . . . x(tn )) |0i h0|xT (0)i e2E0 T
Dividing by hxT (T )|xT (T )i, like in the Gell-Mann-Low formula, removes
the outer parts, leaving
Z xT
h0| T(x(t1 ) . . . x(tn )) |0i = lim ZT
Dx x(t1 ) . . . x(tn )eS
(8.11)
T

xT

110

CHAPTER 8. PATH INTEGRAL FORMULATION OF QFT

where ZT =

R xT
xT

Dx eS and S =

de ( m2x + V (x)).

Remarks
Note the similarity with the thermal average in statistical mechanics
with the Boltzmann weight eH .
x is in general not a fixed value; taking the Gell-Mann-Low quotient, the final result should not depend on it.
The Feynman-Kac formula: limT log hxT (T )|xT (T )i /(2T ) =
E0 . Exercise: derive this result.

8.2

Path integrals in QFT

8.2.1

Framework

Consider real scalar field theory, with L = 12 V (), where V ()


2
could for example be m2 2 + 4 4 . Now, (~x, t) substitutes xi (t) as we go
to infinitely many degrees of freedom. The following correspondences hold:
QM

QFT

Xi (t) (Heisenberg picture)

(~x, t) Heisenberg operator

|xi (t)i state vectors with


Xi (t) |xi (t)i = xi (t) |xi (t)i

Fock space (-eigenvector


states; coherent states)

boundary conditions

vacuum state |0i for t

|0i is unique only without outer fields (currents). With outer fields (or
currents), as we have seen before in the generating functional Z(j), we
have an extra term Lj in the Lagrangian density: Lj = j(x)(x), where
j 0 for large ~x and t. This implies that |t i 6= |t i and |0iout 6=
|0iin .
Going from xi to , our path integral will run over fields , and momentum fields , where the latter are given by
(x) =

L
= 0 = t
(0 )

(8.12)

Now, the - and -integrals become:


Z
Z Y
Z Y
D lim
d(~xi , j ) = lim
dk
i,j

8.2. PATH INTEGRALS IN QFT


Z
D lim

111
Z Y

dk

where the index k represents aR 4-dimensional lattice (the limit  0 will be


taken in the end). Of these, D can be performed like in the QM case,
and starting at the end again, we obtain:
(8.13)
h0| T((x1 ) . . . (xn )) |0i =


 R 4 1
R
D exp i d x 2 V ((x)) (x1 ) . . . (xn )
R
D exp(. . . )
(~ has been set equal to 1 again).
Now, we can sum over all fields with (~x, t) 0 as |~x|, |t| . (In
the quotient, this condition should drop out; strictly speaking, one has to
calculate the Fock space vacuum in x-space. See the exercises.) The integrals
are, like in QM, well defined after Wick rotation (x0 ix4 substitution
and 90-degree rotation in the x4 -plane) to imaginary time; another option is
inserting a term 21 (m2 i)2 (x) in the potential. The Wick rotation results
in:
V ((x)) V ((xE ))
Z
Z
d4 x i d4 xE

(8.14)
(8.15)

Figure 8.3: Wick rotation

Eq. (8.13) is very similar to the Gell-Mann-Low formula. It can also be


written with the help of a source j(x),
Z
Z(j) =

 Z

4
D exp i d x(L + j)

(8.16)

112

CHAPTER 8. PATH INTEGRAL FORMULATION OF QFT

which gives


1
n Z(j)

h0| T((x1 ) . . . (xn )) |0i =
n
(i) Z(0) j(x1 ) . . . j(xn ) j=0

(8.17)

where the Rvariational derivative provides factors i(xi ). Note: the normalization of D is often chosen to give Z(0) = 1.

8.2.2

QFT path integral calculations

Deriving eq. (8.13) in a discretized version, like we did for the QM case,
involves a multi-component version of the standard Gaussian integral, eq.
(8.8):


N Z
Y
1
I(A, b) =
(8.18)
di exp i Aik k + bi i
2

i=1

We diagonalize A by an orthogonal transformation O, whose Jacobi-determinant


is 1. Then, we have
0

I(a, b ) =

N Z
Y
i=1

d0i exp

1
0i aii 0i + b0i 0i
2

Completing the square, we rewrite 21 0i ai 0i + b0i 0i into 21 ai (0i


1 0 1 0
2 bi ai bi , which gives
0

I(a, b ) =

N
Y
i=1

2
exp
ai

1 01 0
b b
2 i ai i

N/2

= (2)

1
exp
(det A)1/2

b0i 2
ai )

1 T 1
b A b
2

where the second step is the result of rotating back. So, in the end, we are
left with
(2)N/2
I(A, b) =
exp
(det A)1/2

1 T 1
b A b
2


= I(A, 0) exp

1 T 1
b A b
2


(8.19)

N/2

(2)
= I(A, 0).
Observe that (det
A)1/2
Differentiation with respect to bi produces vacuum expectation values:



I(A, b)
= I(A, 0)
(A1 )ik
(8.20)
|
{z }
bi bk
b=0
2-point function

Let us apply this to the Z(j) from above, at first without interaction:
L = +

m2 2

(8.21)

8.2. PATH INTEGRALS IN QFT

113

Note that this is the Euclidean, i.e. rotated, form. With this L,
 Z

1
E
E
0 4 E 4 0E
2
2 1
Z0 (j) = Z0 (0) exp
j(x)(E + m )xx0 j(x )d x d x
2

(8.22)

For comparison: in the Minkowski metric, it looks like this:




Z
1
1
0 4
4 0
2
2
Z0 (j) =Z0 (0) exp
j(x)( + m i)xx0 j(x )d xd x =
2i


Z
1
Z0 (0) exp
j(x)DF (x x0 )j(x0 )d4 xd4 x0
2
In the Euclidean metric, the calculation is easier due to the absence of the
i-factors. It is easy to check that

1
2 Z(j)
= DF (x1 x2 )
h0| T((x1 )(x2 )) |0i = 2
i j(x1 )j(x2 ) Z(0)
Remarks
This can also be performed in the Fourier transformed form:
Z

1
d4 p  e
E
2
2 e
e
e
e
e
S0 =
(p)(p
+
m
)
(p)

j(p)
(p)

j(p)
(p)
2
(2)4
e
f (p), since e
with e
j(p) = j (p) and (p)
=
j and are real. Redefining
e0 (p) + (p2 + m2 )1e
e
(p)
=
j(p)
(i.e., completing the square) gives
Z0E (j)

Z0E (0) exp

1
2

d4 p e
j(p)e
j(p)
4
2
(2) p + m2

With

Z
1
W0 (j) =
dxdx0 j(x)(iDF )j(x0 )
2
we obtain the following relation for Z and W :


Z0 (j)
E
= eiW0 (j)
= eW (j)
Z0 (0)
where the factor iDF generates the connected 2-point functions; this
will be generalized later on. This relation resembles the one between
the partition function and the free energy in thermodynamics. Note,
by the way, that Z and W have been exchanged in the text by Ramond
(see literature list).

114

8.2.3

CHAPTER 8. PATH INTEGRAL FORMULATION OF QFT

Perturbation theory

The perturbative approach to path integrals will, again, be introduced in


the Euclidean notation. We will be dealing with a potential V (), which
can be written as
 

VE () VE
(8.23)
j
where the differentiation is acting on the generating functional Z0E (j). (For

4

4!
example, 4!
.) Let us start with
j(x)


n Z(j)
1

h0| T((x1 ) . . . (xn )) |0i =
Z(0) j(x1 ) . . . j(xn ) j=0

(8.24)

Now, expand the exponential in the path integral for Z(j) in powers of V ()

and act on Z0 (j) with V ( j


) as discussed before. Note that also the j(x
i)

of the outer and vertex pairwise


in eq. (8.24) act on Z0 (j). The j
R
1
remove the j-legs of exp( 2 jDF j). Division by Z(0) eliminates the vacuum
graphs (n.b.: Z0 (0) is not the same as Z(0)). For example, a vertex with
4
V = 4!
requires 4 propagators (permutation gives the factor of 4!). Also,
1
several vertices, each with their factor of n!
, can be permuted and require
a combinatorial factor. From all this, the usual rules for Feynman graphs
emerge again:

(i) only graphs with outer lines (no vacuum graphs)


(ii) DF for inner and outer propagators
(iii) and

d4 y for each vertex

(iv) combinatorial factors

Exercise: complex scalar fields:


Z(j, j) =

D exp( A + V () j j)

Here, is a complex vector (with index x). Rewriting it as = (1 +


i2 )/ 2 reduces the problem to two real fields 1,2 . D = D1 D2 , or,
more elegantly, DD with independent , . A = E2 + m2 in case of
the free complex Klein-Gordom field, and Z0 (j, j) = exp jA1 j). Derive

2
the Feynman rules for V () = ( 4) .

8.2. PATH INTEGRALS IN QFT

115

Recapitulation
In the path integral formulation of QFT, vacuum expectation values of time
ordered products of operators are calculated with the following formula:


1
n Z(j)

(8.25)
h0| T((x1 ) . . . (xn )) |0i =
Z(0) j(x1 ) . . . j(xn ) j=0
where

V ( j
)

Z(j) = e

Z0 (j) and
 Z

1
0
0
0
Z0 (j) = Z0 (0) exp
dxdx j(x)DF (x x )j(x ))
2

(8.26)

Z0 (0) drops out in eq. (8.25).


Vacuum diagrams are cancelled by the division by Z(0).
The Feynman rules are the same as those obtained in the framework
of canonical quantization.

116

CHAPTER 8. PATH INTEGRAL FORMULATION OF QFT

Chapter 9

Classification of finite
representations of the
Lorentz group
9.1

Classification of Lorentz transformations

The Lorentz transformations, like the rotations, form a non-commutative,


also called non-Abelian, group, which, however, is non-compact. They are
represented by matrices with the following property:
T g = g

(9.1)

The group operation is defined by executing the transformations sequentially:


x00 = 2 x0 = 2 1 x
T T g2 1 = g
| 1{z 2}

(2 1 )T

Classification
About Lorentz transformations, we know the following:
!

det(T g) = (det )2 det g = det g

(9.2)

det = 1

(9.3)

T0 g 0 = g00

(9.4)

So,

Furthermore, we know:

117

118

CHAPTER 9. LORENTZ GROUP

In orthogonal coordinates, (00 )2

k 2
k ( 0 )

= 1, so

|00 | 1

(9.5)

So, in total, we have 2 2 = 4 possibilities:


det = 1: L+
det = 1: L
00 1: L
00 1: L
L contains a space inversion, represented by the parity operator P; L
contains a time reversal, represented by the time inversion operator T:

1
0
1
0

1
1
;

P=
T=
(9.6)

1
1
0
1
0
1
L+ contains both time and space inversion, whereas L+ has neither. This
last group is called the proper orthochronous Lorentz group, which is a subgroup of the Lorentz group.
Now, consider infinitesimal proper orthochronous Lorentz transformations (i.e. continously connected to 1):
= 1 + M

(9.7)

M is the generator of these transformations. Applying eq. (9.2) to this


, we have (1 + M T )g(1 + M ) = g. Taking this to order O() gives
M T g + gM = 0, or M T = gM g 1 (remember, g 1 = g). In orthogonal
coordinates,
Mki = gii Mik gkk

(9.8)

We will make use of the following vector sets of matrices:


~ := (A(2,3) , A(3,1) , A(1,2) )
R
~ := (A(1,0) , A(2,0) , A(3,0) )
B

antisymmetric
symmetric

(9.9)
(9.10)

with

A(i,k) :=

(M(i,k) )ik = (M(i,k) )ki = 1


zero otherwise

(9.11)

9.1. CLASSIFICATION OF LORENTZ TRANSFORMATIONS

119

The only nonzero entries of M(,) are (M() ) and (M() ) . This means
that, for example, (A(2,3) )kl = (R1 )kl = 1kl . The Ri are the generators
of rotations, whereas the Bi generate boosts. As an exercise, check the
following relations:
[Ri , Rj ] = ijk Rk
[Bi , Bj ] = ijk Rk

(9.12)

[Ri , Bj ] = ijk Bk
~ transforms like a vector under
Note that the last commutator tells us that B
rotations.
Boosts
A short reminder: boosts can be considered rotations with an imaginary
angle. Consider, for example, a boost in the (0, 1)-plane:

cosh u sinh u
0
1 u
0
sinh u cosh u
u 1

infin.

1
1
0
1
0
1

where tanh u = = v/c. Boosting from a particles restframe to one that


is moving with velocity v would look like this (remember, p20 = E 2 /c2 =
p~2 + m2 c2 ):

p=

mc
~0

boost
p=

p~
v
=
;
c
E/c



mc

=
~0

 2 2 1/2
m c
= (1 2 )1/2 =
|
{z
}
p20

E/c
p~

cosh u

which gives back the usual relation for boosts. Exercise: check the commutation relations.
Note: an arbitrary Lorentz transformation, transforming p p0 , can
be represented as
= Bp0 DBp1

with boost

Bp1 : p p(0)
Bp0 : p(0) p0

(9.13)

where D is the so-called small group; in this case, it consists of the rotations.
This is called a Wigner rotation. Later on, we will come back to this, when
discussing induced representations.

120

CHAPTER 9. LORENTZ GROUP


The commutation relations (9.12) can be reformulated by defining
1
Jl = i(Rl + iBl )
2
1
Kl = i(Rl iBl )
2

(9.14)
(9.15)

which are Hermitean. (9.12) now become:


[Jk , Jl ] = iklm Jm
[Kk , Kl ] = iklm Jm

(9.16)

[Jk , Kl ] = 0
These are the independent commutators of two sets of generators of the
rotation group SO(3) and its covering group SU(2), respectively.
Now, representations are characterized by
~ 2 ) = 1 R(,) M (,)
R2 B 2 = 2(J~2 + K
2
1
~ B
~ = (J~2 K
~ 2 ) =  M(,) M(,)
R
8

(9.17)
(9.18)

This makes it easy to write irreducible, finite-dimensional representations of


the proper orthochronous Lorentz group, such as:
irreducible representations of infinitesimal rotations, which are classified by (j); j = 0, 12 , 1, . . .
irreducible representations of infinitesimal L+ -transformations, classified by (j, j 0 ); j, j 0 = 0, 21 , 1, . . .
j is the Casimir operator, whose eigenvalues, which classify the representation, are R2 and N 2 , respectively. However, infinitesimal boosts N J K
generate finite boosts by enN (sic, without i in the power), which is not
represented unitarily. However, the action of boosts in representation space
is not the same as a transformation in QM Hilbert spacae, as we will see
later on.
Remark
Space inversion (parity) P exchanges J and K:

~ axial vector
PK = JP
R
~
PJ = KP
N polar vector
So, V jj V j j are representations of L+ +L . The first non-trivial representation is ( 21 , 0)(0, 21 ), which we will meet again later on, as four-dimensional
Dirac spinor.
0

GROUP
9.2. POINCARE

9.2

121

Poincar
e group

We want a general framework for field equations, which is provided by the


Poincare group, also called the inhomogeneous Lorentz transformations. It
is the analogue of the Galilei group; its generators are P , for translations
and M , for Lorentz transformations. These operators have the following
commutation relations:
[M , P ] = i(g P g P )
[P , P ] = 0

(9.19)

[M , M ] = i(g M g M + g M g M )
In a more elegant representation, one uses the Pauli-Ljubanski vector:
1
W =  M P
2

(9.20)

It has the following properties:


W P = 0
~ = p0 R
~ p~ N
~
W
~
W0 = p~ R
In the restframe of a particle, where p~ = 0 and p0 = m, the latter gives
W0 = 0

and

~ = mR
~
W

~ = ~S, the spin.


where, in the absence of any momentum, R
The Pauli-Ljubanski vector commutes as follows:
[M , W ] = i(g W g W )
[W , P ] = 0

(9.21)

[W , W ] = i W P

Let us define
P 2 = P P

and

W = W W

(and note that W S2 for m2 6= 0, p~ = 0). Since P 2 and W commute with


all generators, they allow for a classification of relativistic particles as unitary representations of the Poincare group. The complete set of commuting
operators consists of:
P 2 , W , P~ , and W3 or

~
~
p~ W
p~ R
=
, the helicity of the particle
p0 |~
p|
|~
p|

In short, the classification is as follows:

122

CHAPTER 9. LORENTZ GROUP

(i) massive particles: P 2 = m2 ; in the rest frame of such particles, one


~ S
~
has p~ = 0, W
(ii) massless particles: P 2 = 0, so also W = 0. Since we cannot go to
the restframe of such a particle, we go to, for example, a frame where
(0)
P = (1, 1, 0, 0).
This leads to
W = P

with =

~ p~
R
p0
|{z}
|~
p|

Now, the commutation relations give


1
= 0, , 1, . . .
2
which correspond to the 1-dimensional (naturally irreducible) representations. For each case, there is a so-called small group, which is simply the
group of transformations that leaves the standard 4-momenta invariant. In
case (i), this group consists of the rotations, in case (ii), the 2-dimensional
Euclidean group (rotations and translations in two dimensions) in the plane
perpendicular to the helicity.
Remark
Irreducible representations of the small group induce irreducible, unitary
representations U of the (noncompact) Lorentz group, which in this case
are infinitely-dimensional, unlike before. Consider a normalized state with
momentum p and spin :

0 0

p0
p , |p, = |N (p)|2 0 3 (~
p0 p~)
k0
We obtain this state by a boost:
|p, i :=N (p)U(L(p)) |k, i

(9.22)

Now, let a unitary representation of a general Lorentz transformation act


on it:
U() |p, i =N (p)U(L(p)) |k, i =
N (p)U(L(p))U(L1 (p)L(p)) |k, i
The second U in the last line sends k to k, but is not necessarily 1, and is
hence an element of the small group (in e.g. the massive case, a rotation).
Let us call it


D0 (W (, p)) k, 0 := U(L1 (p)L(p)) |k, i
(9.23)

GROUP
9.2. POINCARE

123

This allows us to rewrite eq. (9.22) as


U() |p, i =



N (p)
D0 (W (, p)) p, 0
N (p)

which is the induced infinitely-dimensional unitary representation of the


Lorentz group. For further reading, see Wu-Ki Tung, Group Theory in
Physics.
Remarks
The finite-dimensional representations of the Lorentz group are not
unitary. However, field operators (~x, t) do not require direct probability amplitude interpretation, so this is not necessary. We will hear
more about this later, in connection with the Dirac equation.
Remember:
~

eit+k~x = ei(Et~p~x)/~ = eipx/~


where the exponent Et p~ ~x =
invariant.

E
c ct

p~ ~x = p x = p x is Lorentz

A Poincare transformation x0 = x + a can be written as a matrix




a
0 1
Exercise: show this.

124

CHAPTER 9. LORENTZ GROUP

Chapter 10

The Dirac equation and its


quantization
10.1

Spinor representation of the Lorentz group

Remember from last chapter:


i
Jl = (Rl + iBl )
2
i
Kl = (Rl iBl )
2
with
[Jk , Jl ] = iklm Jm
[Kk , Kl ] = iklm Km
[Jl , Km ] = 0
That is, we have a representation of the Lie algebra of the Lorentz group as
the sum of two Lie algebras of the rotation group SO(3). Keep in mind the
distinction between the rotation group SO(3) and its Lie algebra so(3). The
Lie algebra is generally notated by lower case so, although capitals (SO)are
often used as well.
Recall also the Pauli matrices:
1

0 1
1 0

0 i
i 0

1 0
0 1


(10.1)

We have two spinor representations, ( 21 , 0) and (0, 12 ), which can be written


125

126

CHAPTER 10. DIRAC EQUATION

with the help of the Pauli matrices, or with the 2 2 versions of J and K:

l
( 12 ,0) Jl +Kl
Rl
= i = i 2
1
( 2 , 0)
1
B ( 2 ,0) = Jl Kl = l
l
1
2
( (0, 1 )
l
Rl 2 = i 2
(0, 21 )
l
(0, 1 )
Bl 2 = 2
Going from ( 12 , 0) to (0, 12 ), one exchanges the roles of J and K; this is
what the space inversion operator P does, as discussed in last chapter. So,
( 21 , 0) + (0, 21 ) is an irreducible spinor representation, with


( 12 ,0)+(0, 21 )
i l /2
0
Rl
=
and
0
i l /2


( 21 ,0)+(0, 21 )
l /2
0
=
(10.2)
Bl
0
l /2
General infinitesimal Lorentz transformations are given by
1
= 1 + M (,)
2

(10.3)

(For M (,) , see last chapter: M (,) = M (,) ; the only non-zero entries of
M (,) are (M (,) ) and (M (,) ) .) Here, M (,) is represented by 4 4matrices in spinor space, which must fulfill the commutation relations
of the generators of the Lorentz group. If we define the Dirac -matrices
by their anticommutator
{ , } := + = 2g

(10.4)

1
= ( )
4

(10.5)

then

satisfies the commutation relations of the generators of the Lorentz group.


Exercise: check this.
The choice of is not uniquely determined by this requirement, but
Pauli has shown all choices to be equivalent. The Pauli lemma says that,
taking one set of , e.g.




0 1
0
i
0
i
=
and
=
(10.6)
1 0
i 0

one can transform to another set 0 by some transformation S:


S1
0 = S
(Notice that the identity matrices in 0 are 2-dimensional.)

(10.7)

10.2. SPINOR FIELDS, DIRAC EQUATION

127

Remark
The transform like a vector under Lorentz transformations:
[ , ] = ( g g )

10.2

(10.8)

Spinor fields, Dirac equation

A Dirac spinor transforms as follows:


S = 0

(10.9)

These transformations are generated by : infinitesimal ones are given


by
1
S = 1 +
2
with infinitesimal ; finite ones come with the full exponential:


1
S = exp

2

(10.10)

with finite . A Dirac spinor field by definition transforms as follows:


0 (x0 ) = S () (x)

(10.11)

where x0 = x. This begs the question what the corresponding Lorentz


covariant field equation is. To find the answer, begin with remembering
that transforms like a Lorentz vector:
S 1 S = 0 0

(10.12)

and conclude that therefore, if (x) is a solution, 0 (x) must also be one.
The sought-for equation, know as the Dirac equation, is
(i m)(x) = 0

(10.13)

or, in components:

(i
m ) (x) = 0

For the proof, we will use the following notation:


=

and

0 =

x0

128

CHAPTER 10. DIRAC EQUATION

The proof goes as follows: assuming (x0 ) is a solution, and using eq.
(10.12), we have
0

(i 0 m)0 (x0 ) =(i 0 m)S(1 x0 ) =


0

(i (1 )0 m)S(x) =
S(i m)(x) = 0
In the second step, we have rewritten 0 as (1 )0 , and in the third
0
one, we have used eq. (10.12) in the form (1 )0 S = S .

10.3

Construction of the representation matrices


in spinor space

In section 10.1, we started the explicit construction of the representation


matrices in spinor space. Let us finish this:

10.3.1

Rotations

Consider the following mapping:



~x x
= ~ ~x =

x3
x1 ix2
x1 + ix2
x3


(10.14)

Notice that x
is Hermitean. Further, check that
det x
= ~x2
and
1
xi = tr( i x
)
2

(10.15)

(for the last one, use 21 tr i k = ik ). Now, investigate a rotation or inversion


(i.e., an element from the orthogonal group) D:
~x0 = D~x
with, by definition, ~x02 = ~x2 and D D = 1. Consider a unitary representation
x
0 = U
xU

(10.16)

with UU = U U = 1. Notice that, due to the unitarity of U, we have


det x
0 = det x

or

~x02 = ~x2

(10.17)

In other words, U represents a rotation or inversion. Without loss of generality, we can choose det U = 1; this means that
U SU(2)

10.3. REPRESENTATION MATRICES IN SPINOR SPACE

129

where SU(2) stands for the set of special, unitary, 2 2-matrices, special
meaning that its determinant is 1. The general form of U is
U = ei~~/2

(10.18)

which is the general rotation operator of QM acting on spin- 12 objects. Infinitesimal transformations are given by


~
0
x
=
x + i
~ , ~ ~x =
2
 
i
x
+
2i(~
~x) ~
(10.19)
2

10.3.2

Lorentz transformations

For Lorentz transformations, the procedure is very similar. We use the


mapping
 0

x x3 x1 + ix2
xx
= 0 x0 ~ ~x =
(10.20)
x1 ix2 x0 + x3
which again is Hermitean. We also have the 4-dimensional analogues of eqs.
(10.15):
det x
= (x0 )2 ~x2 = x2

and

1
x = tr( x
)
2

(10.21)

Now, let us consider the transformation


x
0 = M x
M

(10.22)

with det x
0 = det x
, and hence | det M |2 = 1. Here, M is some arbitrary
matrix, for which we can, without loss of generality, assume det M = 1.
This implies
M SL(2, C)

(10.23)

i.e., M is a special, linear, 2 2-matrix with complex entries. This means it


has three complex parameters (or six real ones) to be fixed:



M=
with
= 1

Of course, SU(2) SL(2, C).
M acts in the representation space, which in this case is a 2-dimensional
spinor space: the ( 21 , 0) representation discussed before. In an abstract
vector basis, its action is
0 = M

130

CHAPTER 10. DIRAC EQUATION

In the basis


1
0

  
0
,
1

this becomes
s

M =

1/2
X
s0 =1/2

10.3.3

s (s M s )
| {z }
Ms0 s

Boosts

In order to be able to reformulate the Dirac equation in momentum space,


we will need a representation for boosts. Boosting the momentum vector p




m boost p0
,

~0
p~
represented by
p = m1 p0 = p0 1 p~ ~ = M pM ,
is achieved with
M = Lp~ =

p0 + m ~ p~ ! +
= Lp~
(2m(m + p0 ))1/2

(10.24)

(As a link to section 10.1: M = Lp~ is just e(~/2)(~p/|~p|)u with tanh u =


For L+
p
~ , chapter 9.) To prove this, first show that
0
Lp~ m1 L+
p~
p
~ = p 1~

|~
p|
.
p0

and

Lp~ ~ p~ L+
p~
p
~ =~
Boosting in opposite direction is done by means of L1
p . Using
p
~ = L~
(a0 1 + ~ ~a)(a0 1 ~ ~a) = (a0 )2 ~a2 , we get
(p0 1 + ~ p~)Lp~ = mL1
p
~
(p0 1 ~ p~)L1
~
p
~ = mLp

(10.25)

Inverting space by means of the parity operator P sends p~ to ~


p and
s
Lp~ s L~p s = L1
p
~

Defining
usp~ =

Lp~ s
s
L1
p
~


=


(10.26)

10.3. REPRESENTATION MATRICES IN SPINOR SPACE

131

we can rewrite eqs. (10.25) as


(p0 + ~ p~) = m
(p0 ~ p~) = m

(10.27)

Substituting eqs. (10.27) into each other gives


(p0 ~ p~)(p0 + ~ p~) = m2
(p0 + ~ p~)(p0 ~ p~) = m2
which leads us to
(p0 )2 p~2 = m2

(10.28)

which is the mass shell condition. One could, letting go of some mathematical rigour, interpret eqs. (10.27) as the square root of the mass shell
condition.
Remarks
If M is a spinor representation of the Lorentz group, so are (M )1 , M =
(M )T and (M T )1 . However, since


0 1
1
T
M = 2 M 2
with
2 = i
= 21 = 2 , (10.29)
1 0
M is equivalent to (M T )1 , and M to (M )1 .
Exercise: check eq. (10.29). Use the general ansatz
M = a1 + ~b ~ ,

a2 b2 = 1

The actions of the four representations are as follows:


M acts on
(M T )1 acts on
(M )1 acts on
M acts on
Letting 2 act on or raises or lowers the index; the point indicates
which representation is used. The pointed representation M corresponds
to ( 12 , 0), whereas the unpointed (M )1 corresponds to (0, 12 ). The spinor
usp consists of a left-handed and a right-handed component (see below
for more on handedness):
 

left-handed
s
up =

right-handed

132

CHAPTER 10. DIRAC EQUATION

Remark

The vector x
transforms as x : the index transforms with M , and the
with (M )T = M .

10.4

Field equation

10.4.1

Derivation

Now, let us derive the actual field equation. We know that




Lp~ s
usp~ eipx =
L~p s
should be a solution of a (free) field equation. From eqs. (10.27), we obtain





~ p~
0
0 1
0
s ipx
m
usp~ eipx = 0 (10.30)
p 14
up~ e
0
~ p~
1 0
Let us do some renaming, to make it look a little simpler:




~ p~
0
0 1

~ p~ :=
and
:=
0
~ p~
1 0
Observe that 2 = 1. Furthermore, we will name
= 0

and

~
= ~

(Exercise: check that these satisfy the requirements for the Dirac -matrices.)
Now, multiply eq. (10.30) with :
( 0 p0 ~ p~ m) usp~ eipx = 0
| {z }
(x)

and rewrite it into a differential equation, with p = i/x :


(i m)(x) = 0

(10.31)

This is the Dirac equation. It also has other guises: going back to momentum
space, it looks like
( p m) = 0

In the Feynman slash notation, where = / (and similarly p


/ = p ),
it takes the form

(i/ m) = 0

10.4. FIELD EQUATION

10.4.2

133

Choosing the -matrices

In the derivation above, 0 was Hermitean, whereas the i were anti-Hermitean.


Since we know { , } = 2g , as the should be equivalent to by
the Pauli lemma,
= 1

(10.32)

(here with as the 0 from section 10.4.1). Taking the Hermitean conjugate
of eq. (10.32), we get
= 1
Sandwiching this between and 1 gives
= 1
which, combined with eq. (10.32), allows us to conclude that
=

(10.33)

Different representations of the Dirac algebra


The representation used above is called the chiral or Weyl representation.
It has




0 1
0
i
W 0
W i
i
(10.34)
==
,
= =
1 0
i 0
(Itzykson and Zuber use 0 0 .) We can define a fifth -matrix:


1 0
5
0 1 2 3
= 5 = i =
(10.35)
0 1
This has the nice property that (1 5 )/2 projects onto the upper and lower
components of spinors. Representing as (L R )T , it selects the left- or
right-handed component of :





1 + 5 1 5
1 + 5 1 5
L
!
=
+
=
+
=
R
2
2
2
2

 
 

L
0
L
+
=
=
0
R
R
Also often used, mainly in the non-relativistic limit of the Dirac equation,
is the Dirac-Pauli representation:




1 0
0
i
D 0
D i
i
==
,
= =
(10.36)
0 1
i 0

134

CHAPTER 10. DIRAC EQUATION

In this representation, 5 is not diagonal anymore:




0 1
5
= 5 =
1 0

(10.37)

The conversion from the Weyl to the Dirac-Pauli representation is as follows:






1
1 1 W 1
1 1
D

(10.38)
=
2 1 1
2 1 1
Dirac spinors are obtained from their Weyl counterparts by means of the
same matrix:
 

1
1 1

(10.39)

2 1 1
Remarks
The Dirac equation is a differential equation of first order in time, just
like the Schr
odinger equation:
!
~

~
i0 = m +

i
Notice that both terms between the brackets are Hermitean.
Weyl spinors are very common in GUTs, neutrino physics and supersymmetry. In other fields, the Dirac notation dominates.

10.4.3

Relativistic covariance of the Dirac equation

The fact that the Dirac equation is relativistically covariant is rather obvious, since we have been talking about Lorentz covariant objects all the time.
However, let us check it explicitly. Supposing we have a solution (x), we
want to show that its transformed form
0 (x0 ) = S () (x)

(10.40)

is also a solution. In the above equation, as usual, x0 = x and S is a


representation of the Lorentz group ( 12 , 0) + (0, 21 ). If is written in terms
of Weyl spinors, S becomes


M
0
S=
0 (M )1
The proof that 0 (x) also solves the Dirac equation was given at the end of
section 10.2; it holds under the condition that
0

(1 )0 S = S

10.5. COMPLETE SOLUTION OF THE DIRAC EQUATION

135

or
0

S 1 S =

(10.41)

This corresponds to the transformation property of


if we write S in the
above form.
Remarks
Remember that in its infinitesimal form, this S is just the 1 + 12
we have seen before.
~ and B
~ are:
The Dirac versions of R
~ D = i 5 0~ 2 ,
R
4

~ D = i 0~ 2
B
4

Space inversion S() = 0 exchanges with .

10.5

Complete solution of the Dirac equation

In deriving the Dirac equation, we began with writing down a solution.


However, this is not necessarily a unique one; let us therefore now invert the
procedure in order to find all solutions. So far, we have one:
(x) =

1
up~ eipx
(2)3

with positive p0 = p~2 + m2

(10.42)

Going to the momentum basis (by means of a Fourier transform), this becomes
( p m)up~ = 0
with
usp~

m1/2
=
(2m(m + p0 ))1/2

p0 + m ~ p~
p0 + m + ~ p~

(10.43)

The factor of m1/2 is put in by convention (rescaling usp~ is always possible).


Eq. (10.43) is written in the chiral basis; going to the Dirac-Pauli basis, we
will use D = B W B 1 with




1
1
1 1
1 1
1
B=
and
B =
(10.44)
2 1 1
2 1 1
In this basis, we have
Busp~

1
=
(m + p0 )1/2

p0 + m
~ p~

(10.45)

136

CHAPTER 10. DIRAC EQUATION

In analogy to the case of the Klein-Gordon equation (chapter 4), we


might expect a negative energy solution to exist. Let us make the ansatz
1
(x) =
vp~ eipx
(10.46)
(2)3
In momentum space, this would mean
( p m)vp~ = 0
Now we still need to find out what vp~ actually is. To that end, let us go to
another representation of the s. First, we define C:
C T C1 =

(10.47)

We know that this C must exist because of the Pauli lemma: if fulfill the
requirements, so do T and . It is given by
C = i 0 2

(10.48)

Exercise: check this. Now, we need only one further definition:


:=

(also u
sp~ = usp~ ) and
!
:= =

(10.49)

Now, we can simply obtain vp~s from the already known solution usp~ : we begin
with
(p m)usp~ = 0
and take the barred conjugate of this equation:
u
sp~ (
p m) = 0
Then, we take the transpose and multiply that with C:
C( T p m)C1 C u
sp~ T =
(C T pC1 m)C u
sp~ T =
(p m)C u
sp~ T = 0
From this, we can see that
vp~s = C
usp~ T

(10.50)

Exercise: check the following results:


0

u
sp~ usp~ = 2m ss
0
0
0
u
vp~s vp~s = usp~ T CC
sp~ T = 2m ss
0

u
sp~ 0 usp~ = 2p0 ss
vp~s 0 usp~ = 2p0 ss

=1
with CC

vu = 0
v 0 u = 0

(10.51)

10.6. LAGRANGIAN FORMALISM

137

Remarks
The notation us (p) is often used instead of usp~ .
We can define projectors to the positive and negative energy solutions:
p
/+m
+ (p) =

2m

2m

usp~ u
sp~

s=1,2

p
/+m
(p) =

vp~s vp~s

s=1,2

General solution
The most general solution to the Dirac equation is of course a linear combination of the two independent ones we have found so far:
Z
X 

d3 p
s
s
ipx
s
s
ipx
(x) =
(10.52)
a
(p)u
(p)e
+
b
(p)v
(p)e

(2)3 2p0
s=1,2

with a Dirac index . This is the analogue of the complex solution of the
Klein-Gordon equation.

10.6

Lagrangian formalism

Let us develop the Lagrangian formalism for the Dirac equation. We have
the following Lagrangian and Lagrange density:
Z

L = d4 xL;
L = (x)(i
m)(x)
(10.53)
(like we varied the Klein-Gordon LaVarying L with respect to and
grangian with respect to and before)
L
L

= (i m)(x) = 0

L
L
L

=

= (x)(i
m) = 0

( )

(10.54)

gives back the Dirac equation in two guises (the second one is the barred
one could distribute the
form). Notice that L is not symmetric in and ;

derivative
with respect to and symmetrically via partial integration in
R
S = d4 xL.
There is, of course, also a momentum field:
=

L
0 ) = i
= i(
(t )

(10.55)

138

CHAPTER 10. DIRAC EQUATION

The barred equivalent gives


L

) = = 0
(t
=
if it does not take the symmetrized form mentioned above. This result (
0) means that we have a constrained system, requiring special attention and
special tools, like the Dirac bracket. There exists extensive literature about
quantizing under constraints; here, we will not go into it.
The Hamiltonian density is given by

0 0 (x)(i

H = t L = i
p m)(x)
0

= i 0

(10.56)

where in the last step we have applied the Dirac equation, which comes
down to requiring the mass shell condition.
The Hamiltonian, after inserting eq. (10.52), finally takes the form
Z
H=

d3 xH(x) =

X
d3 p
0
p
{as (p)as (p) bs (p)bs (p)} (10.57)
(2)3 2p0
s

Exercise: check this (use eqs. (10.51)). Notice the minus sign in front
of the b-term: this is different from the Klein-Gordon case, and will be of
importance later on. Also note that the ordering of the as s and bs s within
their respective terms does not yet matter now, but will be important for
the quantization procedure.

10.6.1

Quantization

We begin the quantization in the usual way, by promoting the functions


as() and bs() to operators as() and bs() , accompanied by the usual prescription that Poisson brackets are to be replaced with commutators. Now,
we immediately run into trouble: this allows for limitlessly negative energy
solutions , because of the minus sign in the Hamiltonian. To avoid this, we
do not use commutators here, but take anticommutators instead:
0

{as (p), as (p0 )} = (2)3 2p0 3 (~


p p~0 ) ss
{bs (p), bs (p0 )} = (2)3 2p0 3 (~
p p~0 ) ss

(10.58)

{a, a} = {b, b} = {a , a } = {b , b }
= {a, b} = {a , b} = {a, b } = {a , b } = 0
This is equivalent to
(~x0 , t)} = 3 (~x ~x0 ) 0
{ (~x, t),

(10.59)

10.6. LAGRANGIAN FORMALISM

139

Of course, we could have obtained this directly, replacing the Poisson bracket
0 ) by an anticommutator. Using this result,
of and = i = i(
we get a more familiar-looking Hamiltonian:
Z
o
Xn
d3 p
s
s
s
s
0
a (p)a (p) + b (p)b (p) 1
(10.60)
p
H=
(2)3 2p0
s
from which we can conclude that the b-particles also have positive energy.
This is one of the highlights of early QFT.
Fock space
We have already discussed this in the context of non-relativistic QM, in
chapter 3. A general state |i is given by
Z
Z

h
X
1 X
d3 k1
d3 kn

|i = c0 +
.
.
.

(2)3 2kn0
n! s1 ...sn (2)3 2k10
n=1
i
cn (k1 , s1 , . . . , kn , sn )a (k1 , s1 ) . . . a (kn , sn ) |0i +
similar for b, b +

(10.61)

mixed terms
The fact that {a , a } = 0 forces a a = 0, so there can never be more than
one particle with momentum k and spin s; this is the Pauli principle. Due
to the complete antisymmetry in the a s, the coefficients cn can be taken
to be completely antisymmetric itself: any symmetric parts would vanish.
Notice, by the way, that the ordering of the as and bs is also important,
due to the presence of the mixed terms.

10.6.2

Charge conjugation

We want to show that a-particles and b-particles have opposite charge. To


this purpose, consider the coupling to the electromagnetic field, also called
minimal coupling:
D = + ieA
so

L = (x)(i ( + ieA ) m)(x)

(10.62)

Let us now introduce the charge conjugation:


C

T (x)
(x) C (x) = C

(10.63)

(This should correspond to exchanging and in the Klein-Gordon case.)


C is given by

C = i 0 2 = C

(10.64)

140

CHAPTER 10. DIRAC EQUATION

i.e., it is the one we have used in sec. 10.5, defined in eq. (10.47). The

C.
barred equivalent of eq. (10.63) is (x)
T C =
T
(x):
Let us further investigate C
T (x) = C 0T =
C
Z
o
d3 p X n s
0 T s
ipx
s
0 T s
ipx
a
(p)C(
)
u
(p)e
+
b
(p)C(
)
v
(p)e
(2)3 2p0 s
Using that v s (p) = C T us (p) and us (p) = C T v s (p), we see that this
simply comes down to exchanging the roles of as (p) and bs (p); in other words,
particles and antiparticles are exchanged. Now consider the interaction term
in the Lagrangian (eq. (10.62)): the Dirac equation becomes
(i ( + ieA m)(x) = 0
Taking the bar-conjugate of the equation gives

(i
( ieA m) = 0
Letting C act on this, we get
T = 0
(i T ( ieA m)C1 C
And in the end, we see that in the final form,
T = 0
(i T ( ieA m)C
the sign of the charge has changed indeed.

Chapter 11

Feyman rules for theories


with fermions
11.1

Bilinear Covariants

The Lagrange density is a Lorentz scalar, which has to be constructed to


include fermions. The algebra of Dirac matrices has 16 elements, which are
linearly independent 4x4 matrices:

1; ; =

i
[ , ] ; 5 = i 0 1 2 3 ; 5
2

(11.1)

(Check: there are 1 + 4 + 6 + 1 + 4 = 16 of them and they are linearly


independent.) Remember that a spinor transforms like
0 (x0 ) = S () (x)

(11.2)

Its barred conjugate transforms like


0 (x0 ) =
(x)S

with S = 0 S 0

(11.3)

From S 1 () S() = it follows that S = S 1 . Using this, we can


is a Lorentz scalar:
see that
0 (x0 )SS
0 (x0 ) =
(x) (x)
0 (x0 )0 (x0 ) =

(11.4)

Similarly, we obtain

0 (x0 ) 0 (x0 ) =
S
S = (x)

(x)

(11.5)

(x) is a Lorentz vector. More generally, if we denote the

That is, (x)


(i) (x) is a tensor with the indices

16 Dirac matrices by (i) , then (x)


is a symmetric, traceless
of the sandwiched Dirac matrix. Indeed,
tensor of rank 2.

141

142

CHAPTER 11. FERMION FEYNMAN RULES

With the transformation properties of 5 under proper Lorentz transformations and space inversion
S 1 () 5 S() = 5
S 1 () 5 S() = 5
(exercise: check this; remember that 5 =
5 is a pseudoscalar

(S() = 0 )
i

4!  )

we see that

5 is a pseudovector

This will be very important for constructing the coupling terms in the Lagrangian. For example, in the electromagnetic case we have
( + ieA ) m)
Lelmg = (i

(11.6)

(minimal coupling) with as a possible additional coupling


F
Lint =
Another possibility is Yukawa coupling of a Dirac field to scalars and pseu
doscalars, respectively. The latter one can, for example, be found in N N
coupling.

Lsc
int = f (x)(x)(x)

(11.7)

Lps.sc.
= f (x)
(x)(x)
int

(11.8)

Note that in both, may be real or complex valued.

11.2

LSZ reduction

To calculate S-matrix elements with fermions one has to apply the LSZ
reduction again. To this end, similar to the scalar case, one can use:
Z
s
ain (k) = d3 x
us (k) exp(ikx) 0 in (x)
Z
s
v s (k) exp(ikx) 0 in (x)
(11.9)
bin (k) = d3 x
The reduction formula for n incoming particles (u(ki )), n0 incoming antiparticles (
v (ki0 )), m outgoing particles (
u(qi )) and m0 outgoing antiparticles
0
(v(qi )) then reads
h0| bout (q10 ) . . . aout (q1 ) . . . ain (k1 ) . . . bin (k10 ) . . . |0i =
Z
1/2 (n+m)
1/2 (n0 +m0 )
(iZ2 )
(iZ2 )
d4 x1 . . . d4 y10
(11.10)
n X
o
exp
(k x + k 0 x0 q y q 0 y 0 ) u
(q1 )(i6y1 m) . . . v(k1 )(i6x01 m)


10 ) . . . (y1 )(x
1 ) . . . (x01 ) |0i (i
h0| T (y
6 x1 m)u(k1 ) . . . (i 6 y10 m)v(q10 )

11.3. FEYNMAN RULES

143

where the disconnected part has been left out like before. Dirac and spin
indices have been suppressed in this formula, to improve the legibility. Be
careful: anticommutations in the T-product will produce minus signs!
Remark
For photons, we have
1/2

h; k,  out|in i = (i)(Z3 )

d4 x exp(ikx) hout | j(x) |in i

with 2 A = j

(11.11)
(11.12)

Here, further reduction requires a redefinition of the T-product (called T product), adding distributions (6= 0 for x = y) in order to obtain covariant
expressions.

11.3

Feynman rules

Here we can be rather short:


(i) One derives the Gell-Mann-Low formula again, this time formulated
in fields.
in in and
(ii) The analogous Wick theorem for fermionic fields leads to contractions of the spinor fields. Since the Dirac theory is similar to that of
complex scalars, it is pretty obvious (exercise: check this) that only
the propagator

(y) |0i
h0| T (x)
is not zero
Note
Fermions can also be treated in the path integral formalism very elegantly.
We will come back to this in chapter 13.

11.3.1

The Dirac propagator

Let us start off by defining



(y) |0i
SF, (x y) := h0| T (x)

(11.13)

where and are spinor indices. It is often notated as SF (x, y), and fulfills
the equation
((i6 m)x ) (SF (x, y)) = i 4 (x y)

(11.14)

144

CHAPTER 11. FERMION FEYNMAN RULES

We can check that SF (x y) = (i6 + m)x DF (x y) is the solution with


Feynman boundary conditions using
(i6 m) (i6 + m) = 2 m2 = ( 2 + m2 )
The Fourier transform turns out to be
i(6p + m)
SF (p) = 2
p m2 + i

(11.15)

(11.16)

Of course, this propagator has direction, and the corresponding propagator


line has an arrow, since is a complex field.
Note
In the time ordering of the LSZ reduction formula for Dirac fermions commutations cause minus signs. Similarly, in the Wick formula, contraction
can only be done after commuting through the field in between and thereby
picking up minus signs. There are u, u
, v, v in the LSZ-formula corresponing
to ingoing and outgoing particles and ingoing and outgoing antiparticles.
The vacuum graphs again cancel. Inserting the vacuum expectation value
into the LSZ formula the outer propagators are cancelled,
which is also re
ferred to as truncated, and we are left with a Z factor. This will be
important for loop calculations later on.
Later, we will see that in the dressed propagator for fermions, the selfenergy insertions also have the Dirac structure:
iZ2 (6p + m)
i(6p + m0 )
+ SF (p, m0 )SF (p, m0 ) + = 2
+ rest
2
m + i
p m2 + i

p2

We then obtain the Feynman rules for S-matrix elements (in momentum
space):
For Yukawa theory
(i) Dirac propagator (eq. (11.16), instead of the scalar propagator before)
(ii) if or if 5 , for scalar and pseudoscalar vertices, respectively
(iii) statistical factors
(iv) vertex integration
(v) us for incoming particles, u
s for outgoing particles, v s for outgoing
s
antiparticles, v for incoming antiparticles

(vi) Z for outer scalar, Z2 for outer fermion fields


(vii) 1 for closed (inner) fermion lines (to be explained later)
Note: fermion propagator lines end in outer particles or are closed, but do
not cross each other.

11.3. FEYNMAN RULES

145

Figure 11.1: Yukawa scalar and QED vertices

For quantum electrodynamics


The rules will just be given here, in the Lorentz gauge ( A = 0). The
derivation will come later, in the context of the fermionic path integral.
The interaction Hamiltonian is

Hint = eQ(x)
(x)A (x)

(11.17)

(Q is the charge number, e.g. 1 for electrons). The rules are:


ig

(the m-term in the denominator disap(i) photon propagator: p2 +i


pears, since m = 0 for photons); Dirac propagator as before

(ii) ie Q for vertices


(iii) statistical factors
(iv) vertex integration
(v) us , u
s , v s , vs as before
(vi)

Z3 for outer photons

(vii) 1 for fermion loops

Note
Spin averaging over incoming particles and spin summation over outgoing
particles are often required when calculating cross sections. This gives, for
example,

146

CHAPTER 11. FERMION FEYNMAN RULES

1X s
0
0
u
(p2 )5 us (q2 )(
us (p2 )5 us (q2 )) =
2 0
s,s

1X s
0
0
u
(p2 )5 us (q2 )
us (q2 )5 us (p2 ) =
2 0
s,s


6q2 + m
1X
s
s
tr u (p2 )u (p2 )5
5 =
2 s
2m


1
6p2 + m 6q2 + m
tr
5
5 =
2
2m
2m
1
tr((6p2 + m)(6q2 + m)) = 2(p2 q2 + m2 )
2
(the factor of

11.4

1
2

comes from spin averaging).

Simple example in Yukawa theory

Let us consider nucleon scattering with pion-exchange as force mediator


as an example (between quotes because, of course, we know nowadays that
nucleons are composite objects). The interaction Hamiltonian density is
given by
5
Hint = f

(11.18)

N-N scattering

Figure 11.2: t-channel exchange


This is t-channel exchange: t = (p1 q1 )2 . The expression corresponding
to the diagram (see figure) is the following:
i

(p1 q1 m2 + i
(Z2 )2 (2)4 4 (p1 + p2 q1 q2 ) =

(if )2 u
(p1 ) 5 u(q1 )
u(p2 )5 u(q2 )

)2

iM(2)4 4 (p1 + p2 q1 q2 )

(11.19)

11.4. SIMPLE EXAMPLE IN YUKAWA THEORY

147

There is a second diagram for N-N scattering, where the two outgoing lines
are exchanged. According to the above recipe, this will give a minus sign;
this can also be derived from a careful evaluation of the LSZ-formula.

Figure 11.3: u-channel exchange

The Z-factors are not of great interest here; they will become important
in the discussion of renormalization. The propagator 1/(p1 q1 )2 m2 in
the t-channel gives, after Fourier transformation of the evaluated iM (the
amplitude), the Yukawa potential, which is well known in nuclear physics.
scattering
N-N

Figure 11.4: s-channel exchange


This is s-channel exchange: s = (q1 + p2 )2 . The diagram is the same,
except for the exchange of p1 and q2 . The corresponding expression is
i

(p1 q1 + m2 + i
(Z2 )2 (2)4 4 (p1 + p2 q1 q2 )

(if )2 v(p1 ) 5 u(q1 )


u(p2 )5 v(q2 )

)2

(11.20)

148

CHAPTER 11. FERMION FEYNMAN RULES

Another diagram that represents this type of scattering is shown below,


and has the following expression:
(if )2 u
(p1 )5 u(q1 )
v (p2 )5 v(q2 )

(p1 q1

)2

i
...
+ m2 + i

(11.21)

Notes
In Peskin & Schr
oder, in the spirit of time-ordered perturbation theory,
outer one-particle states are directly contracted with vertex-fields:
5
5 |q1 p2 i
hp1 q2 |
The required permutations give rise to extra minus signs, which include
signs related to the fermion statistics of the outer states. Altogether,
this gives the u
, u, v and v as above.
A closed fermion loop gives a minus sign, as mentioned above:
1 )(x1 )(x
2 )(x2 )) |0i
h0| T((x
2 )(x2 )(x
1)
(x1 )(x

Figure 11.5: Closed fermion loop

Feynman approach vs. time-ordered perturbation theory


In the Feynman approach the propagators are off-shell and one has 4momentum conservation at the vertices.

11.4. SIMPLE EXAMPLE IN YUKAWA THEORY

149

Figure 11.6: Cases () (left) and () (right)

In time-ordered perturbation theory, on the other hand, one has the


well-known energy denominators, like in QM, and, which is new in field
theory, 3-momentum conservation at the vertices. Of course the incoming
and outgoing particles have the same total energy, but in between, energy
is not conserved, as a consequence of the energy-time uncertainty relation.
The intermediate particles are on shell in this approach.
For tree-level diagrams, the relation between the two approaches can be
seen by decomposing the propagator:
p2

1
1
= 2
2
2
m + i
p0 p~ m2 + i

In the Feynman case, one rewrites this as


1

p
p

p0 p~2 + m2 + i p0 + p~2 + m2 i

1
2p0,+

In the time-ordered perturbation theory, one has the following denominators:


p
p
E10 + p~2 + m2 + E2 E1 E2 = p0 + p~2 + m2
()
p
p
0
0
0
E1 + p~2 + m2 + E2 E1 E2 = p0 + p~2 + m2
()
These are the same as the ones from the Feynman case.
For loop integrals, the Feynman approach has
Z
d4 k
1
1
4
2
2
2
(2) (p k) m + i k m2 + i

150

CHAPTER 11. FERMION FEYNMAN RULES

Time-ordered perturbation theory has


Z
1
d3 k
energy denominators
Here, the connection between the two approaches can be seen by doing
the dk 0 -integration using Cauchys theorem (the residue theorem), closing
the integration contour in one of the two halfplanes. One gets poles in k0 ,
yielding the required energy denominator in the residue.
For example, in 3 scalar theory, for the graph in figure 11.6, one has
Z
i
i
d4 k
=
(i)2
(2)4 k 2 m2 + i (p k)2 m2 + i
!
Z
4k
1
d
1
1
p
p

(i)2
4
0,+
(2)
k 0 ~k 2 + m2 + i k 0 + ~k 2 + m2 i 2k

1
1

q
q

0
0
2
2
0
0
2
2
~
~
(p k ) (~
p k) + m + i (p k ) + (~
p k) + m i
2(p0

1
k 0 )+

(This is a rough calculation, just to see the principle; we will not worry about
infinities here.) Closing the integration contour in the lower half plane and
applying the residue theorem gives
Z
1
1
d3 k
2
(i)
(2i) 0,+

4
0
(2)
2k
2(p k 0 )+

1
1

q
q

p
p
2
2
0
2
0
2
~
~
~
p k) + m
p + k + m + (~
p ~k)2 + m2
p k + m (~

1
1

q
q
+
p
p
0
2
2
2
0
2
2
2
~
~
~
~
p k + m + (~
p k) + m
p k + m + (~
p k) + m

Chapter 12

Quantum mechanical
interpretation of the Dirac
equation
In this chapter, we will investigate the relation between the Dirac and
Schrodinger equations in the non-relativistic limit, and study the FoldyWouthousen transformation.

12.1

Interpretation as Schr
odinger equation

If one reinserts the - and -matrices used in deriving the Dirac equation
(chapter 10), it takes the form
!
~

~
it (~x, t) = m +
(~x, t)
(12.1)
i
|
{z
}
H

This looks very similar to the Schr


odinger equation. It is Hermitean ( =

2
2
2
and
~ =
~ ) and H = m + p~ . Defining by
0
(~x, t) = =

(12.2)

one obtains a continuity equation:


) = ( )
+
= 0
(

(12.3)

where in the last step, we have used the Dirac equation: (i m) = 0,

= 0,
and its barred couterpart (i
) m
All in all, it looks like we could try to interpret the Dirac equation like a
Schrodinger equation without second quantization. However, there is one
important difference: the Dirac equation has negative energy solutions. The
151

152

CHAPTER 12. QM INTERPRETATION OF DIRAC EQUATION

Hamiltonian above is not bounded from below in any way; we would have
to put this in artifically, since the Schrodinger equation has only positive
energy solutions. This means we would assume
Z
d3 p X s
(x) =
a (p)us (p)eipx
(12.4)
2p0 (2)3 s
without the b-term we had in chapter 10. If we do this,
Z
Z
Z
0 = d 3 p
e (p)(p)
e
d3 x = d3 x

(12.5)

has a probability interpretation. Note that


e
(p)
=

X as (p)us (p)
((2)3 2p0 )1/2
s

(12.6)

We want to be able to calculate expectation values


Z
h| A |i = d3 x A
~ which acts on (x) like a multiplication by ~x,
like in QM. Assuming an X
we can calculate, by Ehrenfests theorem,
Z
Z
d
~ = i d3 x [H, X]
~
d3 x X
dt
Writing out the commutator in the integrand, we have
~ = i~
~ + m, X]
[i~

d
h~xi = h~
i
dt

so
(12.7)

i.e., acts like a velocity. Since


[H,
~ ] 6= 0
we know that this velocity changes. Its expectation value, assuming both
positive and negative energy solutions, is
Z

d3 p
p~ X  s 2
|a | + |bs |2 +
h~
i =
(12.8)
3
(2) 2p0 p0 s
Z
2
1 X n s0
d3 p
0
0
0
a (~
p)bs (~
p)
us (~
p)~
v s (~
p)e2pi0 x +
(2)3 2p0 2p0 0
s,s =1
o
0
0
s0
s
b (~
p)a (~
p)
v s (~
p)~
us (~
p)e2ip0 x


12.1. INTERPRETATION AS SCHRODINGER
EQUATION

153

The first term reproduces the usual group velocity p~/p0 ; the last term mixes
the positive and negative energy solutions (more properly said, it mixes the
a- and b-modes; the energies p0 are always positive). If
~ is a velocity, and
2
i = 1 as we have seen before, this must be interpreted as a zitterbewegung: a
movement with the velocity of light and changing direction. This movement
~ lead out of
arises since operators corresponding to interactions, e.g. X,
the space of positive frequencies. This leads to the question why radiative
transitions from positive to negative energies are impossible.

Figure 12.1: The Dirac sea, with occupied negative energy states

Here, we will introduce the concept of the Dirac sea (fig. 12.11 ). Dirac
postulated that all states with negative energy are occupied in the ground
state of the system. In this case, the Pauli principle forbids transition to
these states, which together form the Dirac sea. In this view, the excitation
of an electron with negative energy, Ee < me , due to electromagnetic radiation with E > 2me results in an electron with positive energy and a hole,
i.e. an unoccupied negative energy level. This hole behaves like a particle
with opposite charge, i.e. a positron, and therefore, the above process is
just that of pair creation. Pair annihilation occurs when an electron emits
electromagnetic radiation and jumps down into an existing hole:
e+ + e 2

(12.9)

The fact that there are two photons being produced is a consequence of
certain conservation laws we will return to at a later time. The relation
between positive and negative energy solutions is given by a kind of charge
conjugation. Be careful, however: in this picture, there are no anticommuting Fourier coefficients. It may at times pop up in old literature or in solid
state physics, but we will use the second quantized picture.
1
Taken from http://www.phys.ualberta.ca/ gingrich/phys512/latex2html/node63.html,
accessed 30-09-2007.

154

CHAPTER 12. QM INTERPRETATION OF DIRAC EQUATION

12.2

Non-relativistic limit

12.2.1

Pauli equation

In the non-relativistic limit, the Dirac-Pauli choice of -matrices is convenient. A short reminder:




1 0
0
i
D 0
D i
i
==
= =
0 1
i 0


1
(p0 + m)s
D s
sW
up = Sup =
(12.10)
~ p~s
(p0 + m)1/2
Note that this u is a factor m1/2 smaller than the last time we saw it (chapter
10); this is done to ensure that u u = 1, to preserve the analogy to QM.
In the non-relativistic limit,
p0 = m +

1 p~2
+ ...
2 2m

for |~
p|  m

(12.11)

In this limit, the upper components of u are much bigger than the lower
ones. This distinction between big and small components extends to the
covariant quantities:
=1

= 0

5 = 0 5

z }| {
0 = ( 0 )2 are big
and
i = 0 i are small
and

=1

z }| {
0i = 0 0 i is small

ik = 0 i k is big

As mentioned before, we write the Dirac equation as


!
~

~
(~x, t)
it (~x, t) = m +
i
We will call the Hamiltonian HD :
!
!
~
~

~
m
~ /i
m +
= HD
=
~
i
~ /i
m

(12.12)

Introducing electromagnetic interaction, we replace by D :


+ ieA = D
With the ansatz
imx0

=e

1
2


(12.13)

12.2. NON-RELATIVISTIC LIMIT

155

we obtain two coupled equations:


~ + ieA)
~ 2 e1
i0 1 = i~ (
~ + ieA)
~ 1 e2 2m2
i0 2 = i~ (

(12.14)

= E m  m and
We want stationary solutions with positive energy E
0
i
Ex
e
. This condition implies
|i0 2 |  2m2
|e|  m

and
(weak potential)

(12.15)
(12.16)

This gives for the small components 2 :


2 i~

~ + ieA)
~
(
1
2m

(12.17)

Inserting this into the equation for 1 gives


#
"
~ + ieA)~
~ (
~ + ieA)
~
~ (
e 1
i0 1
2m
For the vector product between the brackets, we have the following identity:
~ ~a~~b = ~a ~b + i~ (~a ~b)
(Exercise: check this.) Plugging this in gives
"
#
~ + ieA)
~ 2
(
e
~ e 1
i0 1
+
~ B
2m
2m
with
~
~ = B ~r
A
2

(12.18)

~ This is the coordinate gauge, also known as the Fockand constant B.


Schwinger gauge. In this simple case, writing out the square between the
brackets and rearranging terms leads us to the Pauli-equation:
#


~2

e
~

~ +2
~ e 1
+
L
B
i0 1
2m 2m
2
"

(12.19)

The factor 2 in front of ~ /2 should actually be g, the magnetic dipole moment. In the Pauli equation, which therefore is an approximation, it is
assumed to be 2.

156

CHAPTER 12. QM INTERPRETATION OF DIRAC EQUATION

Remark
The Fock-Schwinger gauge can also be formulated as follows:
x A (x) = 0
Z
A (x) =

d x F (x)

12.2.2

Foldy-Wouthousen transformation

It would be useful to diagonalize HD , in order to have the positive energy


solution only in the upper components. This can be done by the FoldyWouthousen transformation:
H0FW = UHD U

(12.20)

Free Dirac equation


For the free Dirac equation, U is given by

U=

p0
2(p0 + m)

1/2

p0 + m ~ p~
p0

(12.21)

Note that with 0 = 0 and ~ = ~ , this means that U U = 1. One can


also rewrite the numerator of the second fraction as (p m) + 0 p0 + p0 ;
Bj
orken & Drell, and Itzykson & Zuber use yet another notation:
p~
(p)
|~
p|
|~
p|
and tan(2(p)) =
m
The action of U on a spinor u is as follows:

1/2
p0
s
Uu (p) =
(1 + 0 )us (p)
2(p0 + m)
U = eiS

with S = i~

(12.22)

(12.23)

Here, we have used the Dirac equation to get rid of the term with (p m).
The factor (1 + 0 ) selects the upper components, like (1 + 5 ) in the Weyl
representation (see chapter 10). The Hamiltonian itself, finally, takes the
form
p0
1
H0FW = UHU =
(~ p~ + m + p0 )( 0~ p~ + m 0 )
2(p0 + m) p20
(12.24)
(~ p~ + m + p0 ) = 0 p0 = 0 (m2 + p~2 )1/2
Note that 0 is diagonal in the Dirac-Pauli basis, so this Hamiltonian is
diagonal as well, which was the original aim of the transformation.

12.2. NON-RELATIVISTIC LIMIT

157

Remark
~ FW in x-space should be multiplication with ~x. This implies
The action of X
~ p ; hence,
that it must be i
~ D = U X
~ FW U
X
must be given by
~ p~]
~ p~ pp~0 + i[
i~
~
~
+i
XD = ip
2p0
2p0 (p0 + m)
with
i
l = lik ik = ( i k k i )lik
2
Exercise: check this.
~ D , we have the following commutation relations:
With this X
i
k
[XD
, XD
] =0

N.b.: P~D = P~FW

i
, P j ] = i ij
[XD

[H0FW , XFW ] = i 0

(12.25)

p~
p0

The above relations imply


D
E
D
E
d D~ E
~ D ] = i [H0 , X
~ FW ] =
XD = i [HD , X
FW
dt
 
p~
p0

(12.26)

~ D, X
~ D (+) contains
in accordance with Ehrenfests theorem. With this X
only positive energy solutions. The above means that we can treat problems
in the Foldy-Wouthousen representation and then transform to the Dirac
representation by means of U.

12.2.3

Foldy-Wouthousen representation in the presence of


a constant electromagnetic field

To evaluate the Dirac equation in the presence of a constant magnetic field,


we will use the Foldy-Wouthousen representation in the notation mentioned
in eq. (12.22). We expand the exponential:
3
2
FW = eiS HD eiS = HD + i[S, HD ] + i [S, [S, HD ]] + i [S, [S, [S, HD ]]]
H
2!
3!

158

CHAPTER 12. QM INTERPRETATION OF DIRAC EQUATION

We want to evaluate the equation up to order O(m2 ). Observe that S


m1 and HD = m + e +
~ p~ m, since the mass is much larger than the
momentum. With the above equation in mind, we can make the following
ansatz:
(1)
FW = m 0 + h(0) + h + . . .
H
m

(12.27)

Let us decompose S:
eiS = eiS2 eiS1
into the following S1 and S2 :
i
~
~ (~
p eA)
2m 

0
i 1

3
~
~
S2 = 2
ie~
E
(~
(~
p eA))
m
4
6m
S1 =

Since [S1 , S2 ] m3 , we dont have to consider these terms. This gives for
H(2) :
!2
~
e
1

(2)
~
~ e 8+
eA

~ B
(12.28)
H = m + e +
2m i
2m
gm2
e 1
~ + O(m3 )
(~ L)
(12.29)
4m2 r r
The sixth term represents the spin-orbit coupling, with the Thomas factor.
The fifth term (the last one on the first line) is known as the Darwin term.
is the charge density, given by = e 3 (~x). The electron experiences a
potential V (~r + ~r) because of the zitterbewegung. Varying V and Taylor
expanding V gives roughly this term:
1 1 1 ~2
e 1 3
V
V =
(~x)
2 3 m2
6 m2
The Foldy-Wouthousen representation is fine for concrete non-relativistic
calculations. It will not do for fundamental questions concerning interactions, in particular since minimal coupling does not hold. That is, it works
only for fundamental particles like e, , q, etc., but not for protons or neutrons.

12.3

The hydrogen atom

For the hydrogen atom, we have


~

~
e2
H=
+ m
i
r




0 ~
1 0

~=
, =
~ 0
0 1


and

1
2

=


~ 0
~
and =
0 ~

(12.30)

12.3. THE HYDROGEN ATOM

159

Plugging all of this into the stationary Dirac equation gives




2
~ 2 = E1
m er 1 i~


2
~ 1 = E2
m er 2 i~

(12.31)

~ + 1 ,
~ i.e. with ~J2 and J3 . Letting the space
H commutes with ~J = ~x P
2
inversion operator act on gives
(~x) = (~x)
~ + 1 ~ and formed out
So, = j,m is characterized by j and m, with ~j = L
2
1

of Y and . Since j = l 2 , and the l-term gives a parity factor (1)l , we


()

introduce j,m . Equations (12.31) can be solved:


En,j =
with n j +
removed:

(1 +
1
2

e4 (n

(j +

and j =

1
2)

m
+ ((j + 12 )2 e4 )1/2 )2 )1/2

1 3
2, 2, . . . .

(12.32)

Thus the degeneracy in l is partially

[figure]

Remarks
~ and S
~ are coupled, L
~ and
~ do not
J~ commutes with H, but in case L
separately commute with H. Therefore, it is preferable to consider the
these commutators in the Fouldy-Wouthousen representation,
~ FW eiS
~ 0 = eiS L
L
D
~ 0 = eiS
~ FW eiS

(12.33)

~ FW and
~ FW each separately commute with HFW = 0 p0 . Hence
where L
~ 0 and
~ 0 also separately commute with HD .
we can see that L
D
D
The relation to the total angular momentum is given by
~ 0D + 1
~ 0D = L
~D + 1L
~ D = J~
L
2
2
0 p
~ p~)
~) p~ (
~ 0D = i (~

p0
p0 (p0 + m)

(12.34)

~ 0 p~ =!
~ p~ commutes with HD . Normalizing it by
The helicity operator,
D
division by |~
p|, we have
!2
~ p~

=1
|~
p|

160

CHAPTER 12. QM INTERPRETATION OF DIRAC EQUATION

~ p~/|~
i.e.,
p| has eigenvalues 1 with eigenvectors .
There also exists a split between the 2S 1 and 2P 1 levels; this is the
2
2
so-called Lamb shift, which one can try to explain by the zitterbewegung.
We do not take this approach, but instead treat it in the context of 1-loop
perturbation theory; see later.

Вам также может понравиться