Sei sulla pagina 1di 127

Ll.

Garrido
Barcelona, 2015

FISICA DALTES ENERGIES


Versio 2.0
The necessity of a quantum field theory

QM as formulated by Schrodinger in 1926 could not be used to describe the physics of


relativistic particles, the non-conservation of particle number due to production (creation of
electron-positron pairs) or annihilation (annihilation of an electron and a positron to produce
a photon) or the requirement of locality, understood as the requirement that physical processes
cannot influence each other if they are outside each others light cone.
If we restrict ourselves to phenomena of atomic physics, positrons are absent, nuclei stable,
and the energy transfers lie below the threshold for electron-positron pair creation. Moreover,
particle number is conserved and the original QM can be used even for the emission, absorption,
or scattering of photons, using a semi-classical treatment of such processes for atoms.
One of the main diculties in elaborating a relativistic QM arises from the non-conservation
of the number of particles. To be a complete theory, its dynamical states should incorporate the
number and the nature of the elementary particles of which they are composed. Such a theory
is obtained by what is called the second quantization or quantized fields, otherwise known as
quantum field theory (QFT). These notes are a journey into the formulation of QFT for electrons
and photons, known as quantum electrodynamics (QED). Historically this was not the way QED
was originally formulated in 1949 by Feynman (R.P. Feynman The theory of Positrons, Phys.
Rev:, 76 (1949) 749-759; R.P. Feynman Space-Time approach to Quantum Electrodynamics,
Phys. Rev:, 76 (1949) 769-789).

Author:
Llus Garrido (garrido@ub.edu)
Departament dEstructura i Constituents de la Materia &
Institut del Ciencies del Cosmos
Facultat de Fsiques
Universitat de Barcelona
Contents

1 A relativistic wave for spin-0 particles: the Klein-Gordon equation 1


1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1.1 Maxwells wave equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1.2 Constructing a wave equation for a particle with mass . . . . . . . . . . . 2
1.2 The Schrodinger equation (a reminder) . . . . . . . . . . . . . . . . . . . . . . . . 4
1.3 The Klein-Gordon equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.3.1 Lorentz covariance of the Klein-Gordon equation . . . . . . . . . . . . . . 5
1.3.2 Plane-wave solutions of the Klein-Gordon equation . . . . . . . . . . . . . 6
1.3.3 Conserved charge and current . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.3.4 The Klein paradox for spin-0 particles . . . . . . . . . . . . . . . . . . . . 7
1.3.5 The hydrogen atom . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

2 Relativistic wave equations, for spin-1/2 particles. The Dirac equation 11


2.1 The Dirac equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.1.1 Conserved charges and current . . . . . . . . . . . . . . . . . . . . . . . . 13
2.1.2 Cliord algebra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.2 Lorentz invariance (or covariance) of the Dirac equation . . . . . . . . . . . . . . 15
2.2.1 Rotations and boosts in the spinor representation . . . . . . . . . . . . . . 17
2.2.2 Bilinear covariants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.3 The spin of the electron . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.3.1 Review of spin-1/2 in non-relativistic QM . . . . . . . . . . . . . . . . . . 19
2.3.2 Conservation of the angular momentum L = r p . . . . . . . . . . . . . 19
2.3.3 Conservation of J . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.3.4 Non-relativistic limit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.4 Solutions of the Dirac equation. Particles and antiparticles . . . . . . . . . . . . 21
2.4.1 Treatment of the negative energy solutions . . . . . . . . . . . . . . . . . 23
2.4.2 The spin operator and the spin projection operator . . . . . . . . . . . . . 25
2.5 P , T , C and CP T . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.5.1 Parity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.5.2 Time reversal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.5.3 Charge conjugation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.5.4 CPT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
ii CONTENTS

3 From classical to quantum field theory 29


3.1 Lagrangian formalism of classical mechanics . . . . . . . . . . . . . . . . . . . . . 29
3.2 Lagrangian formulation of classical waves . . . . . . . . . . . . . . . . . . . . . . 29
3.2.1 Canonical momenta . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
3.3 Noethers theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
3.3.1 Classical mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
3.3.2 Field theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.3.3 Space-time symmetries and the energy-momentum tensor . . . . . . . . . 36
3.3.4 Continuous global symmetries. . . . . . . . . . . . . . . . . . . . . . . . . 37
3.4 Quantum fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
3.4.1 Canonical quantization rules . . . . . . . . . . . . . . . . . . . . . . . . . 38
3.4.2 Second quantization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

4 Quantization of a scalar field 43


4.1 Second quantization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
4.1.1 (x) as a creation operator . . . . . . . . . . . . . . . . . . . . . . . . . . 45
4.1.2 The causality issue . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
4.1.3 The Klein-Gordon propagator . . . . . . . . . . . . . . . . . . . . . . . . . 47
4.2 Canonical quantization rules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
4.2.1 The Hamiltonian operator . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
4.2.2 The total momentum operator . . . . . . . . . . . . . . . . . . . . . . . . 52
4.3 Quantization of charged scalar fields . . . . . . . . . . . . . . . . . . . . . . . . . 52
4.3.1 Noether current: the total charge . . . . . . . . . . . . . . . . . . . . . . . 54
4.3.2 The total Hamiltonian . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

5 Quantization of the Dirac equation 55


5.1 Second quantization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
5.2 Canonical commutation relations . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
5.3 Noether currents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
5.3.1 Space-time symmetries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
5.3.2 Internal symmetries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
5.4 Locality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
5.4.1 The Dirac propagator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
5.5 Majorana fermions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63

6 Quantization of the photon field 65


6.1 Maxwells equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
6.1.1 The Canonical momentum conjugate . . . . . . . . . . . . . . . . . . . . . 66
6.1.2 Gauge invariance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
6.2 Quantizing the field A . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
6.3 Noether currents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
6.3.1 Space-time symmetries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
6.3.2 The spin of the photon . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
CONTENTS iii

7 Interacting Fields 73
7.1 Photon-electron interaction term in the Lagrangian density . . . . . . . . . . . . 73
7.2 Representation of interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
7.3 The S matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
7.4 0 e+ e . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
7.5 Decay rates a 1 + 2 + ...n . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80

8 QED 83
8.1 e+ e 2: the electron propagator . . . . . . . . . . . . . . . . . . . . . . . . . 83
8.1.1 Amplitude of the process e+ e 2 . . . . . . . . . . . . . . . . . . . . . 83
8.1.2 The e+ e 2 cross section . . . . . . . . . . . . . . . . . . . . . . . . . 89
8.2 The gamma propagator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
8.3 The scattering cross section a + b 1 + 2 + ...n . . . . . . . . . . . . . . . . . . . 95
8.4 QED Feynman rules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96

9 Elementary processes in QED 97


9.1 e scattering by a central potential . . . . . . . . . . . . . . . . . . . . . . . . . 97
9.2 Kinematics of the process a + b 1 + 2 . . . . . . . . . . . . . . . . . . . . . . . 99
9.3 e (p1 )e+ (p2 ) (q1 )+ (q2 ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
9.3.1 Polarizations in e (p1 )e+ (p2 ) (q1 )+ (q2 ) . . . . . . . . . . . . . . . 102
9.4 e (p1 )+ (p2 ) e (q1 )+ (q2 ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
9.5 Bhabha scattering e (p1 )e+ (p2 ) e (q1 )e+ (q2 ) . . . . . . . . . . . . . . . . . . 104
9.6 Moller scattering e (p1 )e (p2 ) e (q1 )e (q2 ) . . . . . . . . . . . . . . . . . . . 105

Apendix 106

A Lorentz invariance 107

B The harmonic oscillator 111

C Equivalent descriptions 113

D The Time evolution operator 119

E Gauge Theories: a brief introduction 121

F Renormalization 131
iv CONTENTS
Chapter 1

A relativistic wave for spin-0


particles: the Klein-Gordon equation

In this chapter a wave equation is developed for particles of spin-0 compatible with the postulates
of special relativity. Schrodinger first attempted this in 1920, but eventually gave up because
of the problems of negative probability and negative energy. He ended up dealing with a non-
relativistic version which is none other than the Schrodinger equation. When he realized that
the relativistic equation was still of value, Oscar Klein and Walter Gordon, but also a few others,
had independently published it.

1.1 Introduction
Electrons and photons behave in a very similar fashion: both exhibit diraction eects, as in
the double slit experiment; and both have particle-like or quantum behavior. Photon behavior
is described by Maxwells equations and the probability that a photon is in a given small volume
of space dxdydz, is proportional to |E|2 dxdydz, the energy density. Similarly, electron behavior
x, t), such that ||2 dxdydz gives the
must be described by a wave function (x, t) analogous to E(
probability of finding the electron in a small volume dxdydz around the point x at time t. The
purpose of this introduction is to give a plausible derivation of such an equation by examining
how the Maxwell wave equation works for a single-particle (photon) wave, and constructing
parallel equations for particles which, unlike photons, have non-zero rest mass.

1.1.1 Maxwells wave equation

Maxwells equations in empty space (no sources):


E(
r, t) = 0

r, t) + B(r, t)
E( = 0,
t
B(r, t)
= 0,

r, t) 1 E(x, t)
B( = 0 (1.1)
c2 t
2 Chapter 1. A relativistic wave for spin-0 particles: the Klein-Gordon equation

To derive the wave equation, the curl of the second equation is taken:

r, t) = B(r, t) = 1 E(x, t)
2


E(
2
(1.2)
t c t2
as

. = (
.) 2 . we obtain:

x, t)
1 2 E(
2 E
= (1.3)
c2 t2
For a plane wave moving in the x-direction, this reduces to:

2 E(x, t)
1 2 E(x, t)
2
2 2
=0 (1.4)
x c t
with a solution of the form:

E(x, 0 ei(kxt)
t) = E (1.5)
Applying the dierential wave equation operator to this plane wave solution gives:
( ) ( )
2 1 2 2
0e
E i(kxt)
= k 2 2 0 ei(kxt) = 0 = ck
E (1.6)
x2 c2 t2 c

0 ei(kxt) as
Taking into account that E = h = h and p = h/ = hk allows us to rewrite E
0 ei/h(pxEt) . Thus, the wave equation operator applied to the plane wave describing
E
the particle propagation yields the energy-momentum relationship for the particle:
( ) ( )
2 1 2 i 1 E2 0 e hi (pxEt) = 0 or E 2 = c2 p2
0e
E h
(pxEt)
= 2 p2 2 E (1.7)
x2 c2 t2 h c

1.1.2 Constructing a wave equation for a particle with mass


The discussion above suggests how we might extend the wave equation operator from the photon
case (zero rest mass) to a particle with a rest mass m0 . We need a wave equation operator that,
i
when it operates on a plane wave (x, t) = Ae h (pxEt) , where A is a constant, yields the desired
E(p, m0 ) relation.

The Klein-Gordon equation

We are looking for a wave equation operator that, when it operates on a plane wave, yields

E 2 = c2 p2 + m20 c4 (1.8)

We can see that this operator is:


( )
2 1 2 m20 c2
(x, t) = 0 (1.9)
x2 c2 t2 h2
i
Writing the plane wave function (x, t) as Ae h (pxEt) , where A is a constant, the previous
operator acting over this wave gives:
( ) ( )
2 1 2 m20 c2 i 1 E2 i
Ae h
(pxEt)
= 2 p2 2 + m20 c2 Ae h (pxEt) = 0 or E 2 = c2 p2 +m20 c4
x2 c2 t2 h2 h c
(1.10)
1.1 Introduction 3

This wave equation is called the Klein-Gordon equation and correctly describes the propagation
of relativistic particles of mass m0 . As we will see, this equation has two apparent problems: it
admits negative energy and negative density solutions. We will see in this course that neither
of these problems is insurmountable and that both can be reinterpreted within the framework
of QFT. Yet in the 1920s, these diculties were considered serious enough for an alternative
to the Klein-Gordon equation to be sought. Two dierent approaches were considered: trying
to find its non-relativistic version, as followed by Schrodinger; and trying to find a relativistic
wave equation operator that includes only first derivatives, as proposed by Dirac.

A non-relativistic wave equation: the Schrodinger equation

Continuing along the same lines, let us assume that a non-relativistic electron in free space (no
i
potentials) is described by a plane wave: (x, t) = Ae h (pxEt) . We need to construct a wave
equation operator which, applied to this wave function, just gives us the ordinary non-relativistic
energy-momentum relationship, E = p2 /2m.
We can see that this operator is:
( )
h2 2
ih + (x, t) = 0 (1.11)
t 2m x2

i
Writing the plane wave function (x, t) as Ae h (pxEt) , where A is a constant, the previous
operator acting over this wave gives:
( ) ( )
h2 2 i p2 i
ih + Ae h
(pxEt)
= E Ae h (pxEt) = 0 or E = p2 /2m (1.12)
t 2m x2 2m

This wave equation is the Schrodinger equation for a free particle. With a non-zero potential
present, the energy-momentum relationship for the particle becomes the energy equation E
p2 /2m V (x) = 0 and the appropriate wave equation is:
( )
h2 2
ih + V (x) (x, t) = 0 (1.13)
t 2m x2

which is the standard one-dimensional Schrdinger equation

A relativistic wave equation: the Dirac equation

Dirac argued that the source of negative solutions of the Klein-Gordon equation was related to
the E 2 term that comes from the second time derivative. For that reason, he proposed a wave
equation operator that included only first-order derivatives, so that its square still gives the
relation E 2 = p2 c2 + m20 c4 . From this approach, the problem of negative densities was solved
but the negative solutions remained. Interpreting these solutions as antiparticles, the Dirac
equation is the relativistic equation that describes particles and their antiparticles of spin 1/2.
Full details can be found in chapter 2.
4 Chapter 1. A relativistic wave for spin-0 particles: the Klein-Gordon equation

1.2 The Schrodinger equation (a reminder)


The familiar Schrodinger equation can be obtained by the correspondence rule. We start with
a classical dynamical system represented by a Hamiltonian H(q1 , ..., qn ; p1 , ...., pn ; t), which de-
pends on the coordinates qi of the system in configuration space, on the momenta pi , and on
the time t. The total energy of the system is:

E = H(q1 , ..., qn ; p1 , ...., pn ; t) (1.14)

There is a quantum system that corresponds to this classical system whose dynamical state is
represented by a wave function (q1 , ...., qn ; t) defined in configuration space and whose wave
equation can be obtained by performing the following substitutions on both sides of equation
1.14:

E ih , pi ih (1.15)
t qi
For a classical system with H = p2 /2m + V (x) this leads to:
(x, t) h2 2
ih = ( + V (x))(x, t) (1.16)
t 2m
which is the well-known Schrodinger equation. This equation allows us to construct a non-
relativistic probability density and current such that a continuity equation is obeyed:

ih (||2 ) = ih + ih = (H) (H ) =
t t t
h2 2 h2 2 h2 2
( ) ( ) = 2)
( (1.17)
2m 2m 2m
which leads to:
ih ) = 0
(||2 ) (
t 2m

+j =0
t
ih
(x, t) = |(x, t)|2 0 , j = ( () ())
(1.18)
2m
For a free particle, the classical Hamiltonian is H = p2 /2m, and its wave equation is:
h2 2
iht (x, t) = (x, t) (1.19)
2m
whose solutions are of the form:
N exp(it + ikx) (1.20)

with the eigenvalue of energy E = h ( as (ih t ) = ih(i) = h ) and the eigenvalue of

momenta p = hk (as (ih) = ih(ik) = hk). Plugging this back into the Schrodinger

equation we obtain E = p2 /2m (notice that the eigenvalue of the operator iht is always posi-
tive).
From the definition of and j given in 1.18 we obtain:

= N2 , j = 1 N 2 ( p)) = N 2
p + (+
p
(1.21)
2m m
since in classical mechanics p = mv , this corresponds to j = v .
1.3 The Klein-Gordon equation 5

1.3 The Klein-Gordon equation


Since we are after a relativistic equation, it seems natural to replace the non-relativistic Hamil-
tonian by its relativistic counterpart:

H= p2 c2 + m2 c4 , (1.22)

leading to:


ih = (ih)
2 c2 + m2 c4 = h2 c2
2 + m2 c4 (1.23)
t
The best we can attempt here is to expand the square root by its power series expansion:

(hc)
2 (hc)
4
h2 c2
2 + m2 c4 mc2 (1 ...) (1.24)
2(mc2 )2 8(mc2 )4

thus arriving at a non-local dierential equation (of arbitrarily high order in the space deriva-
tives) which violates causality, so clearly this possibility is not satisfactory. In addition, rela-
tivistic invariance is not clearly exhibited since there is lack of symmetry between the space and
time coordinates.

Alternatively, we can take another time derivative of the l.h.s. of the Schrodinger equation to
give us:
2
h2 2 = H 2 = (h2 2 c2 + m2 c4 ). (1.25)
t
Defining:
1 2
2 2 2 2 = , (1.26)
c t
we then get:
m2 c 2
(2 + 2 ) = 0. (1.27)
h
This equation, which is manifestly relativistic, is called the Klein-Gordon equation. Its form is
not surprising since the only two Lorentz invariants with dimensions of inverse length squared
available to us are and m2 . If the particle has no rest mass, m = 0, we obtain the classical
wave equation, which is also relativistically invariant.

1.3.1 Lorentz covariance of the Klein-Gordon equation


According to the special theory of relativity, all physical laws of nature must have the same
form in any two coordinate system which arise from each other by proper orthochronous Lorentz
transformations (det() = +1, 00 +1). A relativistic theory need not be invariant under P
(spatial reflexion) or T (time reflection). The dAlambert operator 2 is invariant under Lorentz
transformation because it is a scalar product of the four-vector . Also, the mass m is a scalar,
so the operator (2 + m2 ) is Lorentz invariant. It remains for us to show that is invariant under
a Lorentz transformation. Transforming from one inertial reference frame to another denoted
by primes:
m2 c2 m2 c2
(2 + ) (x ) = 0 (2 + ) (x ) = 0 (1.28)
h2 h2
6 Chapter 1. A relativistic wave for spin-0 particles: the Klein-Gordon equation

if (x ) = (x), with || = 1, the Klein-Gordon equation is Lorentz invariant. Since the


Lorentz transformation operator is real, must be real and hence = 1.
If the Lorentz transformation is continuously connected to the identity operator, then =
+1. For space inversion, as P 2 = I, we have two possibilities (x , t ) = (x, t) = 1(x, t).
Therefore, solutions of Klein-Gordon equation are invariant under proper orthochronous Lorentz
transformation and invariant (scalar) or change sign (pseudo-scalar) under space inversion. This
last value, 1, is called the intrinsic parity of the particle and can only be determined by studying
its interactions.
From now on, we will work in units where h = c = 1.

1.3.2 Plane-wave solutions of the Klein-Gordon equation


Plane waves are a solution of this equation (exactly as in the non-relativistic case), but now the
dispersion relation is relativistic. Using four-vectors, we write a plane wave solution as:

(x) expipx , where p = (p0 , p), x = (t, x), px = p0 t px. (1.29)

The choice of a relative minus sign between the 2 terms in the exponential is arbitrary since the
Klein-Gordon equation is second order in the derivatives. Our convention is in accordance with
non-relativistic quantum theory. Plugging this back into the Klein-Gordon equation we indeed
obtain:

0 = ( + m2 ) = ((ip )(ip ) + m2 ) = ((p0 )2 + p2 + m2 )



p0 = p2 + m2 . (1.30)

However, we can see that because of the square root we get negative as well as positive energy
solutions! The negative solutions cannot be discarded without good reason, since the Klein-
Gordon equation is meant to be a fundamental equation of nature and we have naturally to
accept all solutions.
The solutions can be written as:

ipx
()
p (x) = N e = N ei(Etpx) (1.31)

where now p is the four energy-momentum (E, p), with E = + p2 + m2 , and x is the four-
vector (t, x). According to this convention, the positive (negative) energy solution has momen-
tum p(p) as can be seen by applying the energy-momentum operator i = i(t , ):

i () = i(i)p () = p () (1.32)

1.3.3 Conserved charge and current


However, that is not the only diculty. We know from QM that the Schodinger equation leads
to a continuity equation which can be interpreted as the conservation of probability:

i )) =
= ( () ( j, (1.33)
t 2m
1.3 The Klein-Gordon equation 7

with = . Let us try to derive a similar equation from the Klein-Gordon equation. First we
multiply the equation by the complex field:

(( ) + m2 ) = 0
( ) ( )( ) = m2 . (1.34)

Secondly, we multiply the complex conjugation of the Klein-Gordon equation by the field
itself:

(( ) + m2 ) = 0
(( )) ( )( ) = m2 . (1.35)

Now, subtracting one equation from the other we obtain:

( ( ) ( )) = 0, (1.36)

which is indeed a continuity equation j = 0 and is normally taken as:

j = i( ( ) ( )) (1.37)

Therefore, we may try to interpret j 0 as a probability density:

j 0 (x) = i( (t ) (t )) (1.38)

but this would-be probability density is not positive definite. The dierence between this and
the non-relativistic case is the presence of an extra t which completely spoils positivity, as Dirac
pointed out. For the planar wave solution given in 1.31, we obtain:

j = iN 2 (ip (ip )) = 2N 2 p = 2N 2 (E, p) (1.39)

and its 0-th component is j 0 = 2N 2 E. Thus the problem of negative energy and that of
negative probability are related. Let us advance that the Dirac equation still proves impossible
to retain a positive definite probability density for a single particle, and at the same time
provide a physical interpretation of the negative energy solution. This means that the Klein-
Gordon equation is no worse than the Dirac equations which respect the physical interpretation
of probability for a single particle. The equation 1.39 suggests that we could interpret |q|j as
a current density four-vector. and have opposite charge, while the real solutions represent
uncharged particles.
The Klein-Gordon equation can be used to describe spin-0 particles such as pions (real
solutions for 0 and complex ones for , both pseudo-scalar). However, pions are not elemen-
tary: they are made of quarks. To date, the only elementary spin-0 particle is the Higgs boson
discovered at LHC in 2012.

1.3.4 The Klein paradox for spin-0 particles


Consider the scattering of a Klein-Gordon particle of energy E and momentum pz from a step
function potential, as shown in Figure 1.1. The Klein-Gordon equation in the presence of this
8 Chapter 1. A relativistic wave for spin-0 particles: the Klein-Gordon equation

potential becomes:

2
+ 2 m2 = 0, z < 0
t2

(i V0 )2 + 2 m2 = 0, z > 0 (1.40)
t
An incoming positive energy beam is a plane wave of the form ei(Etpz) . We look for solutions

Figure 1.1: Electrostatic potential

of the form:

< = eiEt (eipz + Reipz ), z < 0



> = T eiEt eip z , z > 0 (1.41)

where R and T are reflected and transmitted amplitudes, p is the momentum in region II, and:

E 2 p2 m2 = 0 E = p2 + m2

E 2 + Vo2 2EV0 p2 m2 = 0 p = (E V0 )2 m2 (1.42)

In region II, there are three distinct cases, depending on the strength of the potential:

weak potential: V0 < E m E V0 > m p is real. Standard transmission and


reflexion.

intermediate potential: (E m) < V0 < (E +m) |E V0 | < m p is purely imaginary.


Total reflexion, no transmission.

strong potential: V0 > E + m |E V0 | > m p is real. New transmission () and


reflection () ( with > 1 and < 0).

Imposing the boundary conditions that and /z be continuous at z = 0 gives:

2p p p
T = , R= (1.43)
p + p p + p
1.3 The Klein-Gordon equation 9

Calling ji = 2p, j< , j> the incident, left and right currents, the transmission and reflection
coecients are:
ji j< p p 2 j>
= = |R|2 = |
| , = =1 (1.44)
ji p+p ji
For the case of the strong potential, these coecients are:

p + |p | 2 4p|p |
=( ) >1 , = <0 , +=1 (1.45)
p |p | (p |p |)2

The probability is still conserved but at the cost of having a negative transmission coecient
and a reflexion greater than 1. We are stuck in a many-body problem. This apparent paradox
can be explained if we consider the potential strong enough to create a particle-antiparticle pair.
Antiparticles are create a negative current on the right; while an excess of particles is moving
to the left.

1.3.5 The hydrogen atom


For stationary states = eiEt (x) we can write:

Ze2 2
(E + ) + (h2 c2 2 m2 c4 ) = 0 (1.46)
r
For a solution of definite angular momentum l, (r, , ) = R(r)Yl m, l(, ), and writing the
Laplacian in spherical coordinates, the equation for R(r)becomes:
[ ] [ ]
(E + Ze2 /r)2 m2 c4 1 d d l(l + 1)
2 2 R = 2 (r2 ) + R (1.47)
h c r dr dr r2

Solving this equation for R and E, the eigenvalues for the energy are: E = mc2 (1 + 2 /2n2
4 /2n4 (n/(l + 1/2) 3/4)), with Ze2 /hc. At the lowest order, they are the same as those
we can obtain with the Schrodinger equation, by altering the classical kinetic energy as follows:

p2 1 p4
p2 /2m p2 + m2 m2 (1.48)
2m 8 m3 c2
10 Chapter 1. A relativistic wave for spin-0 particles: the Klein-Gordon equation
Chapter 2

Relativistic wave equations, for


spin-1/2 particles. The Dirac
equation

2.1 The Dirac equation

In the course of what follows, we will see that neither of the two problems mentioned in the
previous chapter (negative energy and negative density) are insurmountable and that they can be
re-interpreted in the framework of QFT. Yet in the 1920s they were considered serious enough
for an alternative to the Klein-Gordon equation to be sought. Dirac proposed an equation
including only first-order derivatives:

ih = ihc

+ mc2 H. (2.1)
t
This could be seen in some way as taking the square root of the Klein-Gordon equation.
Obviously Klein-Gordon, which in practice implies the relativistic dispersion relation E =

p2 c2 + m2 c4 must still be satisfied. The condition H 2 = (ihc)
2 + m2 c4 then becomes:

H 2 = (ihci i + mc2 )(ihcj j + mc2 )


= i j (ihci )(ihcj ) + mc2 i (ihci ) + mc2 j (ihcj ) + 2 m2 c4

= (i )2 (ihci )2 + (i j + j i )(ihci )(ihcj )
i>j

+mc2 (i + i )(ihci ) + 2 m2 c4 (2.2)

This implies that:

i j + j i = 0, (i = j)
i + i = 0
2 = i2 = 1 (2.3)

It is clear that these conditions cannot be fulfilled by ordinary numbers.


12
Chapter 2. Relativistic wave equations, for spin-1/2 particles. The Dirac equation

Due to the hermiticity of the Hamiltonian, these 4 matrices are Hermitian: i = i , = .


Since i2 = 2 = I, their eigenvalues are 1. Also, it is easy to prove that they are all traceless.
As i = i , then:

T r(i ) = T r(i ) = T r( 2 ) = T r(i ) T r(i ) = 0 (2.4)

Similarly, if we start from i i = , we get T r() = 0. Since the trace does not change
under the transformations A S 1 AS , and 0 should be obtained as the sum of eigenvalues
1, the rank of these matrices should be even. The simplest solution involves 4 4 matrices. A
possible solution is: ( ) ( )
0 i I 0
i = , = (2.5)
i 0 0 I
where I is the 2x2 unit matrix and i are the usual Pauli matrices:
( ) ( ) ( )
0 1 0 i 1 0
1 = , 2 = , 3 = (2.6)
1 0 i 0 0 1
which allows us to write = 3 I, and i = 1 i .
It is desirable to write equation 2.1 in a way that time and space are treated on an equal
footing. To do this, let us define:
( ) ( )
I 0 0 i
=
0
, i =
i
(2.7)
0 I i 0
The gamma matrices are called Dirac matrices and they satisfy the anti-commutation relations:

{ , } = 2g . (2.8)

Note that 0 is Hermitian ( 0 = 0 ), but i is not ( i = (i ) = i = i = i = i )
, but the following is always true: = 0 0 .
Multiplying the original Dirac equation ih
t + ihc
mc2 = 0 by 0 we get:


ih mc = 0. (2.9)
x
Defining the symbol:
, (2.10)
the Dirac equation now reads:
(ih mc) = 0, (2.11)
and the fully expanded Dirac equation in the standard representation is:

1 0 0 0 0 0 0 1

0 1 0 0 0 0 1 0
( i h 1 + ih
0
1 0 c t 0 1 0 0
0 x
0 0 0 1 1 0 0 0

0 0 0 i 0 0 1 0 a (t, x)

0 0 i 0 0
+ ih
0 0 1 (t,
mc) b
x)
+ i h 0
= 0 (2.12)
i 0 0 y 1 0 0 0 z
(t,
c x)

i 0 0 0 0 1 0 0 d (t, x)
2.1 The Dirac equation 13

2.1.1 Conserved charges and current


It is easy to see that one of the problems, the sign problem, is still there and we will eventually
have to consider just how to interpret the solutions. The other one is, however, solved: a
probability current can be defined and with > 0. Let us multiply the Dirac equation by
= (1 , 2 , 3 , 4 ):

0 = (ih mc) = ih mc . (2.13)

and the Hermitian conjugate of the Dirac equation by :

0 = (ih mc) = (ih( ) mc ) = ih( ) mc (2.14)

Subtracting the two equations we get:

+ ( ) = 0 (2.15)

which could easily be interpreted as a conservation-of-probability equation j if = , but


it is not! The solution is to introduce the conjugate of :

0 = (1 , 2 , 3 , 4 ) (2.16)

and multiply the Hermitian conjugate of the Dirac equation by 0 from the right:

0 = ih( ) 0 0 0 mc 0 = ih( ) 0 0 mc = ih( ) mc


(2.17)
In contrast, multiplying the Dirac equation by , we obtain ih ( ) m = 0 and

subtracting it from the previous equation:

( ) + ( ) = ( ) = 0 (2.18)

Thus the conserved current for the Dirac field is , which looks Lorentz invariant, but to
be sure we have to find out how the 4-component spinor transforms under Lorentz transfor-
mations. That will be tackled in the following sections, where we show that the Dirac equation
is Lorentz invariant. For now, however, let us look at the 0-th component of j :

j 0 = 0 = = |1 |2 + |2 |2 + |3 |2 + |4 |2 0 (2.19)

which is positive and so solves the problem we had with the Klein-Gordon equation.

2.1.2 Cliord algebra


This is defined by the relation
{ , } = 2g . (2.20)

In addition the matrix 5 = i 0 1 2 3 = 4!i is also introduced. We will see its


utility later. There are infinite representations of this algebra, but three of them are commonly
used.
14
Chapter 2. Relativistic wave equations, for spin-1/2 particles. The Dirac equation

Dirac representation:
( ) ( ) ( )
0 I 0 k 0 k 5 0 I
= , = , = . (2.21)
0 I k 0 I 0

Chiral representation:
( ) ( )
0 0 I k 0 k
= , = . (2.22)
I 0 k 0
Majorana representation:

( ) ( ) ( ) ( )
0 0 2 1 i3 0 2 0 2 3 i1 0
= , = , = , = . (2.23)
2 0 0 i3 2 0 0 i1

In d = 4 there are 16 independent combinations of gamma matrices (i , 1 = 1, ..., 16):

I,
0 , i 1 , i 2 , i 3 ,
0 1 , 0 2 , 0 3 , i 1 2 , i 1 3 , i 2 3 ,
i 0 5 , 1 5 , 2 5 , 3 5 ,
5.

All these matrices satisfy

(A )2 = I
T rA = 0
A B = C
1
X, X = T r(XA )A
4 A
A = 1, B; {A , B } = 0.

Other important properties following directly from the definition of 5 and the anti-commutation
relation defining Cliord algebra are:

{ , 5 } = 0, (2.24)

( 0 )2 = I, ( i )2 = I, ( 5 )2 = I, (2.25)

( 0 ) = 0 , ( i ) = i , ( 5 ) = 5 , 0 ( ) 0 = , (2.26)

A A = A2 , A B = 2A B B A (2.27)
2.2 Lorentz invariance (or covariance) of the Dirac equation 15

Traces of Dirac matrices:

T r( A) = 0
T r( A1 A2 A2n+1 ) = 0
T r( A B) = 4AB
T r( A B C D) = 4(A B C D A C B D + A D B C)
T r( ) = 4(g g g g + g g )
T r( 5 ) = 0
T r( 5 ) = T r( 5 ) = T r( 5 ) = 0
T r( 5 ) = 4i .

with
0123 = 1, 0123 = 1. (2.28)

2.2 Lorentz invariance (or covariance) of the Dirac equation


We know the transformation properties of vectors under Lorentz transformations (see Appendix
A):
x x = x = (2.29)

Spinors, i.e. solutions of the Dirac equation, are 4-dimensional objects, but how do they
transform under the Lorentz group in order to preserve covariance?

(x) (x ) = S()(x), (2.30)

where S represents the action of the Lorentz group in the space of spinors. Obviously S(1 ) =
S 1 ().

Taking as our starting point the Dirac equation:

(i m)(x) = 0, (2.31)

(i m)S 1 () (x ) = 0 (2.32)

multiplying by S():
( )
iS() S 1 () m (x ) = 0 (2.33)

S() S 1 () = . (2.34)

or:
S 1 () S() = . (2.35)

This equation tells us that the transformation of matrix under S is like the transformation
of a 4-vector: x . One can say that transforms as a 4-vector in the spinor representation
16
Chapter 2. Relativistic wave equations, for spin-1/2 particles. The Dirac equation

of Lorentz transformation. This should be expected to make the in the Dirac equation
unchanged under Lorentz transformations.
Another property required for S() is that it should have the same multiplication rule as
the Lorentz transformation: S(1 2 ) = S(1 )S(2 ). In other words S is a representation of
the Lorentz group.
Any proper Lorentz transformation can be expressed as = ei Ki +i Li (see Appendix A).
It is convenient to write it in a compact form defining:
{ {
Ki M 0i Li M jk
(2.36)
i a0i i ajk

where i,j,k are cyclic and we take M = M and a = a . Note that the matrix elements
of M can now be written as:

(M ) = g g g g (2.37)

and
1
= ei Ki +i Li = e 2 a M (2.38)

As S and have the same group structure, let us try the mapping:
1 1
= e 2 a M S = e 2 a B (2.39)

where B is a set of 6 independent generators (4x4) which should satisfy the same commutation
relations as M , and in addition verify that S 1 S = .
Let us now consider infinitesimal transformations (a small):

S 1 () S() = e 2 a B e 2 a B
1 1

1 1
= (1 a B ) (1 + a B ) + ....
2 2
1
= + ( a B a B ) + ...

2
1 [ ]
= + ,B a + .... (2.40)
2
However:
1 1
= (e 2 a M ) = (1 + a M ) + ...
2
1 1
= (g + a (M ) ) + ... = + a (M ) + .. (2.41)
2 2
equating the 2 expressions above, we obtain the requirement for B :
[ ]
, B = (M ) (2.42)
[ ]
It is left as an exercise to prove that B = 14 , is a solution of the previous expression,
which has been obtained for infinitesimal transformations of 2.35. Prove also, that it is the
solution for finite Lorentz transformations.
2.2 Lorentz invariance (or covariance) of the Dirac equation 17

The Lorentz transformation of a spinor is some times written as:


i
S = e 4 a , with = 2iB =
i
[ , ] (2.43)
2
which reproduces the same S for a given a . The are the Poincare generators acting on
spin-1/2 fields. They are not all Hermitian, but:

( ij ) = ij , ( 0i ) = 0i . (2.44)

One advantage of this definition is that ij becomes very much like k (i,j,k cyclic) acting in
the spinor space:
( )( ) ( ) ( )
i [ i j] 0 i 0 j i j
k 0 0
(i = j), =
ij
, = i i j = i =i =
2 i 0 j 0 0 k
0 i j
(2.45)
we can define k (i,j,k cyclic), and observe that k /2i satisfy the same commutations as
ij

Li of equation 2.38. More will be said regarding k when the spin of the electron is discussed.
The boost are generated by B 0i = 41 [ 0 , i ] = 12 0 i = 12 0 0 i = 12 i that satisfy the same
commutations as Ki in equation 2.38. This fact and the previous one show that we have found
the generators of the Lorentz group in the spinor representation.

2.2.1 Rotations and boosts in the spinor representation

A Lorentz transformation acting on a spinor is given by the expression (2.30), with S() given
by:
S() = e 4 a .
i
(2.46)
i
where = 2 [ , ]. In the case of rotations:

S() = e 4 aij = exp ik k /2


i ij
(2.47)

as k is diagonal with regard to the Pauli matrices k , the general rotation matrix eis/2 will

be: ( )
is/2
eis/2 0
e = (2.48)
0 eis/2
showing that a rotation in the spinor space acts as a double rotation over two spin-1/2 states.
A boost in the Dirac representation is given by:

S() = e 2 a0i = ea0i B = ei B = e 2 i i


i 0i 0i 0i 1
(2.49)

where we have used B 0i = 41 [ 0 , i ] = 12 0 i = 12 0 0 i = 21 i .


Writing = n, and noting that (n )2 = ni nj (i j + j i )/2 = ni nj 2ij /2 = n2 = 1, the
matrix S can now be expanded:
( ) ( )
1 1 2 1 3
S = e 2 i i = e 2 n = 1 + (n) + )2 +
(n )3 + ...
(n
2 2! 2 3! 2
( ( ) ) ( ( )3 )
1 2
= 1+ ... + + + ... n = cosh + sinh n
2! 2 2 2 2 2
( ) ( )
cosh 2 sinh 2 n 1 tanh 2 n
= = cosh (2.50)
sinh 2 n cosh 2 2 tanh 2 n 1
18
Chapter 2. Relativistic wave equations, for spin-1/2 particles. The Dirac equation

to express it in terms of and remember that = cosh and



+1 1
cosh = , sinh = , tanh = (2.51)
2 2 2 2 2 +1

in addition E = m, p = m, then:

E+m p
cosh = , tanh = (2.52)
2 2m 2 E+m

and S can be written as:


( )
p

E+m 1 E+m
S= p
(2.53)
2m E+m 1

2.2.2 Bilinear covariants

We have defined the conjugate spinor as = 0 . Similarly, let us define the conjugate of a
4x4 matrix in spinor space M by:
M 0M 0 (2.54)

then we have AB = 0 (AB) 0 = 0 B 0 0 A 0 = B A, and the relation 0 0 = can be


written as = . Then B = B .
Using this last expression we get:

0 S 0 S = e1/2a B (1/2a B )k /k! = e1/2a B = S 1

= (1/2a B )k /k! =
k k
(2.55)
Note that 0S0= S 1
= S0S 0,
which can be compared to T G = G. We can say
0
that acts as a metric for the spinor representation.
The conjugate spinor transforms in the following way:

(x ) = S()(x) = (S()) 0 = S () 0 = S 1 (). (2.56)

Thus (x)(x) is a scalar.


Similarly it can be shown that the following bilinear forms O behave as:

O = I scalar(R)

O = vector(V)

O = tensor(T)

O = 5 axialvector (A)

O = 5 pseudoscalar(P)

These are 16 O independent operators.


2.3 The spin of the electron 19

2.3 The spin of the electron


2.3.1 Review of spin-1/2 in non-relativistic QM
In spin-1/2 space, the angular momentum operator, L, is represented by the Pauli matrices:
Li = i /2. The spin operator along the unit vector s is represented by s /2 with eigenvalues
1/2 and eigenvectors:
( ) ( )
1 1 + sz 1 1 sz
s+ = , s = , where s = sx isy (2.57)
2(1 + sz ) s+ 2(1 sz ) s
Any state vector v in the spin 1/2 space can be expressed as a combination of the two
eigenvectors of s /2 by the following projection operator:
1 1
v = P+ (s)v + P (s)v, with P (s) = s (2.58)
2 2

2.3.2 = r p
Conservation of the angular momentum L
We want to compute [Li , H] with Li = ijk rj pk and H being the Dirac Hamiltonian. It is left as
an exercise to prove that:
[Li , H] = i 0 ijk j pk (2.59)
using the facts that ri ( ) commutes with all but pi ( , = ). This shows that the angular
momentum is not a conserved quantity.

2.3.3 Conservation of J
We have seen that the ordinary angular momentum in not conserved. Now let us calculate the
and H. It is left as an exercise to prove that:
commutation relation of

[i , H] = 2i 0 ijk j pk . (2.60)

Therefore we have a conserved quantity J r p + 12 as [Ji , H] = 0. At this point, recall that

/2 is very much like /2 except that it operates in the 4-component spinor space. i /2, i =
1, 2, 3 have the same commutation relations as Li , from their definition it is clear that their
eigenvalues are 1/2 (both double) and furthermore, we have:
1 2 1 2 3 1
| | = (1 + 22 + 23 ) = = j(j + 1), with j = (2.61)
2 4 4 2
This suggests that any fermion that satisfies the Dirac equation carries an intrinsic angular
momentum of 1/2 unit. This was an entirely unexpected result for Dirac himself. It also turned
out it had the correct magnetic moment, as we will see next.

2.3.4 Non-relativistic limit


We now want to prove that this equation actually describes particles of spin-1/2. We will do
that by comparing the Dirac equation with the non-relativistic Pauli equation that describes
the propagation of an s = 1/2 particle in a background electromagnetic field. That equation is:
p ec A)
( 2
ih =[ gs B S
B + e]. (2.62)
t 2m
20
Chapter 2. Relativistic wave equations, for spin-1/2 particles. The Dirac equation

In the equation above, B is the magnetic field induced by the vector potential A, S is the
spin operator ( /2), and B = eh/2mc is the Bohr magneton. is a 2-component vector
representing the two spinorial degrees of freedom of the particle. The gyromagnetic factor of
the spin has to be determined experimentally; it is known to be very close to 2 for electrons.
In order to make contact with (2.62), we have to couple electromagnetism to the Dirac equation.
This is done via the so-called minimal coupling:
e
p p A , (2.63)
c
At this point this is just a prescription; it will be justified later in the
where A = (c, A).
course.
After this, the Dirac equation reads:
e
( p A mc) = 0, (2.64)
c
or:
+ e A)
ih = c
(ih + (mc2 + e). (2.65)
t c
We now make the change ( )
imc2 t/h
=e (2.66)

and we obtain:
( ) ( ) ( ) ( )
e 0
ih p A) I
= c ( + e 2mc 2
. (2.67)
t c

The non-relativistic limit is obtained by taking c . Then, keeping the leading terms in the
previous equations, from the lower components we get:

p ec A)
(
= . (2.68)
2mc
Note that we can also conclude (from the upper components) that 0 (to leading order in
a 1/c expansion), which of course is implied by the previous equation too. Plugging back in,
we get: [ ]
p ec A)][
[ ( (p ec A)]

ih = + e . (2.69)
t 2m

All we need to do now is to use the property

i j = ij + iijk k (2.70)

to simplify the r.h.s. further. The result is:


e e e 2 eh
p A)][
[ ( p A)]
( p A)
= ( 2 S B. (2.71)
c c c c
Thus we reproduce the Pauli equation, predicting that gs = 2 ! This is a major triumph of
the Dirac equation. Of course, the previous exercise shows the validity of the Dirac equation to
describe the electromagnetic interactions of spin-1/2 particles.
2.4 Solutions of the Dirac equation. Particles and antiparticles 21

2.4 Solutions of the Dirac equation. Particles and antiparticles


Let us consider a particle at rest: p = 0. In this case, the Dirac equation reduces to:
( )
I 0
i = m = m . (2.72)
t 0 I

It is immediately clear that the 4-dimensional vectors (spinors):



1 0 0 0

0 1 0 0
1 = N
0
2 = N
0
3 = N
1
4 = N
0, (2.73)

0 0 0 1

are a base of the spinor space, and each one is part of a solution of the Dirac equation:

1 0

0 imt 1 imt
1 = N
0e
, 2 = N
0e


0 0

0 0

0 +imt 0 +imt
3 = N
1e
, 4 = N
0e
(2.74)

0 1

We can see at once that, just as for the Klein-Gordon equation, the Dirac equation has negative
energy solutions (3 , 4 ) that will be interpreted later. Let us see how these solutions respond

to the spin operator /2. We have seen that L + /2
is a constant of motion and at rest L = 0.
Taking the expression 3 /2 in the Dirac representation we can see that 1 and 3 represent
spin up in the z-direction, while 2 and 4 represent spin down.

Using N = 2m as the normalization constant for these 4 at-rest solutions, for each i-solution
we obtain (j = ) :

j 0 = i 0 i = i i = 2m, i = 1, 2, 3, 4
j k = i k i = i 0 k i = i k i = 2m(k )ii = 0 (2.75)

since the diagonal elements of k are null. Thus, for all the at-rest solutions, we have obtained:

j = (2m, 0) (2.76)

Since it is known that j transforms as a vector, for the boosted solutions we should obtain:

j = (2E, 2
p) = 2p (2.77)

Recall that j 0 is interpreted as the probability density: the number of particles per unit volume.
This is 2m at rest and 2E for the boosted solutions, as we must expect due to the Lorentz
contraction (the number of particles per unit volume is 2m = 2E).
22
Chapter 2. Relativistic wave equations, for spin-1/2 particles. The Dirac equation

Note that i i = +2m(2m) for positive(negative) solutions. Since is a scalar, this is


also true for the boosted solutions.
In order to obtain the solution for p = 0, we can apply a boost. While it is clear how the
space-time dependence appears:
eimt eipx , (2.78)
the transformation properties of the spinors are given by 2.53 Now we are in a position to apply
these transformations and obtain the general solutions for the spinors. The transformed spinors
after a boost of 3-momentum p are:

1 0

0 ipx 1 ipx
1 = E + m
p3
e
, 2 = E + m
p1 ip2
e

E+m E+m
p1 +ip2 p3
E+m
p3 E+m
p1 ip2
|E|+m |E|+m
p3
p1 +ip2
+ipx +ipx
3 = E + m

|E|+m e
, 4 = E + m

|E|+m e
. (2.79)
1 0
0 1
Note an interesting feature of the boosted solutions: while in the rest frame i are eigenstates
of 3 , this is not the case for the boosted solution given above (except when the boost is along
the spin-quantized axes: Z in this case). In general the spin state is defined as usual in the rest
frame.
In order to construct a set of solutions with the spin quantization in a general direction s,
all we have to do is to construct a set of 4 states at rest which are spin-quantized along s and
to boost them.

The 2 states with positive energy and eigenstates of s/2 (spin up and spin down) will be:
( ) ( )
s+ imt
s
1 = 2m e , 2 = 2m eimt (2.80)
0 0
Similarly, for negative energy we have:
( ) ( )
0 0
3 = 2m eimt , 4 = 2m eimt (2.81)
s+ s
Then, applying the boost operator 2.53 to boost in the direction of p by a velocity given by
= E/m, we get a set of solutions with definite energy-momentum (p = (E, p) for +ve energy
solutions and p for -ve energy solutions), and spin-quantized (in the rest frame) in a general
direction, s:
( )
s+
1 = E+m p

eipx w(
p, s)eipx ,
E+m s+
( )
s
2 = E+m p

eipx w(
p, s)eipx ,
E+m s
( )
p


E+m s+
3 = E+m eipx w(
p, s)eipx ,
s+
( )
p


E+m s
4 = E+m eipx w(
p, s)eipx , (2.82)
s
2.4 Solutions of the Dirac equation. Particles and antiparticles 23

where w( p, s) denotes the energy/momentum and the spin in the s-direction. This set
of 4 spinors constitutes a base of the spinor space.
The positive and negative energy is defined by the eigenstate of the energy operator it . For
positive energy:

0 = (i m)w exp(ip x ) = ( p m)w exp(ip x ), (i p ) (2.83)

while for negative:

0 = (i m)w exp(+ip x ) = ( p m)w exp(+ip x ), (i p ) (2.84)

thus, the solution of the equations (in momentum space) obeys:

( p m)w = 0 for + energy, ( p + m)w = 0 for energy (2.85)

and taking their adjoin:

0 = ( p m)w = w( p m) = w( p m) for + energy


0 = ( p + m)w = w( p + m) = w( p + m) for energy
(2.86)

In the expressions for the spinors w( p, s) we have explicitly written E 2 = p2 + m2 . In


fact, this is a condition on w so that w exp ipx is a solution of the Dirac equation. As an
example, for positive solution spinors ( ( p m)w(+p, s) = 0) we can deduce it thus:

0 = ( p + m)( p m)w = ( p p m2 )w = (p2 m2 )w (2.87)

since p p = p p = 21 p p ( + ) = p p g = p2 . And similarly for negative energy


spinors.
We can construct a pair of energy-projection operators ( p) = ( p + m)/2m where
0
p = +E. For any given 4-component column vector w:

p)weipx and (
+ ( p)weipx (2.88)

are both solution of the Dirac equation; the first for positive energy and the second for the
negative energy. For example, for any spinor w (which can be expressed as a combination of the
spinor base w( p, s) with p2 = m2 ), + (p)w selects the positive energy spinor part, since it
satisfies:
( p m)+ (p)w = ( p m)( p + m)w/2m = (p2 m)w/2m = 0 (2.89)

As expected, 2 (
p) = (
p), + (
p) (
p) = (
p)+ (
p) = 0, + (
p) + (
p) = 1

2.4.1 Treatment of the negative energy solutions

We have to face the problem of interpreting negative energy solutions. Dirac proposed that the
vacuum state is such that all negative energy states are filled up. Since we are dealing with
fermions, it is impossible for any such state to be further occupied. This is called the Dirac sea.
24
Chapter 2. Relativistic wave equations, for spin-1/2 particles. The Dirac equation

If we add fermions, they can only occupy a positive energy state. Furthermore, there is a 2m
gap between the highest negative energy state and the lowest possible positive energy state.

If we inject energy into the system (for instance, via a virtual photon or a couple of real photons)
it is possible to promote one of the negative energy states to positive energy. We have thus
produced a particle with positive energy and a hole in the sea of negative energy, i.e., also
positive energy (and opposite momentum too). This is an antiparticle.
Thus the correspondence between the missing negative emerged electron and existing positron
means that, when missing a e of spin s, momentum p and energy E (E > 0) it corresponds
+
to a e of spin s, momentum p and energy E(E > 0). Note also that a negative energy electron
in the initial state corresponds to a positive energy positron in the final state in the transition
process.

The solutions of positive energy are usually denoted by u, while those corresponding to negative
energy are denoted by v. In order to work with positive energy solutions always (electron and
positron) it is convenient to relabel the spinors w:

p, s) w(
u( p, s)
p, s) w(
u( p, s)
p, s) w(
v( p, s)
p, s) w(
v( p, s)
(2.90)

The solutions associated with u(p, s) correspond to particles of 3-momentum p, energy E =

p) + m and spin in the s-direction; while those associated with the v(


( 2 2 p, s) correspond

to states of 3-momentum p, energy E = ( p) + m and spin in the s-direction. Now,
2 2

using the antiparticle interpretation (the absence of a state with energy E, momentum p
and spin in the s-direction is equivalent to the presence of an antiparticle state with energy
E, momentum p and spin in the s-direction) we can conclude that v( p, s) corresponds to

antiparticles of 3-momentum p, energy E = ( p) + m and spin in the s-direction.
2 2

As i i = 2m (the sign denotes a positive or negative energy solutions), the following property
must hold for the spinors:

u( p, s) = v(
p, s)u( p, s) = 2m, ,
p, s)v( (2.91)

also:

1 5 s 1 5 s
p, s)u(
u( p, s) = ( p + m) , p, s)v(
v( p, s) = ( p m)
2 2
u(
p, s)u(
p, s) = p + m, v( p, s) = p m.
p, s)v( (2.92)
= =
2.4 Solutions of the Dirac equation. Particles and antiparticles 25

2.4.2 The spin operator and the spin projection operator

We want to construct a spin operator which gives the correct, or physical, spin for electrons and
positrons in the rest frame. In the rest frame u and v are certainly eigenstates of s/2 but as
we have defined v, this will have the sign inverted from the one we want. This can be solved
easily by adding a 0 , since its eigenvalues are 1 for energy solutions. Thus, the operator
we are looking for is written in the rest frame as (s ) 0 /2. It is easy to show (which is left as
an exercise) that:
1 0 = 1 5 s = 1 5 s where s (0, s)
(s ) (2.93)
2 2 2
Now, if we assume that the 4-vector defined at rest s = (0, s) transforms as a 4-vector under
Lorentz transformations, then 12 5 s represents the spin component along s in the rest frame
even if it is applied to a boosted spinor. This means that having a state at rest which is an
eigenstate of 5 s, 5 s = at rest, this equation remains invariant if we transform it by a
Lorentz transformation:

5 sat

rest
at rest
= at rest

5 s S 1 = S 1
5 S s S 1 = 5 s =
5 s = (2.94)

The two spin projection operators that filter out the two eigenstates of 1
2 5 s (or the
component of physical spin along s) are:

1 5 s
(s) = (2.95)
2
where in the expression s is a 4-vector such that at rest it is (0, s).
The helicity is defined as the spin operator in the direction s = p . In the rest frame:

1 0 = 1 5 s where s (0, p)
h = (p ) (2.96)
2 2
As has been seen before, the helicity along p in the rest frame can also be obtained by applying
the operator 12 5 s over a boosted solution if s is the boost of the polarization vector at rest
s = (0, p). The boost considered must be such that it goes from the rest frame to the frame
where the particle has: p = (E, p); or by (, ) = (E, |p|)/m
( ) ( )( ) ( ) ( )
s0 0 |
p|/m
= = = (2.97)
s 1 E/m

p|/m, pE/m).
and s = 0, thus s = (|
The two helicity projection operators that filter out the two eigenstates of the the component
of physical spin along p will be:

1 5 s
(p) = p|/m, pE/m)
with s = (| (2.98)
2
26
Chapter 2. Relativistic wave equations, for spin-1/2 particles. The Dirac equation

In the high energy limit, |


p| = E and s (E/m, | p|p/m) = p /m. Then, in the high energy
limit, the helicity operator applied to a positive energy solution u gives:

5 p/m 5 m/m 5
u= u= u (2.99)
2 2 2
applied similarly to negative energy it gives:

5 p/m 5 m/m 5
v= v= v (2.100)
2 2 2
We have thus found that the helicity operator, at the high energy limit (or when m=0), acts as
5 /2 for u(v):

1 1 1 1
p, p) = u(
hu( p, p) 5 su(
p, p) 5 u(
p, p) = u( p, p)
2 2 2 2
1 1 1 1
p, p) = v(
hv( p, p) 5 sv(
p, p) 5 v(
p, p) = v(p, p) (2.101)
2 2 2 2
The two helicity projection operators at the high energy limit (or when m=0) that filter out
the two eigenstates of the the component of physical spin along p will be:
1 5
(p) = (2.102)
2
where = 1 for energy solutions.
Let us now consider the following chirality operators:
1
PR (1 5 ) (2.103)
L 2
and define the chirally projected fermions R,L :

R = PR , L = PL (2.104)

so R = PL and L = PR . From what we have seen, at high energy (or m = 0), chirality PR
L
acts as helicity projectors for the +ve energy solutions, u; and it acts as helicity projectors
for the -ve energy solutions, v. Namely PR filters out right-handed helicity (+) when applied to
a fermion, and it filters out left-handed helicity (-) when applied to an anti-fermion. PL filters
out left-handed fermions and right-handed anti-fermions.
Chirality is not a conserved quantum number for massive particles.

2.5 P , T , C and CP T
2.5.1 Parity
We need to look for a S() when is the space inversion operation. In this case, the relation
S 1 S = is:

S 1 0 S = 0
S 1 i S = i (2.105)
2.5 P , T , C and CP T 27

A solution (except for an irrelevant phase) is S = 0 . This means that:



1 (x)

2 (x)
0
(x ) = (x) = with t = t and
x = x (2.106)

3 (x)
4 (x)
This shows that the particle and antiparticle have dierent parity. Applying this transformation,
for a example, to the explicit solution for a fermion of momentum p and spin up in the z-direction,
we obtain the solution that represents a spin up fermion with momentum p:

1
( )
I 0 0 i(Etxp)
0
(x ) = 1 (x) = E + m
p3
e

0 I E+m
p1 +ip2
E+m

1 1

0 i(Et +x p) 0 i(Et x
)
= E + m
p3
e
= E + m
p3
e

p
(2.107)
E+m E+m
p1 +ip2
pE+m
1 +ip2
E+m

with p =
p. As expected, parity inverts the momentum and maintains the spin.

2.5.2 Time reversal


We are now looking for a ST which satisfies ST1 ST = T , where T is the time reversal
operation (equal to I except for the element 00 = 1). Such a matrix does indeed exist, and is
given by: ( )
1 2 3 0 I
St = i = (2.108)
I 0

If we use this in (x ) = ST (x) to see how the solutions transforms (together with x0 =
x0 , x = x) we find that it transforms an electron, with spin up and momentum p, to a
positron, with spin down and momentum p. This is not really what we should expect: we
expect the flip of the spin and the 3-momentum, but not the change from a particle to an
antiparticle.
This is solved by considering the following transformation:

(x ) = T (x), with T = i 1 3 (2.109)

The price we have to pay with this definition is the loss of linearity; T is an anti-lineal operator
that does not belong to the Lorentz group spinor representation.

2.5.3 Charge conjugation

Next we turn to charge conjugation C. Its meaning is best seen by considering the coupling
of fermions to an external gauge field. According to the minimal coupling principle, the Dirac
equation reads:
i e A m = 0 (2.110)
28
Chapter 2. Relativistic wave equations, for spin-1/2 particles. The Dirac equation

(We will often use the covariant derivative D = + ieA .) It is clear that the equation
describing the coupling to antiparticles should be:

i c + e Ac mc = 0 (2.111)

Let us find the relation between and c . We complex conjugate the first equation:

i e A m = 0 (2.112)

Let us now multiply by a matrix C such that:

C C = (2.113)

and such that: C 2 = 1, C C = 1. Elementary manipulations allow us to show that:

iC C C eC CA C mC = 0 (2.114)

so, clearly,
c = C (2.115)

It can be seen that for Dirac fermions C = i 2 , which is an anti-lineal operator.

2.5.4 CPT
Let us try to combine CPT together:

(i 1 3 ) 0 (i 1 3 ) i 2 ( 0 (i 1 3 ))
(x0 , x) (x0 , x) (x0 , x) (x0 , x) (2.116)

Thus CPT transforms (x) i 2 0 (i 1 3 )(x) = i 5 (x)


Applying this to the electron solution with spin up and 3-momentum p, we see that it
exchanges particle antiparticle, flips the spin and keeps p, as expected.
It is interesting to note that x = x is a Lorentz transformation and in the spinor represen-
tation is implemented by SP T = 5 , which is the same as the one obtained by applying CPT,
up to a phase.
Thus the action of CTP on a Dirac field is just to flip the time and space coordinates
and transform the spinor accordingly. In the process, a particle-antiparticle flip occurs as well
since the t t also meant the sign of the energy changed. Thus, positrons are negative
energy electrons running backwards in space-time. CPT turns incoming particles into outgoing
antiparticles, while flipping the spin.
There is a general theorem that proves that if a theory is locally described by Hermitian H,
Lorentz invariant and obeys spin-statistics relations, then the H is invariant under CPT.
Chapter 3

From classical to quantum field


theory

3.1 Lagrangian formalism of classical mechanics


Newtonian classical mechanics is based on the equations of motion:
( )
d L L
. =0 (3.1)
dt qi qi
.
where L is the Lagrangian L(qi , qi ) = T V , T is the kinetic energy, V is the potential energy
and qi the coordinates.
The Lagrangian equation can be obtained by minimizing the action, S, along the path :

S Ldt (3.2)

Changing the path slightly:


qi (t) qi (t) + qi (t) (3.3)
and keeping the extremes invariant qi (t0 ) = qi (t1 ) = 0, we obtain:
t1 t1 ( ) t1 ( ( ) ( ) )
L . L d L d L L
S = Ldt = . qi + qi dt = . qi . qi + qi dt
t0 t0 qi qi t0 dt qi dt qi qi
t1 ( ( ) ) t1 t1 ( ( ) )
d L L L d L L
= . + qi dt + . qi = . + qi dt (3.4)
t0 dt q i q i q i t0 t0 dt qi qi
This shows that the equation of motion emerges from the principle of minimum action (in
. . . . .
its derivation we have used the fact that: qi = qi since: qi qi + qi and by definition:
. . .
qi qi + qi ). Classical mechanics chooses the path that minimizes the action.

3.2 Lagrangian formulation of classical waves


Consider the string of masses (m) of Figure3.1. The coupling between the masses is provided
by the tension, T, of the rubber band connecting them. In addition each mass is also connected
to a spring. The spring located at position xi has a restoring force ki , where i is the
perpendicular displacement with respect to its equilibrium position.
30 Chapter 3. From classical to quantum field theory

Figure 3.1: String of masses coupled by a rubber band


The potential energy of the rubber band (w.r.t. rest) is i T di , where di is the variation
of the length of the rubber band between consecutive masses (i 1 and i).
( )2
1
d = 2 + x2 x x (3.5)
2 x
The Lagrangian of this discrete wave is:
( ( )2 )
. 1 . 2 1 1 i i1
L(i , i ) = m i k2i T x (3.6)
i
2 2 2 x

In terms of a mass per unit length ( = m/x) and of the spring constant per unit length
( = k/x), we can rewrite the Lagrangian as:
( ( )2 )
. .2 2 T i i1
L(i , i ) = x i i (3.7)
i
2 2 2 x

Then, the equation of motion for i is:


( )
d L L .. (i i1 ) (i+1 i ) 2 i
. = i = i T = i + T (3.8)
dt i i (x)2 x2

The same equation of motion can be obtained more elegantly. The Lagrangian can be written
in its integral form: . .
L(, , /x) = dxL((, , /x) (3.9)
. ( )
2
where L is the Lagrangian density: 2 2 2 2 T
2 x
When we vary the function

(x, t) (x, t) + (x, t) (3.10)

but keeping (t1 , x) = (t2 , x) = 0, x and (t, ) = (t, ) = 0, t, the variation of the
action should be zero for the real motion.
3.2 Lagrangian formulation of classical waves 31

We can proceed as in the previous section:


( )
t2 L L . L
S = dt dx + . () + (x ) (3.11)
t1 (x )
. .
as () = t ( + ) t = and similar for (x ), we can write:
( )
t2 L L L
S = dt dx + . t () + x ()
t1 (x )
( ( ) ( ) ( ) ( ) )
t2 L L L L L
= dt dx + t . t . + x x
t1 (x ) (x )
( ( ) ( ) )
t2 L L L
= dt dx t . x (3.12)
t1 (x )

since is arbitrary, the equation of motion that minimizes S(S = 0) should satisfy:
( ) ( )
L L L
= . + (3.13)
t x (/x)
. ( )
2
Now, if we take as the Lagrangian density L = 2 2 2 2 T
2 x the equation of motion
will be:
2 ..
= T (3.14)
x2
as expected. Setting T / = 1 and / = m2 this is just the Klein-Gordon equation in one
dimension:
.. 2
2 + m2 = 0 (3.15)
x
Extending the case we have just studied to 3 dimensions, we must just obtain the well-known
Klein-Gordon equation. In this case, the equation of motion will be:

..
3
2
0 = + m2 = (2 + m2 ) (3.16)
i=1
x2i

which should be obtained from the Lagrangian



1( )
L= d3 xL = d3 x m2 2 (3.17)
2
applying the equations:
( ) ( ) ( ) ( ) ( )
L L L L L L
= . + + = +
t x (/x) y (/y) ( ) z (/z)
(3.18)
In the general case in 3 dimensions and several fields i , i = 1, ..., n, we can write:

L= d3 xL, S = dtL = d4 xL(i , i ) (3.19)

And applying the principle of minimum action, we get the equations:


( )
L L
= , i = 1, ...n (3.20)
i ( i )
32 Chapter 3. From classical to quantum field theory

3.2.1 Canonical momenta


By analogy with point-like dynamics qi (t), for the field i (t, x), we can define its conjugate
momentum as:
L
i (x) . (3.21)
i
and then construct the Hamiltonian:
.

H= d3 x( i i L) d3 xH (3.22)
i
( )
To illustrate this, consider the Klein-Gordon Lagrangian density: L = 1
2 m2 2 .
In this case the conjugate momentum of the field is:

L .
i = . = (3.23)

and the corresponding Hamiltonian:


.

1 1 2 1 2 2 1 1 2 1 2 2
H= d3 x( L) = d3 x( 2 2 + () + m )= d3 x( 2 + () + m )
2 2 2 2 2 2
(3.24)

3.3 Noethers theorem


The relationship between symmetries and conservation laws is summarized in Noethers theorem.

3.3.1 Classical mechanics


In general, any transformation of the coordinates (for simplicity we consider only infinitesimal
transformations and no change on t):

qi (t) qi (t) = qi (t) + qi (t) (3.25)

is called a symmetry if it leaves the equation of motion invariant (the equation of motion for qi
is identical to the one obtained from the equation of motion of qi by changing qi qi ). This
means that qi (t) is also a solution of our problem which allows us to give another definition of
a symmetry: a transformation is a symmetry if it transforms solutions into solutions.
Under a general transformation, from the definition of the Lagrangian as L = T V , it is
clear that it transforms as a scalar:
.
.
L (qi , qi ) = L(qi , qi ) (3.26)

(from now on, for simplicity we will drop the sub-indices and the velocities, and write the previous
expression as just L (q ) = L(q)). If the transformation is a symmetry, both Lagrangians should
give rise to the same equation of motion. This is guaranteed if:

L (q ) = L(q ) (3.27)
3.3 Noethers theorem 33

(they are equal from the functional point of view). In fact, adding a total time derivative of a
function of the coordinates to the Lagrangian:
. d f .
f= f (q) = q (3.28)
dt q

does not change the equations of motion because:


( . ) . ( ) ( . ) .
d (L + f ) (L + f ) d L L d f f
. = . + .
dt q q dt q q dt q q
( ) ( ) ( )
d L L d f 2f . d L L
= . + 2q = . (3.29)
dt q q dt q q dt q q

Thus, the requirement given in 3.27 for a transformation of the coordinates to be a symmetry
must contemplate this last possibility and finally it should be expressed by:

d
T is a symmetry L (q) = L(q) + f (q) (3.30)
dt
Noethers theorem proves that any transformation of a symmetry has a conservation law asso-
ciated with it.
Defining L L(q) L (q) and taking into account that for any transformation L (q ) =
L (q + q) = L(q), in general we have:

L = L(q) L (q) = L(q) L(q q) = L(q + q) L(q) = L(q ) L(q) (3.31)

thus:
L L .
L = L(q ) L(q) = L(q + q) L(q) = q + . q
q q
( ) ( ) ( )
L d L d L d L
= q + . q . q = . q (3.32)
q dt q dt q dt q

If the transformation is a symmetry we have:

d
L = L(q) L (q) = f (q) (3.33)
dt
which combined with the previous result, leads to the conserved quantity:
( )
d L L
. q + f (q) = 0 . q + f (q) = Constant (3.34)
dt q q

Example: The harmonic oscillator

The Lagrangian, equation of motion and solution of the harmonic oscillator are:
.
L(x) = 21 mx2 12 kx2
..
mx + kx = 0

x(t) = A cos(t + )
34 Chapter 3. From classical to quantum field theory

Making the transformation:


x(t) x (t) = x(t) + (3.35)

the new Lagrangian, transformed equation of motion and transformed solution are:
. 2
L (x ) = L(x) = 12 x 12 k(x )2
..
mx + k(x ) = 0

x (t) = A cos(t + ) +

We observe that this is not a symmetry because: L (x ) = L(x ) 21 k 2 +kx = L(x )+df (x )/dt.
In fact, the equation of motion is not invariant and does not transform solutions into solutions.

Example: Free fall

The Lagrangian, equation of motion and a solution of the free fall are (taking y as vertical and
x as horizontal):
.2
L(x, y) = 21 mx mgy
.. ..
mx = 0 , my + mg = 0

x(t) = (vx t, 12 gt2 )

Making the transformation:


(x, y) (x , y ) = (x + , y) (3.36)

the new Lagrangian, transformed equation of motion and transformed solution are:
. 2
L (x , y ) = L(x, y) = 21 mx mgy
.. ..
mx = 0 , my + mg = 0

x (t) = (vx t + , 12 gt2 )

This is a symmetry because L (x , y ) = L(x , y ) (which means, as we can see, that the equation
of motion is invariant and, as expected, transforms solutions into solutions).
The conserved quantity is:
L . .
Constant = . (x) = mx mx = C (3.37)
x
Following this example, let us now consider the transformation:

(x, y) (x , y ) = (x, y + ) (3.38)

the new Lagrangian, transformed equation of motion and transformed solution are:
. 2
L (x , y ) = L(x, y) = 21 x mg(y )
.. ..
mx = 0 , my + mg = 0
3.3 Noethers theorem 35

x (t) = (vx t, 21 gt2 + )


which is also a symmetry because L (x , y ) = L(x , y ) + dt
d
f (x , y ) with f (x , y ) = mgt (which
means , as we can see, that the equation of motion is invariant and, as expected, transforms
solutions into solutions).
The conserved quantity is:
L . . .
Constant = . y + f (x, y) = my + mgt my + mgt = C my = mgt (3.39)
y

3.3.2 Field theory


In general, any transformation of the fields:

(x) (x) = (x) + (x) (3.40)

is called a symmetry if it leaves the equation of motion invariant. This is ensured if the action

is invariant under such transformations. As S = d4 L, any transformation that leaves the
Lagrangian density invariant up to a total derivative:

T is a symmetry L ( ) = L( ) + J ( ) (3.41)

and will have a conserved associated current. Let us proceed to obtain this current.
Defining L L() L () and taking into account that for any transformation L ( ) =
L ( + ) = L(), we have in general:

L L() L () = L() L( ) = L( + ) L() = L( ) L() (3.42)

thus
L L L L
L = L( ) L() = L( + ) L() = () + ( ) = () + ()
( ) ( )
( ) ( ) ( )
L L L L
= () + () () = () (3.43)
( ) ( ) ( )
where we have used the equation of motion to cancel out some of the terms.
If the transformation is a symmetry, from 3.41 we have:

L = L() L () = J () (3.44)

which combining with 3.43 leads to:


( )
L
+ J =0 (3.45)
( )
This result states that
L
j
+ J (3.46)
( )
is a conserved current. This conservation can also be expressed by saying that for charge Q:

Q= d3 x j 0 (x, t) (3.47)

is a constant in time:

dQ d j i
= d xj (x, t) =
3 0
d x i (x, t) =
3
d x
3 j = ds j = 0 (3.48)
dt dt V V x V S
36 Chapter 3. From classical to quantum field theory

3.3.3 Space-time symmetries and the energy-momentum tensor


Noethers theorem also applies to space-time transformations that can be interpreted as trans-
formations of the fields. For simplicity we will consider a scalar field theory where (x ) = (x)
and x = T (x) (thus, (T (x)) = (x) (x) = (T 1 (x))).
Let us consider displacements in space-time:

x x = T (x) = x + a, ( T 1 (x + a) = x T 1 (x) = x a )
= (x) (x) = (T 1 (x)) (x) = (x a) (x) = a (x) + ..... (3.49)

Under such a transformation, the Lagrangian density is not invariant, but rather changes as:

L = L( (x)) L((x)) = L(x a) L(x) = a L = (a L) = J (3.50)

This is a total derivative which ensures that the transformation is a symmetry. The conserved
currents (from 3.46) is:
( )
L L L
j = + J = (a (x)) + a L = (x) g L a T a
( ) ( ) ( )
( ) (3.51)
where T = ( ) (x) g L is the energy-momentum tensor, which has 4 conserved
L

currents associated with it ( T = 0), each one with the corresponding conserved charges:

P
d3 xT 0 (3.52)

For = 0 we obtain: ( )
L .
0
P = 3
d xT 00
= 3
d x . L =H (3.53)

just the Hamiltonian; and for i = 1, 2, 3 we have:

L
Pi = d3 xT 0i = d3 x . i = d3 x i P = d3 x
(3.54)

Lorentz transformations are slightly more complicated to implement. In this case, the corre-
sponding transformations are:
1
x x = x = x + a (M ) x + ... = x + a x + ... (3.55)
2
where we have used the fact that 12 a (M ) = 21 a (g g g g ) = 12 (a a ) = a .
The corresponding inverse transformation is:
1
(1 ) x = x a (M ) x + ... = x a x + ... (3.56)
2
Taking as an example the Dirac spinor field, where we know that the spinors transform as
(x ) = S()(x), so for each of the four field components we have:

1
(x) (x) = S() (1 x) = ( + a (B ) + ...)( (x a x + ...) )
2
1
= ( + a (B ) + ...)( (x) a x + ...)

2
1
= (x) a x + a (B ) + .... (3.57)
2
3.3 Noethers theorem 37

where M and B are the matrices obtained in section 2.2. The corresponding conserved
Noether current can be expressed as:

M = 0 (3.58)

with
L
M = T x T x (B ) . (3.59)
( )
Note the anti-symmetry in , . The corresponding conserved charges are:

M = d3 xM 0 (3.60)

The i, j components correspond to rotations (boosts correspond to the 0, i components). Thus,


the corresponding charge is just angular momentum! The actual relation is:
1
J i = ijk Mjk . (3.61)
2

3.3.4 Continuous global symmetries.


Very often Lagrangians are invariant under global transformations of the fields; this gives us
new conservation laws. As an example, let us take as a Lagrangian density:

L = + m 2 (3.62)

which is invariant under a global phase transformation:

ei , ei = +i, = i (3.63)

This leads to a conserved current:


L L
j = + = (i) + (i ) = i (( ) ( )) (3.64)
( ) ( )
As is arbitrary , we take j = i (( ) ( )). Later we will see that in this case, the
. .
conserved charge Q = d3 xj 0 = d3 xi( ) will represent the electric charge.
More generally, let us now assume that we have a set of N fields, i . In particle physics,
we are mostly concerned with transformations of the fields that are linear and unitary. Such a
symmetry takes the form:

= eia T a
i = ia (T a )ij j (3.65)

The T a are a set of Hermitian matrices, that may be traceless or not, which are generators of
some symmetry group G. If the group of transformations is continuous:

[T a , T b ] = if abc T c . (3.66)

The constants f abc are the structure constants of the group. The parameters a are real.

Then i i = i ia (T a )ij j , and the Noether currents are:


L
j a = i (T a )ij j (3.67)
( i )
38 Chapter 3. From classical to quantum field theory

Notice that j 0a = ii (T a )ij j . Taking the quantization rules introduced in the next section,
it is easy to prove:
[Qa , Qb ] = if abc Qc (3.68)

with Qa = d3 xj 0a

3.4 Quantum fields


It is worth emphasizing that we are dealing with fields as natural variables. In point-like dy-
namics, the natural variable is x(t). In field theory, we are dealing with objects extended over
space and time (such as an electromagnetic field A (x), or a matter field describing the electron
degrees of freedom, for instance (x)).

3.4.1 Canonical quantization rules


To quantize the fields, our present natural variables, we can proceed by analogy with QM. Recall
that from the postulates of QM one imposes that

[xi (t), pj (t)] = iij (3.69)

In field theory, at any instant t we have a continuous infinity of position variables (t, x),
labeled by x; and likewise a continuous infinity of momentum variables (t, x).
To obtain the canonical relations of the field theory, let us consider the 1-dimensional Klein-
Gordon field represented by the continuous Lagrangian:

1( )
L(, ) = dxL(, ) = dx (t )2 (x )2 m2 (3.70)
2
or its discrete version;
( ( )2 )
. 1 .2 i i1
L(i , i ) = x i m2 2i (3.71)
i
2 x

To quantize the discrete version we just have to follow the QM prescription, finding the canonical
momenta corresponding to i :

L .
i = . = x i , i = 1, .., n (3.72)
i

and then regard i and i as operators that satisfy the quantization condition:

[i (t), j (t)] = iij


[i (t), j (t)] = 0
[i (t), j (t)] = 0 (3.73)

From the expression obtained for i , we deduce:


. ij
[i (t), j (t)] = i (3.74)
x
3.4 Quantum fields 39

which, in the continuous limit, becomes


.
[(t, x), (t, x )] = i(x x ) (3.75)

Taking the continuous Lagrangian, the conjugate momenta is given by equation 3.23:
L .
= . = (3.76)

which, combined with equation 3.75, gives us the canonical commutation relations of field theory
(also called equal time commutators, E.T.C.):

[(t, x), (t, y)] = i(x y)


[(t, x), (t, y)] = [(t, x), (t, y)] = 0 (3.77)

Extending to 3 dimensions and several fields, the canonical commutation relations are:

[ (t, x), (t, y )] = i 3 (x y )


[ (t, x), (t, y )] = [ (t, x), (t, y )] = 0 (3.78)

3.4.2 Second quantization


What is the set of coordinates for a 1D string? One obvious choice is the displacement at each
point x, named (x), and the string can be quantized as explained in the previous section.
A more convenient choice is to describe the configuration of the string as a superposition of
normal modes (at a given time):

(x, t) = ap (t)eipx + h.c. (3.79)
p

where we have added the Hermitian conjugate to make it real. This is just the Fourier transform
of (x). Each normal mode is specified by the wavelength 2/p, and the system is equivalent
to a set of decoupled harmonic oscillators, each one corresponding to a specific normal mode,
with a frequency depending on p. To illustrate this, let us take the Klein-Gordon equation in
momentum space (1D):
( )
(t2 x2 + m2 ) = 0 t2 + (m2 + p2 ) (p, t) = 0 (3.80)

Comparing it with the equation for a harmonic oscillator, (d2 /dt2 + w2 )x = 0, shows that each
normal mode oscillates with w2 = m2 + p2 and independently of the others. The relative weight
of each normal node is given by ap .
Guided by this analogy, we can follow an alternative path to quantize the string: quantizing
the string is equivalent to quantizing each harmonic oscillator associated with each normal node
p.
As shown in Appendix B, the eigenstates of a harmonic oscillator can be labeled by a
quantum number n. These states that can be obtained by the action of the operator a , called
the creation operator, over the vacuum state:
1
|n >= (a )n |0 > (3.81)
n!
40 Chapter 3. From classical to quantum field theory

where these operators satisfy the commutation rules:

[a, a ] = 1, [a , a ] = 0, [a, a] = 0 (3.82)

The idea now is to use the quantum number of the harmonic oscillator labeled by p, to
describe the number of particles with momentum p. This is implemented by considering the
field of equation 3.79 as an operator, as well as the coecients ap (t) and ap (t) that will be
interpreted as the creation and annihilation operators associated with momenta p, as suggested
by the fact in the classical interpretation they give the relative weight of each normal node. This
approach seems to define the operators a up to a constant; but if we take as a normal mode the
one that represents only one particle in the entire volume, in agreement with what we want to
interpret, the operators a are completely identified and their commutation rules will be:

[ap , ap ] = pp , [ap , ap ] = 0, [ap , ap ] = 0 (3.83)

This alternative procedure to quantize the fields is normally called second quantization
because the quantized object is the wave function of standard QM. Other authors call it field
quantization in the sense that there is only one quantization step.
The time dependence of the Fourier transform of (t, x) (equation 3.79), thus the time
dependence of a, a , depends on the type of particle. In general, a oscillates with a time de-
pendence exp iEt, where E is a function of p. For a Schrodinger field, E = p2 /m, while for a
Klein-Gordon field E 2 = p2 + m2 . The dependence of E(p) reflects dierent types of string.
From what we have said, for a scalar field (x) there is one set of operators ap , ap corre-
sponding to each possible momentum p. A state with one particle of momentum p will be:

|p >= ap |0 > (3.84)

and examples of other states with more that one particle are:
1
|p, q >= ap aq |0 >, |p, p >= ap ap |0 > (3.85)
2!
In the case of spin-1/2 particles, two particles cannot occupy an identical state, in accordance
with the Pauli exclusion principle. This means that for a fixed p we desire that (ap )2 = 0. This
suggests that the way to incorporate the exclusion principle into the framework of creation and
annihilation operators is to replace the commutators with anti-commutators:

{ap , ap } = 0, {ap , ap } = 0, {ap , ap } = pp (3.86)

Now the number operator Np = ap ap that counts the number of particles of a given momentum
p, has only two eigenvalues, 0 and 1, as desired. This can be seen from the fact (we omit the
sub-index p):
N 2 = a aa a = a (1 a a)a = a a = N N (N 1) = 0 (3.87)

The two eigenstates are denoted by |1 >, |0 >, and the actuation of the creation and annihilation
operators over them is as expected:

N (a |0 >) = a aa |0 >= a (1 a a)|0 >= a |0 > |1 > = a |0 >


3.4 Quantum fields 41

a|1 >= a(a |0 >) = (1 a a)|0 >= (1 N )|0 >= |0 > a|1 > = |0 >
a |1 > = a a |0 >= 0
a|0 > = aa|1 >= 0 (3.88)
42 Chapter 3. From classical to quantum field theory
Chapter 4

Quantization of a scalar field

In this chapter we will proceed to quantize the scalar Klein-Gordon field. First, we will do it
following the second quantization method explained in section 3.4.2 with the introduction of
creation and annihilation operators. Then, we will use the standard canonical procedure and
prove that the methods are two dierent alternatives to quantize such fields.

4.1 Second quantization


First, let us study the normalization for each normal mode of the Klein-Gordon equation (for
the moment we consider only positive energy solutions):

p (t, x) = N eipx = N ei(Etpx) (4.1)

requiring that it represents only ONE particle in the entire space. From equation ??, we know
that the density associated with this solution is:

= j 0 = 2EN 2 (4.2)

and the required condition, V d3 x = 1, implies that N = 1/ 2EV . Thus the normalized
normal mode of the Klein-Gordon field that represents one particle with momentum p is:
1
p (x) = eipx (4.3)
2EV
and the general field is a superposition of these modes with some coecients:
a(
p)
+ (x) = eipx (4.4)
p
2EV

where p is summed over all possible values with the periodic condition ( px = (2/L)nx , py =
(2/L)ny , pz = (2/L)nz , with ni Z).
Now we consider a(p) as operators. Thus (x) is also an operator and if we want it to be
Hermitian, we must add the Hermitian conjugate:

ap ap
(x) = + (x) + + (x) = eipx + eipx (4.5)
p
2EV 2EV
44 Chapter 4. Quantization of a scalar field

where, as said in section 3.4.2, we can identify ap , ap as the creation/annihilation operators with
the commutation rules:

[ap , ap ] = p,p , [ap , ap ] = 0, [ap , ap ] = 0 (4.6)

These are space-time-independent operators and clearly their time dependence is:

ap (t) = ap eiEt (4.7)

which allows us to write



a (t) ipx ap (t) ipx
(t, x) = p e + e (4.8)
p
2EV 2EV

Let us now follow another argument that will allow us to obtain an identical expression for
the (x) operator. At any given time, say t=0, a general classical real (Hermitian) field can be
expressed by:
( a( p) ipx p) ipx
a(
)
(0, x) = e + e (4.9)
p
2EV 2EV

Its time evolution is given by the Klein-Gordon equation and we know it can be eiEt . If we
want the field to be real, the only solution is:
( a(
p) p) +iEt ipx
a(
)
iEt i
(t, x) = e e p
x
+ e e (4.10)
p
2EV 2EV

The apparent second alternative that can be obtained by multiplying the first term by e+iEt and
the second by eiEt , represents the same solution, just re-labeling p p. Now, to quantize
1
the field, we have to identify the coecient of the mode 2EV e i
px as an creation/annihilation
operator of particles with momentum p .
To obtain the continuous version of equation 4.5, we take into account that the possible values
of p are on a cubic grid of V = L3 , each one occupying a volume dV = (2/L)3 = (2)3 /V .

This means that p dV (...) = d3 p(...) or:

V
d3 p(...) (4.11)
p

(2)3

Replacing the p
in equation 4.5 by its continuous limit, we obtain:

1 ( ipx ) d3 p V ( )
(x) = ap e + ap eipx = a p
e ipx
+ a ipx
p
e (4.12)
p
2EV (2)3 2EV

Very often, a new set of creation/annihilation operators are introduced which are those defined
above up to a constant:

ap 2EV ap , ap 2EV ap (4.13)

with their commutation rules:

[ap , ap ] = 2EV [ap , ap ] = 2EV pp = 2E(2)3 (


p p ) (4.14)
4.1 Second quantization 45

where in the last step we have used the relation:



d3 x ix(p p) V
p p) =
( 3
e = p ,p (4.15)
(2) (2)3

With this new operator equation, 4.12 can be expressed as:



d3 p 1 ( ipx ipx
)
(x) = a p
e + a p
e (4.16)
(2)3 2E

and in an explicitly covariant form as:



d4 p
(x) = (p0 )(p2 m2 )[ap eipx + ap eipx ]. (4.17)
(2)3

where we have used:



1 1
(p2 m2 ) = ((p0 )2 (
p2 + m2 )) = (p 0
p
2 + m2 ) + (p0
+ p2 + m2 ), (4.18)
2|p0 | 2|p0 |

with |p0 | = E = + p2 + m2 , an expression derived from the property of the -function:


n
(x xi )
(f (x)) = (4.19)
i=1
|f (xi )|

where xi , i = 1, ..., n are the roots of f (x) = 0.


From now on, we will stick to the discrete version of (x) given in equation 4.5 as it is easier
to interpret and manipulate. Nevertheless, to reach the continuous limit, we only have to take
into accountat any stepthe equations 4.11 and 4.15.

4.1.1 (x) as a creation operator


The field operator introduced in the last section has a set of desirable properties. As we will see,
it satisfies the Klein-Gordon equation and when operating over the vacuum, it creates a state
localized at x.
It is very easy to prove that it satisfies the Klein-Gordon equation, since the only space-
time dependence of (x) is in the exponentials of the normal modes, which we know satisfy the
Klein-Gordon equation:
1 ( ipx )
(2 + m2 )(x) = (2 + m2 ) ap e + ap eipx
p
2EV
1 ( )
= ap (2 + m2 )eipx + ap (2 + m2 )eipx = 0 (4.20)
p
2EV

To prove that (x)| > creates a state that is sharply localized at x, we will proceed by
similarity with QM, where the wave function of a sate | > is given by (x) =< x| >.
Following this, the wave function of (x)| >, at say, time t=0, is given by:

1 ( ipx ) eipx
< x |(0, x)| >=< x | ap e + ap eipx | >= < x |ap | > (4.21)
p
2EV p
2EV
46 Chapter 4. Quantization of a scalar field

where < x |ap | >=< x | 1


p >= 2EV eipx . (Notice the lack of time dependence; even though not
stated explicitly, we are working in the Heisenberg picture, where operators such as (x) are
time dependent; whereas the states do not evolve in time. Notice also that < x |ap | >= eipx ,

which shows the normalization used for the normal modes when considering the operators a
instead of a.)
Using this we obtain:
eipx
1 d3 p 1 ip(x x)
< x |(0, x)| >= eipx = e (4.22)
p
2EV 2EV (2)3 2E

Except for the factor 1/2E, this is just 3 (x x); in fact, this extra factor is nearly constant for
small (non-relativistic) p. We can therefore say that < x |(0, x)| > 2E 1 3
(x x) represents
a state sharply localized at x; namely (x) creates a particle at x when acting over the vacuum.

4.1.2 The causality issue


QFT is built on the same postulates as QM, but with fields and conjugate canonical momenta
replacing positions and ordinary momenta. A key ingredient, a priori absent in QM, is causality
or locality; i.e. the statement that information cannot propagate faster than light.
Taking into account that (x) it is actually a Hermitian operator acting in the Hilbert space
of the system, and thus (x) is an observable that creates a particle at x when acting over the
vacuum, we can pose the causality issue in the following terms: the value of the field at one
space-time point x cannot influence, or is independent of, any other point x , if they cannot be
connected by a velocity less than the speed of light; or equivalently, if (x x )2 = t2 x2 < 0
(space-like separation). Thus we expect:

[(x), (x )] = 0 if (x x )2 < 0 (4.23)

Actually, we expect that this should be the case if x0 = x0 , as this is the canonical commutation
relation. Physically, this will mean that we can set our initial data for a constant time slice.
Let us now compute the above commutator for any time. Writing: (x) = + (x) + (x),
where + ( = + ) contains only annihilation (creation) operators, the commutator is:

[(x), (y)] = [+ (x) + (x), + (y) + (y)] = [+ (x), (y)] + [ (x), + (y)] (4.24)

where:
1 1 1 1
[+ (x), (y)] = [ ap eipx , ap eip y ] = ei(pxp y) [ap , ap ]
2EV
2E V
2EV 2E V
p

p p p
,

1 ip(xy) d3 p
= e = eip(xy) D(x y) (4.25)
p

2EV (2)3 2E

and:
1 1
[ (x), + (y)] = [ ap eipx , ap eip y ]
2EV
2E V
p

p

1 ip(yx) d3 p
= e = eip(yx) = D(y x) (4.26)
p

2EV (2)3 2E
4.1 Second quantization 47

Finally:

d3 p
[(x), (y)] = D(x y) D(y x) = (eip(xy) eip(yx) )
(2)3 2E

d4 p
= (p2 m2 )(p0 )(eip(xy) eip(xy) ) (x y) (4.27)
(2)3

where we have used the same procedure used in equation 4.17. From the last expression for
(x y), we can see that it is Lorentz invariant, as is D(x y). This means that (x y) =
((x y)), where is any Lorentz transformation.
We can see that indeed, if x0 = y 0 , and changing p p in the second integral (D(y x)),
the commutator [(t, x), (t, y )] = (0, x y ) = 0; as expected from the canonical commutators.
Knowing now that (z) = 0 for any vector with z 0 = 0, (z = (0, z)), let us prove that
(z) = 0 for any z such that z 2 < 0. As mentioned before, since (z) is Lorentz invariant, if
we are able to find a Lorentz transformation from any space-like 4-vector z 2 < 0, to a 4-vector
z with z 0 = 0, we are proving the results mentioned before, since (z) = (0, z ) = 0.
A general boost in the direction of z by speed applied to z is:
( ) ( )( ) ( ) ( )
z 0 z0 z 0 |z| (z 0 |z|)
= = = (4.28)
|z | |z| z 0 + |z| (z 0 |z|)

as z 0 |z| for a space-like vector (z 2 = (z 0 )2 z2 0), we can always choose = z 0 /|z| 1 to


make z 0 = 0. Thus, (z) = (z ) = 0, giving us that:

[(x), (y)] = 0 if (x y)2 < 0. (4.29)

and causality, in the sense above indicated, is now manifest.

4.1.3 The Klein-Gordon propagator

Since [(x), (y)] is a c-number, we can write:

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


= D(x y) D(y x) (4.30)

where we can identify each term, one-to-one, since: 0|(x)(y)|0 = 0|+ (x) (y)|0 =
0|[(x)+ , (y) ]|0 = D(x y).
Let us assume that: x0 > y 0 ,

d3 p
(x y )0|[(x), (y)]|0 = (x y )
0 0 0
(eip(xy) eip(xy) )
0
(2)3 2Ep
( )
d3 p 1 ip(xy) 1 ip(xy)
= (x0 y 0 ) e |p0 =Ep + e |p0 =Ep (4.31)
(2)3 2p0 2p0

where we have used that the integral is invariant under p


p, in order to change the sign of
the second term.
48 Chapter 4. Quantization of a scalar field

H
Using the fact that dzf (z) = +2i singularities ak Res(f, ak ) (counterclockwise loop), the
integral can also be expressed as:

d3 p dp0 1
(x y )
0 0
eip(xy) (4.32)
(2)3 C 2i p2 m2

where C is a circuit containing the two poles at p0 = Ep = p2 + m2 closed through the
lowest half of the complex plane (p0 = i) since x0 y 0 > 0. We have used:
1 1 1 1 1 1 1 1 1
= 0 2 = 0 2 = 0 ( ), ( ).
p2 m2 (p ) (p2 + m2 ) (p ) Ep2 p Ep p0 + Ep 2Ep p0 Ep 2Ep p0 + Ep
(4.33)
Another way of obtain the integration circuit is to add a small and negative imaginary part and
declare that the poles lie at p0 = Ep i; and the integrating p0 over the real line.
Finally let us define the retarded propagator as:

DR (x y) (x0 y 0 )0|[(x), (y)]|0 = (x0 y 0 )(D(x y) D(y x))



d4 p i
= 4 p2 m2
eip(xy) (4.34)
C (2)

where p0 is integrated over the real line with the poles at p0 = Ep i. If x0 y 0 > 0, we
close the circuit through the lowest half of the complex plane and both poles contribute to the
integral; while if x0 y 0 < 0, we close the circuit through the upper half of the complex plane
and the integral is 0.
Likewise, we can consider an advanced propagator as:

DA (x y) (y 0 x0 )0|[(x), (y)]|0 = (y 0 x0 )(D(x y) D(y x))



d4 p i
= eip(xy) . (4.35)
C (2) p m
4 2 2

where p0 is integrated over the real line with the poles at p0 = Ep + i. If x0 y 0 > 0, we close
the circuit through the lower half of the complex plane and the integral is 0; while if x0 y 0 < 0,
we close the circuit through the upper half of the complex plane the two poles contribute to the
integral.
The Feynman propagator is defined as:

DF (x y) = (x0 y 0 )D(x y) + (y 0 x0 )D(y x)


= (x0 y 0 )0|(x)(y)|0 + (y 0 x0 )0|(y)(x)|0
0|T ((x)(y))|0 (4.36)

This means that DF (x y) = D(x y) if x0 > y 0 (only the pole at +E contributes), and
DF (x y) = D(y x) if x0 < y 0 (only the pole at -E contributes). This can be expressed by:

d4 p i
DF = eip(xy) . (4.37)
(2)4 p2 m2 + i

With this prescription, (p0 )2 = p2 + m2 i = Ep2 (1 i) and the poles are in p0 = Ep i


and p0 = Ep + i. Now, if x0 > y 0 , we have to close the circuit in the lower half-plane and
4.2 Canonical quantization rules 49

we pick only the contribution from the pole at p0 = +Ep , which is D(x y), as desired. The
opposite happens if y 0 > x0 ; in that case the non-zero contribution comes from the negative
energy solution.
The dierent prescriptions for the poles are summarized in Figure 4.1.

Figure 4.1: Dierent prescriptions for the poles

We can check that the retarded propagator verifies the following dierential equation:

(2 + m2 )DR (x y) = i (4) (x y). (4.38)

In fact, all propagators are related to solutions of this partial dierential equation (Greens
equation) and they are therefore called Greens functions. If we denote a generic propagator by
D(x y) (retarded, Feynman...), then:

(2 + m2 )D(x y) = i 4 (x y), (4.39)

and thus all of them in momentum space obey:

(p2 + m2 )D(p) = i, (4.40)

and:
d4 p i
D(x y) = eip(xy) (4.41)
(2) p m2
4 2

The dierence between the dierent Green functions (DA (x y), DR (x y), ...) lies in the
prescription of how to handle the poles appearing in the p0 integral. The retarded prescription,
as we have seen, includes both of them in the contour.

4.2 Canonical quantization rules


Now we will follow the standard or canonical procedure to quantize the Klein-Gordon field
applying the canonical commutation relations introduced in the previous chapter:

[(t, x), (t, y )] = i 3 (x y )


[(t, x), (t, y )] = [(t, x), (t, y )] = 0 (4.42)
50 Chapter 4. Quantization of a scalar field

where (x) is the canonical momentum that can be derived from the Lagrangian density by:

L
(x) = . (4.43)

We have seen that the Lagrangian density for the Klein-Gordon field is:
1
L(, ) = ( m2 2 ) (4.44)
2
and the canonical momentum of is:
L .
(x) = . = (4.45)

As we are, for the moment, considering a real classical field that corresponds to a Hermitian
operator (x) which should satisfy the equation (2 + m2 ) = 0, we can expand it in terms of
the normal modes as:

(x) = (ap ep (x) + ap ep (x) (4.46)
p


1
where ep (x) = 2EV eipx , with p0 = E = + p2 + m2 , and ap , ap are operators which do not
depend on x. For the time being, we do not know the commutation rules between these opera-
tors, but we can deduce that they follow the same rules as the creation/annihilation operators
introduced in the previous section (which explains the names used for them) when we demand
the fulfillment of the canonical commutation rules.
As ep (x) = iEep (x), ep (x) = iEep (x), the conjugate field is:
. .

.
(x) = (x) = (iE)(ap ep (x) ap ep (x) (4.47)
p

The orthonormality of the normal modes is given by:



1 0 0
d3 xep (x)ep (x) = ei(p p )x i(pp )x
0
d3 x
2V EE
i(p0 p0 )x0
e p,p
= d3 xei(pp )x = (4.48)
2V EE V 2E
p,p p,p
Similarly, we can obtain: d3 xep (x)ep (x) = e2iEt 2E and d3 xep (x)ep (x) = e2iEt 2E .
With these relations it is easy to prove:
ap 2iEt
ap
d3 xep (x)(x) = + e
2E 2E
a
a
d3 xep (x)(x) = i e2iEt
p
p
(4.49)
2 2

which allows as to write:



ap = d3 ep (x)(E + i)

ap = d3 ep (x)(E i) (4.50)
4.2 Canonical quantization rules 51

These last two expressions allow us to compute the commutators between the operators a, a ,
as we know the commutators between and . Doing so, we obtain the same results as in 4.6,
thereby demonstrating that the system can be regarded as a set of harmonic oscillators labeled
by p and ap , ap as the creation/annihilation operators. We can also demonstrate the inverse:
show that the commutators between the operators , are those given in 3.78, just assuming
the commutators between a, a as those given in 4.6. This shows the equivalence of the two
methods, canonical and second quantization, to quantize the fields.

4.2.1 The Hamiltonian operator


As shown in section 3.2.1, the corresponding Hamiltonian of the Klein- Gordon field is:
.

1 1 2 1 2 2 1 1 2 1 2 2
H= d3 x( L) = d3 x( 2 2 + () d3 x( 2 + ()
+ m )= + m )
2 2 2 2 2 2
(4.51)
Before expressing this in terms of the creation and annihilation operators, let us rewrite the
middle term of the last expression. Applying partial integration to it:

()(
)d
3
x=
2 d3 x (4.52)
L/2
since for each component we have (as an example I take the z-component): L/2 z z =
L/2 L/2 2 L/2 2

z |L/2 L/2 z2 = L/2 z 2 due to periodicity. This, plus the fact that ( +
..
m2 ) =
2 + m2 = 0 allows as to express the Hamiltonian as:

1 .2 1 .2 ..
H= d3 x (
2 + m2 2 ) = d3 x ( ) (4.53)
2 2
..
Taking = (E
p
2 )(a
ep
p + ap ep ) it is easy to prove that:
1 1 1
H= E(ap ap + ap ap ) = E(ap ap + ) = E(Np + ) (4.54)
2 p p

2 p

2

where Np is the number operator corresponding to the normal mode labeled by p. This means
that exciting a normal mode p by one quantum number, the energy of the system increases by

E = p2 + m2 .

Notice that in the expression for H, the term p 12 is infinite but constant. The solution to
this is to drop it, since adding or subtracting a constant to the Hamiltonian does not change
the dynamics. This can be implemented using the normal product symbol:
1
H= E : (ap ap + ap ap ) := Eap ap = ENp (4.55)
2 p p
p

where the normal product means moving the operator a in each term to the left. It is understood
that when the normal product is taken, the operators a, a are not reordered by commutation
(for example : a aa := a a a, and not : a aa :=: a a a a := a a a a).
Finally, let us compute the energy of the state |p >= ap | >:

< p|H|
p >=< p| ENp |
p >=< p| Ep p,p |
p >= Ep (4.56)

p
p

as expected.
52 Chapter 4. Quantization of a scalar field

4.2.2 The total momentum operator

In section 3.3.3 we have seen that the symmetry of the Lagrangian under displacements in
space-time allows us to define 4 conserved charges P :

( )
L .
P = 0 3
d xT 00
= d x 3
. L = d3 xH = H (4.57)

just the Hamiltonian, and for i = 1, 2, 3 we have:

L
Pi = d3 xT 0i = d3 x . i = d3 x i P = d3 x
(4.58)

Using the fact that e (x) = i


pep (x) and e
p (x) = i pep (x), it is straightforward to show:
p

1 1
P = p (ap ap + ap ap ) = p(Np + ) (4.59)
p

2 p

2

as in the case of the Hamiltonian, each quantum carries the momentum associated with its
normal mode p, and the additional infinite constant in the definition of P can be discarded
using the normal product:

1
P = p : (ap ap + ap ap ) := pap ap = pNp (4.60)
2 p p
p

4.3 Quantization of charged scalar fields


Let us take two Klein-Gordon fields as discussed up to now, 1 (x), 2 (x); both Hermitian and
with the same mass. Assuming no interaction between them, the total Lagrangian density is
just:
1( )
L(1 , 2 , 1 , 2 ) = 1 1 + 2 2 m2 (21 + 22 ) (4.61)
2
which leads to the appropriate equation of motion for each of them, as expected from equation
3.20.
Let as now define the non-Hermitian fields:
1
(1 + i2 )
2
1
(1 i2 ) (4.62)
2

both of which satisfy the Klein-Gordon equation, since 1 (x), 2 (x) do. The Lagrangian can be
written in terms of these new fields:
1( )
L = 1 1 + 2 2 m2 (21 + 22 )
2
1( )
= ( 1 i 2 )( 1 + i 2 ) m2 (1 i2 )(1 + i2 )
2
= m2 (4.63)
4.3 Quantization of charged scalar fields 53

Treating and as independent fields now, equation 3.20 applied to the last expression for L
allows us to deduce that both fields obey the Klein-Gordon equation:
( )
L L
= 0 ( + m2 ) = 0
( )
( )
L L
= 0 ( + m2 ) = 0
( )
(4.64)

Expressing both i , i = 1, 2 in terms of creation and annihilation operators:




i (x) = (aip ep + aip ep ) (4.65)
p

allows us to write and in terms of a new set of creation and annihilation operators:

1 1 ( )
= (1 + i2 ) = (a1p + ia2p )ep + (a1p + ia2p )ep ) (4.66)
2 2 p

Defining: ap 1 (a1
2 p + ia2p ) and bp 1 (a
2 1 p + ia2p ), this allows us to write the fields as:

(x) = (ap ep + bp ep )
p


(x) = (ap ep + bp ep ) (4.67)
p

and their conjugate fields are:

L . 1 . . 1
.= = (1 i 2 ) = (1 i2 )
2 2
L . 1 . . 1
. == (1 +i 2 ) = (1 + i2 ) = (4.68)
2 2

The equal-time commutators among the new fields and their conjugates are:

[(t, x), (t, x )] = i 3 (x x )


[ (t, x), (t, x )] = i 3 (x x ) (4.69)

and all the rest are 0. This can be shown using the equal-time commutators of the original fields
( [i (t, x), i (t, x)] = i 3 (x x ), i = 1, 2 and 0 the rest). Similarly, the commutation relations
among the new creation and annihilation operators can be obtained using the commutators of
the original ones. This leads to:

[ap , ap ] = pp , [bp , bp ] = pp , (4.70)

and 0 for all the rest.


54 Chapter 4. Quantization of a scalar field

4.3.1 Noether current: the total charge


In section 3.3.4, we considered the case when the Lagrangians are invariant under global trans-
formations of the fields; which gives us new conservation laws.
In the example we are considering now, the Lagrangian density L = + m2 is
invariant under a global phase transformation:

ei , ei = +i, = i (4.71)

or its equivalent: ( ) ( )( )
1 cos sin 1
= (4.72)
2 sin cos 2
which is just a rotation in the 2-dimensional space spanned by the two Hermitian fields 1 , 2 .
In section 3.3.4, we saw that this symmetry leads to a conserved current:
( )
j = i ( ) ( ) (4.73)

and its associated charge is:


. .
Q= 3
d xj = 0
d3 xi( ) = ... = (ap ap bp bp ) = Nb Na (4.74)
p


with Na p ap ap , Nb p bp bp , where Na (Nb ) is the total number of particles generated by
a (b ). This allows us to interpret Q as the TOTAL CHARGE of the system, where a generates
-1 charged particles and b +1 charged particles (the antiparticles).

4.3.2 The total Hamiltonian


The Hamiltonian density is the sum of the two initial Hamiltonians:

H(1 , 2 , 1 , 2 ) = i i L = (1 1 L1 ) + (2 2 L2 ) = H1 (1 , 1 ) + H1 (2 , 2 ) (4.75)
i

Now, expressing each Hi in terms of the creation/annihilation operators, we have for the total
Hamiltonian:

H = d3 xH = d3 x(H1 + H2 ) = H1 + H2
1 1
= E(a1,p a1,p + ) + E(a2,p a2,p + )
p

2 p

2

= E(a1,p a1,p + a2,p a2,p + 1)
p


= E(ap ap + bp bp + 1) (4.76)
p

where we have used ap ap +bp bp = a1,p a1,p +a2,p a2,p . Notice the very important feature that both
particle types, charges (or particle/antiparticle), contribute positively to the total energy,
eliminating the problem of negative energy for antiparticles. As usual, the infinite constant term

1 can be eliminated by the normal product, finally obtaining:


p

H= E(ap ap + bp bp ) (4.77)
p

Chapter 5

Quantization of the Dirac equation

In this chapter we quantize the Dirac field using the second quantization method. As in the case
of the Klein-Gordon field, we do so by writing the field as a function of the normal modes, but
with two main dierences. First, knowing that we will describe a charged particle/antiparticle
system, and guided by what we saw in section 4.3 above, we will not require that the field
operator be Hermitian. Secondly, as suggested in section 3.4.2, we will replace the commutator
rules for the creation/annihilation operator by anti-commutators.

5.1 Second quantization


Assume that at a given time, let us take t=0, our system is in a given 4D state (0, x). The
spacial function of the 4 components can have any shape, but its time evolution will be dictated
by the Dirac Hamiltonian. So, at t=0, we can Fourier expand each of the components:
i
p
x
1 (0, x) C1,
p pe C1,p

2 (0, x) C eipx C2,p ipx
(0, x) =
(0,
= p 2,p =
C e
(5.1)
3 x)
p C3,p e
i
p
x
p
3,
p
i
px
4 (0, x) C4,
p pe C4,p

Now, at a given p, its 4-component spinor can be expressed as a linear combination of a base
set. As a base we are going to take the spinors up,s , vp,s introduced in section 2.4.1. Notice
that we use a minus sign for the vs for a reason that will become clear soon. Any other base set
can be used, for instance with the positive sign for the vs, but this selection is more appropriate
for our purposes. With this base:

C1,p

C2,p

C =
(ap,s up,s + bp,s vp,s ) (5.2)
3,p s
C4,p

where ap,s , bp,s are just the 4 expansion coecients. Thus at t=0, we can express any (0, x)
as

(0, x) = (ap,s up,s + bp,s vp,s )eipx = (ap,s up,s eipx + bp,s vp,s eipx ) (5.3)
p
,s p
,s
56 Chapter 5. Quantization of the Dirac equation

where we have just re-labeled p


p in the last term. Its time evolution is unique since u(v)
evolves with eiEt and gives:

(x) = (ap,s up,s eipx + bp,s vp,s eipx ) (5.4)
p
,s

Now we can proceed with the second quantization identifying the coecients a, b with cre-
ation/annihilation operators, except up to a possible constant. To fix it, we have to be sure that
the normal modes are normalized. In section 2.4 we normalized the spinors u and v such that:

= j 0 = u 0 u = u u = 2E, = j 0 = v 0 v = v v = 2E (5.5)

thus, the normalization of the normal modes p (t, x) = N up,s eipx implies:

1
1= 3
d x = 3
d x = N 2 up,s up,s d3 x = N 2 2EV N = (5.6)
2EV

Finally, the normalized normal modes are:


up,s ipx vp,s ipx
fp,s e , gp,s e (5.7)
2EV 2EV

and the proper field expansion is:



(x) = (ap,s fp,s + bp,s gp,s ) (5.8)
p
,s

where a, b are the creation/annihilation operators satisfying:

{ap,s , ap ,s } = p,p s,s , {bp,s , bp ,s } = p,p s,s (5.9)

and 0 for all the others (s, s can take the two values s once the quantization axis is fixed s)
The physical action of the operators a and b is as follows (also by comparison to the boson
case):

ap,s annihilates a positive energy state with momentum p and polarization s.

bp,s creates an antiparticle (of positive energy) with momentum p and polarization s (or
annihilates a negative energy state with momentum p and polarization s)

The (Dirac) vacuum will be defined by the relations:

a(
p)|0 = 0, b(
p)|0 = 0. (5.10)

The first defines the vacuum as the state where there are no positive energy solutions (and hence
none can be destroyed). The second relation defines the vacuum as the state where there are no
antiparticles (and hence none can be destroyed), which is equivalent to saying that the vacuum
is the state where negative energy solutions are all filled up (and hence none can be created, due
to the Pauli exclusion principle). This makes it clear that we are indeed dealing with fermions.
5.2 Canonical commutation relations 57

It is easy to prove that the normal modes satisfy:



d3 xfp,s fp ,s = p,p s,s

d3 xgp,s gp ,s = p,p s,s

d3 xfp,s gp ,s = d3 xgp,s fp ,s = 0 (5.11)

which allows us to write the creation and annihilation operators in terms of the fields:

ap,s = d3 xfp,s (x)

bp,s = d3 xgp,s (x) (5.12)

5.2 Canonical commutation relations


By proceeding backwards, the above anti-commutator rules imply that for fermionic fields the
canonical commutation relations have to be implemented by the following ETC:

{ (x), (y)} = { (x), (y)} = 0, x0 = y 0 (5.13)

{ (x), (y)} = i, (3) (x y ), x0 = y 0 . (5.14)


L
To deduce it, we need the expression of the momentum operator =
, where L is the
Lagrangian density of the Dirac field and runs over the four spinor components.
One Lagrangian that leads to the Dirac equation is

L= d3 xL = d3 x(i m) (5.15)

The Dirac equation should be obtained from the variation of the action S = d4 xL. Since
and can vary independently, let us start with the variation of :

S = d4 x (i m) (5.16)

and the principle of minimal action for the true equation of motion, which should be true for all
possible , implies that (i m) = 0 which is the Dirac equation. To obtain the equation
of motion for , first we can rewrite the action S by partial integration as

:S= d4 x(i m) (5.17)

where f f . Varying now , this leads to the equation of motion for :

(i m) = 0 (5.18)

In fact, the expression of the Lagrangian that appears in the action of equation 5.17 is just the
Hermitian of the one used in 5.15. Both of them give the same action because the dierence is
only a total derivative. To be proper, we can ensure the Lagrangian is Hermitian by adding the
two expressions obtained for L and dividing by 2:
(
)
1 1
L= (i m) + (i m) = (i m) (5.19)
2 2
58 Chapter 5. Quantization of the Dirac equation


where = ( ) and is the normal .
As all the Lagrangians shown in this section represent the same action, which gives the Dirac
equation, for simplicity we will use:

L = (i m) = i m = i( 0 ) m (5.20)

where runs over the 4-index spinor. The field conjugate to a will be:

L
= . = i( 0 0 ) = i , = i (5.21)

and the following equal-time anti-commutators can be evaluated in a straightforward manner:

{ (t, x), (t, y )} = { (t, x), (t, y )} = 0


{ (t, x), (t, y )} = , (3) (x y ) (5.22)

which implies 5.13 using the fact that i = ii .


Similarly from the anti-commutators 5.13, we can deduce the anti-commutator relations
between the creation and annihilation operators using 5.12 and equations 5.11, which again
shows the equivalence of the procedures to quantize the fields.

5.3 Noether currents


5.3.1 Space-time symmetries
The Dirac equation is invariant under time-space displacements x x = x +a , changing the
fields by (x ) = (x) (no change in the spinor part). It is immediately clear how to construct
the energy-momentum tensor associated with the Dirac Lagrangian: L(x) = i m.
Following the methodology used in section 3.3.3, we obtain:
( )
L
T
= (x) g L = i g L (5.23)
( )

where runs over the 4 spinor components. This energy-momentum tensor has 4 associated
conserved currents ( T = 0), each one with the corresponding conserved charges:

P d3 xT 0 = d3 xi g 0 L = d3 x g 0 L (5.24)

The total energy

The case = 0 in equation 5.24 gives:


.

P0 = d3 x L = d3 xH = H (5.25)

which is the total energy, as expected. Let us work on the Hamiltonian density H some more.
. . .
H = L = i (i m) = i (5.26)
5.3 Noether currents 59

where we have used the fact that when satisfies the Dirac equation the second term vanishes.
Another way to confirm this result is:
. .
H = i (i m) = i i 0 ( 0 0 + i i ) + m 0
= i 0 i i + m 0 = (i
+ m) (5.27)

where HDirac = i
+ m is the time evolution operator (i = HDirac ) introduced in
t
equation 2.1. Once more we recover:
.
H = i (5.28)
Now, using the relations given at 5.11 and the expressions:

(x) = (ap,s fp,s + bp,s gp,s )
p
,s
.
(x) = (iE)(ap,s fp,s bp,s gp,s ) (5.29)
p
,s

we can calculate the total energy H:


.
H= 3
d xH = d3 xi = ... = E(ap,s ap,s bp,s bp,s ) = E(ap,s ap,s +bp,s bp,s 1) (5.30)
p
s p
s

which, as expected, is just the sum of energy of all particles since: Npa,s = ap,s ap,s and Npb,s =
bp,s bp,s are the number operators for particle and antiparticle in state p, s. As in the charged
Klein-Gordon field, both particles contribute positively to the total energy and the infinite
constant term can be discarded using normal ordering:
.
H =: d3 xi :=: E(ap,s ap,s bp,s bp,s ) := E(ap,s ap,s + bp,s bp,s ) (5.31)
p
s p
s

Note that in this case a - sign appears when bp,s bp,s is replaced by its normal-ordered prod-
uct bp,s bp,s . In general, for FERMIONS, normal ordering acquires a -(+) sign if odd(even)
permutations are needed.

The total momentum

The case = i in equation 5.24 gives:



Pi = d3 x i P = d3 x
= d3 x (i)
(5.32)

since i is the momentum operator, this result can interpreted as the expectation value of
i for a given field . The same interpretation can be used for the total energy:


H= d3 x (i ) (5.33)
t
Substituting and in the expression for P given in equation 5.32, we obtain:

P = p(ap,s ap,s bp,s bp,s ) = p(ap,s ap,s + bp,s bp,s 1) (5.34)
p
s p
s

as expected. Again, we could remove the constant divergence term by normal ordering:

P = d3 x : (i)
:= p(ap,s ap,s + bp,s bp,s ) (5.35)
p
s
60 Chapter 5. Quantization of the Dirac equation

5.3.2 Internal symmetries

The Dirac Lagrangian L = (i m) also has a global U (1) symmetry, namely:

(x) (x) = ei (x), (x) (x) = ei (x) (5.36)

The corresponding Noether current can immediately be obtained:


L L
j = + = (i ) i + 0 = (5.37)
( ) ( )
as is arbitrary, the conserved current is finally defined as:

j = (5.38)

and the corresponding charge is:



Q= d3 xj 0 = d3 x (5.39)

which we have identified as a probability at 2.1.1, which is always positive. When we move to
quantum fields, making use of the expressions for and as functions of the creation and
annihilation operators, we obtain:

Q= d3 x = (ap,s ap,s + bp,s bp,s ) = (ap,s ap,s bp,s bp,s + 1) (5.40)
p
s p
s

Thus, the total number of particles minus the total number of antiparticles is a conserved
quantity. The physical interpretation of this symmetry depends on the assignment of particles.
It may correspond to electric charge, lepton number, baryon number, etc. As we can see,
particles and antiparticles have opposite values of this conserved charge.
As before, the additive infinite constant can be done away with through the use of normal
ordering:

Q= d3 x : := : (ap,s ap,s + bp,s bp,s ) := (ap,s ap,s bp,s bp,s ) (5.41)
p
s p
s

5.4 Locality
When discussing the Klein-Gordon field we found that locality, or microscopic causality, is guar-
anteed by the fact that [(x), (y)] = 0, if (xy)2 < 0. In the case of the Dirac field, the relevant
relations for causality are still commutators, but commutators between bilinear covariants be-
cause the physically relevant quantities are not the fields themselves, but the combinations which
are the sources of interaction: the bilinear covariants. Thus, we would like to prove that:

[ (x) (x), (x ) (x )] = 0, for (x x )2 < 0 (5.42)

which is the case if:

{ (x), (y)} = { (x), (y)} = 0,


{ (x), (y)} = 0 if (x y)2 < 0 (5.43)
5.4 Locality 61

Using these last relations, and assuming (x y)2 < 0, we can deduce:

(x) (x) (x ) (x ) = (x ) (x) (x) (x ) = (x ) (x ) (x) (x)


[ (x) (x), (x ) (x )] = 0 (5.44)

The first two anti-commutators of 5.43 are easy to prove since they do not contain any term like
{a, a } or {b, b }. The third one requires some care:
( ) ( )
{ (x), (x )} = { ap,s fp,s (x) + bp,s gp,s (x) , ap ,s fp ,s (x ) + bp ,s gp ,s (x ) }

p
,s ,s
p
( )
= (fp,s (x)) (fp,s (x )) + (gp,s (x)) (gp,s (x ))
p
,s
1 (
)
= (up,s ) (up,s ) eip(xx ) + (vp,s ) (vp,s ) eip(xx )
p
,s
2EV
1 (
)
= ( p + m) eip(xx ) ( p + m) eip(xx )
p

2EV
1 (
)
= (i + m) eip(xx ) (i + m) eip(xx )
p

2EV
1 ( ip(xx )
)
= (i + m) e eip(xx )
p

2EV
( )
d3 p ip(xx ) ip(xx )
= (i + m) e e
(2E)(2)3
= (i + m) (x x ) (5.45)

which proves the locality of the Dirac equation as (x x ) has this property.
Notice that we have proved the microscopic causality of the Dirac equation using the anti-
commutators for the creation/annihilation operators, while we used the commutator relations
for the Klein-Gordon equation. This is general: microscopic causality requires quantization
using anti-commutators for fermions and commutators for bosons.

5.4.1 The Dirac propagator


Similarly as for the Klein-Gordon field (eq 4.36), the Feynmam propagator for the Dirac field is
defined as:

SF (x y) = (x0 y 0 )0| (x) (y)|0 (y 0 x0 )0| (y) (x)|0 0|T (x)(y) |0,
(5.46)
where the minus sign appears on account of the anti-commuting character of the variables and
will become clear when we discuss interacting fields. Note that SF is a matrix.
The two contributions to this propagator come from:

S+ (x y) 0|(x) (y) |0


= 0| (ap,s fp,s (x) + bp,s gp,s (x) ) (ap ,s fp ,s (y) + bp ,s gp ,s (y) ) |0
p
,s ,s
p
62 Chapter 5. Quantization of the Dirac equation



= 0| (ap,s fp,s (x) ) (ap ,s fp ,s (y) ) |0 = fp,s (x) fp,s (y)
p
,s ,s
p p
,s
(up,s ) (up,s ) +ipy 1
= eipx e = ( p + m) eip(xy)
p
,s 2EV 2EV p

2EV

d3 p
= ( p + m) eip(xy) . (5.47)
(2)3 2Ep
and
(vp,s )
ipy (v ) d3 p
S (yx) 0|(y) (x) |0 = e p,s e+ipx = ( pm) eip(xy) .
p
,s 2EV 2EV (2)3 2Ep
(5.48)
We can now see that: 0|(x)(y)|0 propagates particles (positive energy solutions) from y to
x while 0|(y)(x)|0 propagates antiparticles from x to y.

Figure 5.1: Feynman prescription

The Feynman propagator can be expressed as (see Figure 5.1):


( )
d3 p
SF (z) = ( p + m)eipz (z 0 ) ( p m)eipz (z 0 )
(2)3 2Ep
( )
d3 p
= ( 0 E p + m)ei(Etpz) (z 0 ) ( 0 E p m)ei(Etpz) (z 0 )
(2)3 2Ep
( )
d3 p
= ( 0 E p + m)ei(Etpz) (z 0 ) ( 0 E + p m)ei(Et+pz) (z 0 )
(2)3 2Ep
( )
d3 p
= ( 0 E p + m)ei(Etpz) (z 0 ) + ( 0 E p + m)ei(Et+pz) (z 0 )
(2)3 2Ep

dp0 i( 0 p0 p + m)eip
0 z0
d3 p +ipz
= e
(2)3 (2) (p0 )2 E 2 + i

d4 p i( p + m)eipz
= (5.49)
(2)4 p2 m2 + i
The Feynman propagator contains both contributions too:

SF (x y) = (x0 y 0 )S+ (x y) (y 0 x0 )S (y x), (5.50)


5.5 Majorana fermions 63

but not simultaneously! If x0 > y 0 , we have particle propagation; while for y 0 > x0 , we have
antiparticle propagation.
As in the case discussed above for the Klein-Gordon field, all propagators are related to
solutions of the partial dierential equations:

(i x m)S(x y) = i (4) (x y). (5.51)

(Greens equation) and they are therefore called Greens functions.


We can easily solve Greens equation in momentum space:

( p m)S(p) = i (5.52)

and the result is:


i
S(p) = . (5.53)
p m
which actually means:
i p + m p + m
=i =i 2 . (5.54)
p m ( p m)( p + m) p m2
The dierence between the propagators:

d4 p p + m ip(xy)
S(x y) = i .e (5.55)
(2)4 p2 m2

lies in how we deal with the poles appearing in the p0 integral. The Feynman propagator is
obtained by considering:
p + m
i 2 (5.56)
p m2 + i

5.5 Majorana fermions


When discussing the neutral scalar field, we required it to be real (x) = (x). This is manifest
in the expansion of creation and annihilation operators:
1 ( ipx )
(x) = ap e + ap eipx (5.57)
p
2EV

showing that a particle coincides with its own antiparticle.


It is thus tempting to require a similar condition of fermion fields (x) = (x). This
imposes some restrictions on the transformation of these fields under the Lorentz group:

(x ) = (x ) S()(x) = S() (x) = S() (x) S = S (5.58)

and the requirement over S() is only verified in the Majorana representation. Now, there are
only two independent spinors that can be obtained by solving the Dirac equation ( p m)u = 0
in this representation, or boosting the normalized rest frame solutions:
( )
s
s = m eimt + h.c. (5.59)
2 s
64 Chapter 5. Quantization of the Dirac equation

finally obtaining the general solutions:


( p

)
E + m (1 + 2 E+m )s
s = p
eipx + h.c. up,s eipx + h.c. (5.60)
2 (2 E+m )s

Thus, the most general real field solution can be expressed as a combination of the previous
ones: ( )
up,s ipx up,s ipx
(x) = ap,s e + ap,s e (5.61)
p
,s 2EV 2EV

Following the second quantization prescription for this field, we must now interpret a, a as
creation/annihilation operator satisfying the usual anti-commutator rules for fermions:

{ap,s , ap ,s } = p,p s,s (5.62)

and 0 for all the rest.


Since neutrinos carry no charge, they could perfectly well be Majorana particles. We must
not confuse the fact that a fermion is a Majorana particle with the fact that we may use the
Majorana representation to describe ordinary Dirac fermions.
Chapter 6

Quantization of the photon field

A massive spin-1 field at rest has 3 degrees of freedom corresponding to the 3 eigenvalues of the
spin in the direction of quantization. In the case of a massless particle, such as the photon, the
number of degrees of freedom is reduced to two and they correspond to the two helicity states
(the spin component along the p-direction.

6.1 Maxwells equations


The electric and magnetic fields obey, in the absence of sources, Maxwells equations. Using the
rationalized Gaussian system of units, these equations are:


E
=0 ,
B
= 0,

+ B = 0 ,
E E = 0.
B (6.1)
t t

These electric and magnetic fields are expressed using the 4-vector potential A = (A0 , A),
is the vector potential. The expressions are:
where A0 is the scalar potential and A

= A
E 0 0 A,
=
B A,
(6.2)

The equation that this 4-vector A should obey to guarantee that the electric and magnetic
fields defined by 6.2 obey equations 6.1, will be made explicit soon.
Maxwells equations 6.1 can be expressed in a more covariant notation by introducing the
anti-symmetric field-strength tensor:

0 Ex Ey Ez

Ex 0 Bz By
F
E
(6.3)
y Bz 0 Bx
Ez By Bx 0

which can be expressed as F = A A :

F 0i = i A0 0 Ai = i A0 0 Ai = E i
F ij = j Ai i Aj = i Aj j Ai = ijk B k . (6.4)
66 Chapter 6. Quantization of the photon field

Then Maxwells equations read F = 0:

F 0 = i F i0 = i F 0i =
E
=0
. .
F j = 0 F 0j + i F ij =E j i jik B k =E j (
B)j = 0 (6.5)

and in terms of A , the equations of motion for A are:

F = 2A A = 0. (6.6)

6.1.1 The Canonical momentum conjugate


Maxwells equations can be derived from the following Lagrangian:

1 1 2 2
L= d3 xF F = d3 x (E B ) (6.7)
4 2
using the minimal action principle when varying the fields A :

1 1
S = d xL =
4 3
d xF F
= d4 x(F F + F F )
4 4

1 1
= d4 x(F F ) = d4 x(F ( A ) F ( A ))
2 2
= d4 xF ( A ) = d4 xF (A ) = d4 x( F )A (6.8)

since A is arbitrary, we get F = 0.


A peculiarity of this classical field theory is that since F 00 = 0, the canonical momentum
conjugate of A0 vanishes:
L
0 = .0 =0 (6.9)
A
This is just a constraint rather than a dynamical variable and indicates that A0 is a non-
independent variable. In fact, as E = A 0 0 A and E
= 0, this gives:
.
0=
E
=
2 A0
A (6.10)
.

Thus A0 can be expressed in terms of A:
.


A (x)
A (t, x ) =
0
d x 3
(6.11)
4|x x|
The space components are:
L
i = = F 0i = E i = E
(6.12)
( 0 Ai )

6.1.2 Gauge invariance


We immediately notice one problem in equation 6.6. Imagine we have found a solution A (x),
then A (x) + is a solution too.

2(A + ) (A + ) = 2(A ) (A ) + 2( ) ( ) = 0 + 0 (6.13)

Physics is not altered by the gauge transformation and one can define A uniquely by im-
posing an extra condition on A , called choosing the gauge or fixing the gauge. Commonly used
choices are:
6.1 Maxwells equations 67

Lorentz gauge: A = 0

Coulomb gauge:
A
=0

The Lorentz gauge is manifestly covariant while the Coulomb field gives A0 = 0 as deduced from
equation 6.11. In both cases equation 6.6 is reduced to:

2A = 0 (6.14)

thus each of the components of A satisfies the Klein-Gordon equation with m = 0 and the
solutions are:
A (x) = A0 eikx (6.15)
with k 2 = 0 and A00 = 0 for the Coulomb gauge.
From here on, we stay in the Coulomb gauge.

Coulomb Gauge

In the Coulomb gauge


A(x)
= 0, (6.16)
A0 = 0 (6.17)

and the solutions are



A(x) =A ei(k0 tkx)
eikx = A (6.18)
k k

with k 2 = 0 and k A
= 0 (from
k
A(x)
= 0 ). Then:
.
= A=
E ik 0 A(x)

=
B = ik A(x)
A (6.19)

which shows, since A is perpendicular to k, that E points in the direction of A


and B is
perpendicular to the plane defined by A and k, as seen in Figure 6.1.
We have seen that this Coulomb gauge implies k A = 0. Therefore A
k
can be written as a
k
linear combination of 2 unit vectors perpendicular to k:
= A + A
A (6.20)
k k,1 k,1 k,2 k,2

where k,i , 1 = 1, 2 are two unitary vectors perpendicular to k such that:

k,i k,j = ij
k,1 k,2 = k (6.21)

This fixes these vectors except for a rotation around k. We will leave it at that, except for k,
where the vectors are defined relative to k,i as shown in Figure 6.2.

k,1 = k,1 , k,2 = k,2 (6.22)

Thus, the solutions in this gauge can be written as:



A(x) = (Ak,1 k,1 + Ak,2 k,2 )eikx (6.23)
68 Chapter 6. Quantization of the photon field

Figure 6.1: Spatial configuration of the fields

Figure 6.2: polarization base


6.2 Quantizing the field A 69

6.2 Quantizing the field A


At this point we may write an expansion for the photon field, proceeding as we did for the
Klein-Gordon field. The most general expansion for a Hermitian field will be:
2 (
)

A(x) = ak,i k,i ek (x) + a k,i ek (x) (6.24)
k,i
k i=1


where ek (x) = eikx / 2V w, (w k 0 = |k|). Note that for a given k, we have only two degrees
of freedom, as expected for the photon. Thus, the quantization is obtained by considering ak,i
and a as operators with the standard commutators:
k,i

[ak,i , a ] = ij kk (6.25)
k ,j

and 0 for all the rest. a is the creation operator that corresponds to a photon with momentum
k,i
k and polarized linearly in the direction .
k,i
Very often it is more convenient to use circular polarized bases. This can be done by defining:
1
a (a ia )
k, 2 k,1 k,2
1
k, (k,1 ik,2 ) (6.26)
2
It is easy to prove that these new operators obey the rules:

[ak,s , a ] = ss kk , (s, s = ) (6.27)


k ,s

Now, taking into account that ak,+ k,+ + ak, k, = ak,1 k,1 + ak,2 k,2 , the field operator can be
rewritten as
( )

A(x) = ak,s k,s ek (x) + a k,s ek (x) (6.28)
k,s
k s=

where a is the creation operator that corresponds to a photon with momentum k and right-
k,+
handedly polarized with respect to k. This can be seen by looking at the time dependence of
the associated wave function:
( )
Re(k,s ek (x)) Re(k,1 ik,2 )eiwt ) k,1 cos(wt) + k,2 sin(wt) (6.29)

which corresponds to a photon of helicity +1 as we will see. Similarly for a .


k,
In 3-D space k,1 , k,2 , k will be a complete set and we can express it as:


2
(k, )i (k, )j + ki kj = ij (6.30)
=1

and it is easy to prove that:


2
2
(k,s )i (k,s )j = (k, )i (k, )j = ij ki kj (6.31)
s= =1
70 Chapter 6. Quantization of the photon field

6.3 Noether currents


6.3.1 Space-time symmetries
In section 3.3.3 above, we saw that the symmetry of the Lagrangian under displacements in
space-time allowed us to define 4 conserved charges P :
( )
L
P d3 xT 0 = d3 x Ai g 0 L (6.32)
(0 Ai )
For = 0 we obtain, as expected, the total Hamiltonian:
( ) .
L
0
P = d x3
0 Ai L = d3 x( i Ai L) = d3 xH = H (6.33)
(0 Ai )
Taking into account that the Lagrangian density is L = 14 F F = 1 2
2 (E B
2 ), the total
Hamiltonian is:
( . ) ( )
H = d3 xH = A
d3 x (E)
d3 x E L =
(E 0 ) 1 (E
+ A 2 B
2)
2
( ) ( )
1 2 2 0 = d3 x 1 (E
A 2 + B 2 ) (
E)A
0
= d3 x (E + B ) + E
2 2

1 2 + B
2)
= d3 x (E (6.34)
2
where we have not explicitly used that A0 = 0 for the Coulomb gauge, just to show that the
and B
result is valid for any gauge. Now, substituting E as a function of the field operator

A(x), it is straightforward to show that:



2
P0 = H = w(a ak, + 1/2) (6.35)
k,
k =1

where, as usual, the infinite constant term can be discarded using normal ordering.
The space components of P are given by:
.
k
Pi = d3 x k i Ak = d3 x(E k )(
i Ak ) = d3 x A i Ak (6.36)

Whereas:
B)
(E i = ijk E j B k = ijk E j (
A)
k = ijk klm E j
l Am

= (il jm im jl )E j
l Am = E j
i Aj E j
j Ai (6.37)

thus,

B)
d3 x(E i = d3 x(E j
i Aj E j
j Ai )
.j
.j
= d x( A
3 i Aj
j (E j Ai ) + (
j E j )Ai = d3 x A i Aj = P i

P = B
d3 xE (6.38)

just as predicted by classical electrodynamics. Substituting the momentum expansion, it is easy


to show:

2
P = k(a a + 1/2) (6.39)
k,
k,
k =1
6.3 Noether currents 71

6.3.2 The spin of the photon


In section 3.3.3 above, we saw that symmetry under Lorentz transformations implies, among
other things, that there are 3 conserved currents corresponding to the conservation of angular
momentum:
1
J i = ijk Mjk . (6.40)
2

with M = d3 xM 0 and

L
M = T x T x (B ) . (6.41)
( )

In our case:
A (x) A (x) = A (1 x) (6.42)

which means we should use (M ) instead of (B ) in 6.41:

L L
M = T x T x
(M ) A = T x T x (g g g g )A
( A ) ( A )
L L
= T x T x A + A (6.43)
( A ) ( A )

and
L L
M 0jk = T 0j xk T 0k xj
A + A
(0 A ) (0 A )
= ()
j A xk ()
k A xj + E k Aj E j Ak (6.44)
( ) ( )
Mjk = d3 x ()
j A xk ()
k A xj + E k Aj E j Ak = B)
d3 x (E j xk (E
B)
k xj

( ) (6.45)
and as ijk
xj (E B)
k = x (E
B)
, we finally obtain
i
( )
J = d3 x x (E
B)
(6.46)

This is the total angular momentum operator and it is a conserved charge. It is easy to show
that it has two parts: ( )
Ji = d 3 x E (x )
iA A)
+ (E i (6.47)

the first one can be identified with the total orbital angular momentum. By comparison with

the total momentum, P i = d3 xE we see that the momentum operator
i A, has been
substituted by the orbital momentum operator x .
Thus, the second term should be identified with the spin:

=
S A)
d3 x : ( E := ... = +i k(a ak,1 a ak,2 ) = k(a ak,+ a ak, ) (6.48)
k,2 k,1 k,+ k,
k k

which shows what was stated above: a (a ) is the creation operator that corresponds to a
k,+ k,
photon of helicity +1 (-1).
72 Chapter 6. Quantization of the photon field
Chapter 7

Interacting Fields

In a e+ e collision at S 3GeV the J/ particle can be produced. We can describe this
interaction as 2 initial plane waves, one for e+ and other for e , that overlap in some region. If
there is no interaction, nothing will happen; but if there is an interaction, the overlap can be
the source of a new wave, the J/ wave. The strength and the overlap itself, depend on the
kind of physics we are dealing with. Normally the strength is given by a constant called the
coupling constant and the overlap at x which is relevant to generate some other wave could
be:
e+ e , e+ e , e+ 5 e , .... (7.1)

The correct form in the case of e+ e J/ turns out to be:

c A (7.2)

where c is the coupling constant and A is the J/ field. This is the additional term that will
appear in the Lagrangian or Hamiltonian.

7.1 Photon-electron interaction term in the Lagrangian density


If there were no interaction between the photon and the electron, the total Lagrangian density
would be the sum of the two free Lagrangian densities:

1
L0 = Le + L = (i m) F F (7.3)
4
but if there is interaction, since L = T V , the obvious move would be to subtract the additional
potential energy caused by the interaction. To obtain the expression for the photon-electron
potential, let us move to the rest frame where the potential energy at x is:

A0 q d3 x = A0 (ej 0 )d3 x = ej A d3 x = e A d3 x Lint d3 x (7.4)

Thus, the Lagrangian density with an interaction is:

1
Ltot = L0 + Lint = (i m) F F + e A (7.5)
4
74 Chapter 7. Interacting Fields

Notice that the interaction term is Hermitian, since A and are Hermitian. Since Lint does
not contain any time derivatives, the conjugate fields stay as before:

Ltot L0
r = = , r = , A (7.6)
r r

and the new Hamiltonian density is just:

Ltot L0
Htot = r Ltot = r L0 Lint = H0 Lint H0 + Hint (7.7)
r r

The new equations of motion for the fields can be obtained using the principle of minimal action.
Varying A :

1
S = d4 xLtot = d4 x(Le + L + Lint ) = d4 x F F + d4 xe A
4
= d4 x( F )A + d4 xe A = d4 x( F + e )A (7.8)

where we have used 6.8. As A is arbitrary, the principle of minimal action implies:

F = ej (j = ) (7.9)

Now, varying :

S = d4 xLtot = d4 x(Le + Lint ) = d4 x( (i m) + e A )

= d4 x ((i m) + e A ) = 0
(i m) = e A (7.10)

these two equations indicate that the overlap between the electron and positron, given by j =
, acts as a source of the photon field with a coupling constant e; and the overlap of the
photon field and the electron (or positron) given by A acts as a source of the Dirac field with
the same coupling constant. Both sources come from Lint = e A

7.2 Representation of interactions


As explained in Appendix C, in a QM description, we can use several representations with the
same physical predictions. From here on we will only consider the case where the operators A
do not have an explicit dependence on t in the the Schrodinger representation.
Thus, in the usual, the Schrodinger, representation in QM, the operators do not evolve with
time, while the states do:

i | >S = HS | >S

i AS = 0 (7.11)

where the subscript S indicates that we are in this representation.


7.2 Representation of interactions 75

An alternative representation is that of Heisenberg where, as opposed to the previous case,


the operators evolve with time, while the states do not:

i | >H = 0

i AH = [AH , HH ] (7.12)

as before, the subscript H indicates the we are now in the Heisenberg representation. In the
very usual case that HS do not depend on t, we obtain HH = HS H
Up to now, even though we have not mentioned it explicitly, we have been working in the
Heisenberg representation, since the field operators (x) do depend on time and its commutator
with the Hamiltonian obey equation 7.12. Taking as as example the Klein-Gordon field operator:

(x) = (ap ep (x) + ap ep (x) (7.13)
p

and its total Hamiltonian (notice it does not depend on time):



H= Eap ap (7.14)
p

it is very easy to prove that the fields obey the equation of motion (in the Heisenberg represen-
tation):
.
i = [, H] (7.15)

and similarly for the Dirac field.


Now suppose we add an extra term to the original Hamiltonian HS = HS0 + hS . In this case
we can take the Dirac representation, also called interaction representation, where the operators
evolve with time with HI0 , while the states evolve using hI :
.
i | >I = hI | >I
.
i AI = [AI , HI0 ] (7.16)

as before, the subscript I indicates the we are now in the interaction representation. In the very
usual case that HS0 do not depend on t, we obtain HI0 = HS0 H 0 and hI = hS h.
Thus, if we consider operators and Hamiltonians (H 0 ) that do not have explicit time depen-
dence in the Schrodinger representation, the equations of motion can be summarized as:

Schrodinger:

i | >S = H| >S ,

i AS = 0 (7.17)

Heisenberg:

i | >H = 0,

i AH = [AH , H] (7.18)
76 Chapter 7. Interacting Fields

Dirac:

i | >I = h| >I

i AI = [AI , H 0 ] (7.19)

The important feature of the Dirac representation is that all the field operators satisfy the
same equation of motion of the free fields and thus their time dependence is the same. This
means that, for the Dirac field for example, its time dependence is still given by:

I (x) = (ap,s fp,s + bp,s gp,s ) (7.20)
p
,s

with
up,s ipx vp,s ipx
fp,s e , gp,s e (7.21)
2EV 2EV
and HI0

HI0 = E(ap,s ap,s + bp,s bp,s ) (7.22)
p
s

From now on we will stay in the interaction representation and we will omit the subscript
I.

7.3 The S matrix


In the interaction representation the evolution of the states is:
.
i| >= h| > (7.23)

in particular, each one of the basis states |


p1 >, |
p1 p2 >, ... at time t0 will evolve to become some
linear combination of the vector base at time t.
Let us define the time evolution operator as the operator that transforms states at time t0
to their evolved states at time t:

|(t) >= U (t, t0 )|(t0 ) > (7.24)

substituting this into the evolution equation for the states:



i| >= h| > i U (t, t0 )|(t0 ) >= hU (t, t0 )|(t0 ) > i U (t, t0 ) = hU (t, t0 ) (7.25)

In Appendix D, we deduce the expression for such an operator:


[ ( t )]
U (t, t0 ) = T exp i dt h(t ) , t > t0 (7.26)
t0

where T is called the time-ordered product and is defined as:


{
h(t1 )h(t2 ) if t1 > t2
T [h(t1 )h(t2 )] = T [h(t2 )h(t1 )] = (7.27)
h(t2 )h(t1 ) if t2 > t1
7.4 0 e+ e 77

Now, the S-matrix is defined by taking the limits:


[ ( )]
S = limt0 U (t, t0 ) = T exp i dth(t) (7.28)
t+

The advantage of this form is that the interaction Hamiltonian h is often proportional to
some dimensionless constant and the contribution from higher-order terms becomes progressively
smaller. Taking Htot = H0 + h = H0 Lint , the first order in S is:

S 1i dtT (h(t)) = 1i dth(t) = 1i dt 3
d x(Lint ) = 1+i d4 xLint (7.29)

Once the S-matrix is given, the transition amplitude from an initial state |i > to a final state
|f >, is given by
Sf i =< f |S|i > (7.30)

and the transition probability is just

|Sf i |2 = | < f |S|i > |2 (7.31)

7.4 0 e+ e
First we have to find out the possible expression for the interaction term between 0 , a particle
that happens to be a pseudo-scalar (a spin-0 field that changes sign under parity), and the
electron/positron field. As this is not an example of a weak interaction, the total Lagrangian
should be invariant under parity operations. In particular, the interaction Lagrangian should
contain the pion field, that changes sign under parity, and a bilinear term that must also change
sign under parity. So the interaction is:

Lint = ig 5 (7.32)

where g is a real coupling constant, and i is added to make it Hermitian:

Lint = (ig)( 5 ) = ig( 5 0 ) = ig 0 0 5 0 = ig 5


= ig(i 0 1 2 3 ) = ig(i 3 2 1 0 ) = ig(i 3 2 1 0 )
= ig(i 0 1 2 3 ) = ig 5 = Lint (7.33)

Notice that the coupling constant g is dimensionless. This is easy to see taking into account
that the action S is dimensionless, which implies that dim(L) = E 4 , and the massive terms of
the free scalar and Dirac Lagrangian tell us:

dim(m2 2 ) = E 4 dim() = E
3
dim(m) = 4 dim() = E 2 (7.34)

Thus, in order to have dim(ig 5 ) = E 4 , g must be dimensionless.


Now let us try to compute the transition amplitude 0 e+ e , where the 4 pion momentum
is going to be designated by P and the 4-momentum and spins of e e+ by (p1 , s1 ), (p2 , s2 )
respectively (as shown in Figure 7.1).
78 Chapter 7. Interacting Fields

Figure 7.1: pion decay diagram

Then the initial state is:


|i >= |P >= a |0 > (7.35)
P

and the final state is:


|f >= | p2 (e+ ) >= ap1s1 bp2s2 |0 >
p1 (e ) (7.36)

Notice that there cannot be any confusion between the creation operators for the pion and the
electron because the latter also contains the spin subscript.
We will compute the transition amplitude to first order in the coupling constant, where the
S matrix is given by equation 7.29:

S = 1 + i 4
d xLint = 1 + i d4 x(ig 5 ) (7.37)

So the transition amplitude is:


( )
Sf i =< f | d4 xg 5 |i > (7.38)

which can be computed using the momentum expansions of the fields:


( )
Sf i = < 0|bp2s2 ap1s1 d xg a |0 >
4 5
P

= g < 0|bp2s2 ap1s1 d4 x (aP eP + a eP ) (ap,s fp,s + bp,s gp,s ) 5
P
P p
,s

(ap ,s fp ,s +
bp ,s gp ,s )a |0 >
P
,s
p
(7.39)

the only term in 5 that survives is the term that annihilates a 0 with momentum P and
creates e+ (e ) with p2 , s2 (
p1 , s1 ) . Thus:

Sf i = g < 0|bp2s2 ap1s1 d4 x aP eP ap1 ,s1 fp1 ,s1 5 bp2 ,s2 gp2 ,s2 a |0 >
P

= g d4 x < 0|(bp2s2 ap1s1 aP )(ap1 ,s1 bp2 ,s2 a )|0 > eP fp1 ,s1 5 gp2 ,s2
P
7.4 0 e+ e 79


0 ,s 5 gp ,s
= g d4 x < ep+ 1 ,s1 P |P ep
2 ,s2 ep
0 +
1 ,s1 ep
2 ,s2 > eP
fp1 1 2 2


eiP x up1 ,s1 eip1 x 5 vp2 ,s2 eip2 x
= g d4 x eP fp1 ,s1 5 gp2 ,s2 = g d4 x
2P 0 V 2p01 V 2p02 V

g up1 ,s1 5 vp2 ,s2
= d4 xei(p1 +p2 P )x
(2P 0 V )(2p01 V )(2p02 V )
g up1 ,s1 5 vp2 ,s2 (2)4 4 (p1 + p2 P )
= (2)4 4 (p1 + p2 P ) M(7.40)
(2P 0 V )(2p01 V )(2p02 V ) (2P 0 V )(2p01 V )(2p02 V )

where M = g up1 ,s1 5 vp2 ,s2 is the so-called Lorentz-invariant matrix element.
Now, |Sf i |2 will be the probability of finding the final state |f > after an infinite time T
from the initial state |i >. Notice that this involves the square of a function. To deal with
this, we will use the trick introduced in section 4.1:
( )2
(2) (p1 + p2 P )
4 4
= (2) (p1 +p2 P )
4 4
dt d3 xei(p1 +p2 P )x = (2)4 4 (p1 +p2 P )T V
T V
(7.41)
To be more realistic, we want to compute the probability of finding the final state in a particular
range of momenta, [ p1 d
p1 /2, p1 + d p2 d
p1 /2] for the electron and [ p2 /2, p2 + d
p2 /2] for the
3
positron. Remember that we are considering discrete momenta, but these d p will still contain
many discrete ps, which allows us to use equation 4.11 to obtain:
V
= d3 p (7.42)
d
p p/2
(2)3

Using this, the probability we are looking for is:



p1 d
P rob( p1 /2, p2 d
p2 /2) = |Sf i |2
1 d
p 2 d
p1 /2 p p2 /2
( )( )
V V (2)4 4 (p1 + p2 P )V T
= d3 p1 d 3
p 2 |M|2
(2)3 (2)3 (2P 0 V )(2p01 V )(2p02 V )
(7.43)

which should be equal to the dierential decay rate d (the probability of decaying into the
momentum ranges p1 d p1 /2 and p2 d
p2 /2 per unit time) times the entire time, T, we are
considering from initial to final state. Thus:

(2)4 4 d3 p1 d3 p2 (2)4
d = 0
(p1 + p2 P ) 0 0 |M|2 d2 |M|2 (7.44)
2P 3 3
(2) 2p1 (2) 2p2 2P 0

d p1 d p2 3 3
where d2 4 (p1 + p2 P ) (2)3 2p0 (2)3 2p0 is called the 2-body Lorentz-invariant phase space.
1 2
Let us now compute the Lorentz-invariant matrix element (M) and this 2-body Lorentz-
invariant phase space. Assuming that we do not measure the spins of the final state, we simply
have to add the decay rate of these distinctive final spin states. Thus, the matrix element is:
( ) ( )
|M|2 = g up1 s1 5 vp2 s2 gup1 s1 5 vp2 s2
s1 ,s2 s1 ,s2
80 Chapter 7. Interacting Fields

( )

= g2 vp2 s2 5 up1 s1 up1 s1 5 vp2 s2 = g 2 vp2 s2 ( 5 ) up1 s1 up1 s1 5 vp2 s2
s1 ,s2 s2 s1
( )
= g 2 vp2 s2 5 ( p1 + m1 ) 5 vp2 s2 = g 2 (vp s ) 5 ( p1 + m1 ) 5 (vp2 s2 )
2 2

s2 s2
( )
( ) ( )
= g 2 (vp2 s2 ) (vp2 s2 ) 5 ( p1 + m1 ) 5 = g 2 ( p2 m2 ) 5 ( p1 + m1 ) 5

s2
( ) ( )
= g 2 ( p2 m2 ) 5 ( p1 + m1 ) 5 = g 2 T r ( p2 m2 ) 5 ( p1 + m1 ) 5 (7.45)

Since the pion mass is much greater than that of the electron (M = 135M eV, me = 0.5M eV ),
then:
( ) ( )
|M|2 g 2 T r p2 5 p1 5 = g 2 T r p2 5 5 p1 = g 2 T r ( p2 p1 ) = g 2 4p2 p1 (7.46)
s1 ,s2

We can compute p2 p1 in the center of mass (C.M.) where P = (M, 0), p1 = (p, +
p) and
p2 = (p,
p), with p = M/2. Thus:
M2
|M|2 = g 2 4p2 p1 = g 2 4(2p2 ) = g 2 4 = 2g 2 M 2 (7.47)
s1 ,s2 2

and the dierential decay rate is:

(2)4
d = d2 2g 2 M 2 (7.48)
2P 0
If we are interested in the total decay rate, we need to integrate over the whole of p1 , p2 . Since
in this case M is constant, we only need to compute the integral of the 2-body Lorentz-invariant
phase space, which gives:

|
p2 |
d2 = (7.49)
(2)5 2M
Finally, the total decay rate is:
( )
(2)4 (2)4 M/2 g2
+
( e e ) =
0
d2 |M| 2
= 2g 2
M 2
= M (7.50)
2P 0 s1 ,s2 2M (2)5 2M 8

7.5 Decay rates a 1 + 2 + ...n


We want to find a general expression for the decay of a particle a into n particles. Independently
of the type of particle and interaction, following on from the work in the previous section, it is
clear that the matrix element can be written as:
n
i=1 pi Pa )
(2)4 4 (
Sf i = M (7.51)
(2Pa0 V ) ni=1 (2p0i V )

since each particle in the final state will contribute via a term proportional to eipi x / 2p0i V

and the initial particle with eiPa x / 2Pa0 V , and the d4 x will produce the 4 in the above
expression. The expression is valid to any order of the coupling constant. The matrix element
7.5 Decay rates a 1 + 2 + ...n 81

M depends on the type of particle and the interaction and can be computed as in the previous
section, or systematically following the Feynmam rules that are given in the next chapter.
Thus, the probability we are looking for is:

p1 d3 p1 , p2 d3 p2 , ....) =
P rob( ....|Sf i |2
1 d3 p1 p
p 2 d3 p2

( V )
(2)4 4 ( i pi Pa )V T
= 3
d pi |M|2 (7.52)
i
(2)3 (2P 0 V ) i (2p0i V )

and the dierential decay rate can be written as:

(2)4 d3 pi
d = dn |M|2
, d n = 4
( p i Pa ) (7.53)
2Pa0 i i
(2)3 2p0i

Clearly the total width is such that:

dN/dt
aall = N (t) = N0 et N0 et/ (7.54)
N
where we have naturally defined the mean lifetime as = 1 .
A particle can have dierent decay channels. A useful concept is the branching ratio for a
given process:
aX
BraX = . (7.55)
aall
The activity is defined as the number of decays per unit time:

dN (t)

A(t) = = N0 et/ = A0 et/ (7.56)
dt
and for a specific decay, it is clear that:

A(t)aX = A(t)BraX = Cet/ (7.57)

Thus, measuring the number of decays for a specific channel as a function of time is enough to
obtain the lifetime or total width.
As an example, taking the measured value of the 0 = 1 tot and the measured value of
Br( 0 e+ e ) , we can obtain a value for ( 0 e+ e ) = tot Br( 0 e+ e ) and from this
determine the value of the coupling constant, g, in expression 7.50.

The reason for calling a total (or partial) width is clear. If we consider an unstable state, in
a state of well defined momentum, we can describe it with a wave function:

(x, t) = Ce 2 t ei(E0 tpx) .



(7.58)

It is an easy exercise to show that if we actually measure the energy of such a state, we will find
a Lorentzian distribution whose mean is E0 and its width .
82 Chapter 7. Interacting Fields
Chapter 8

QED

In this chapter we will proceed to compute some processes involving electrons and photons in
the framework known as QED. We will use the interaction or Dirac representation, where the
fields evolve like the free quantized fields introduced in the previous chapters for electrons and
photons, but considering now the interaction term between them given in section 7.1.
We will see that the procedure used to obtain such expressions can be systematized in what
are called Feynman rules.

8.1 e+ e 2: the electron propagator


In this section we want to compute the probability of the reaction:

e ( p2s2 ) (ka a )(kb b )


p1s1 )e+ ( (8.1)

knowing that the interaction density Hamiltonian between electrons and photons is e A .
First we will determine the amplitude of such a process and later its probability, which will be
expressed in terms of an area: the cross section.

8.1.1 Amplitude of the process e+ e 2


The initial and final states are:

|i > = ap1s1 bp2s2 |0 > |e e+ >


|f > = a a |0 > |2 > (8.2)
ka a kb b

where i denotes the polarization state of the final photons. Notice that we use the same letter
a to design the creation/annihilation operators for electrons and photons, but there cannot
be any confusion because the subscripts are dierent.
The amplitude is:


(i)n
Sif = < f |S|i >=< f |1 + dt1 ... dtn T (h(t1 )....h(tn ))|i >
n=1
n!

1
= < f |1 i dth(t) dt1 dt2 T (h(t1 )h(t2 )) + ...|i > (8.3)
2
84 Chapter 8. QED


where h(t) = d3 x(Lint ) = d3 xe : A := d3 xe : j : A (from now on we will
not explicitly write the normal ordering, but it will be taken into account in due course). The
process we are considering involves destroying two electrons and creating 2 gamma particles,
thus S should contain at least:
ap1s1 bp2s2 a a (8.4)
k a a k b b

but h only has products of 3 creation/annihilation operators. This implies that we must go to at
least second order to compute this amplitude. Thus, keeping only the first non-zero contribution:

1
Sif < f| dx0 dy 0 T (h(x0 )h(y 0 ))|i >
2
( )
e2
= < f| dx0 dy 0 T d3 xj (x)A (x) d3 yj (y)A (y) |i > (8.5)
2

but this approximation for S contains 6 creation/annihilation operators. The two extra oper-
ators, in addition to the four needed, account for the creation and annihilation of internal e
particles, as Figure attempts to indicate 8.1. The contribution to the calculation of this internal
line is called the Feynman propagator.

Figure 8.1: e+ e 2 diagram

The aim of this section is to derive the rules associated with external e , and internal e
lines, to compute Sif systematically from Figure 8.1. The same rules apply to other diagrams
that contribute to high orders on the coupling constant e.
As the time ordering is linear and j and A operate in orthogonal subspaces:

e2
Sif = < 2| d4 x d4 yT (j (x)A (x)j (y)A (y)) |e e+ >
2

e2
= d4 x d4 y < 2|T (A (x)A (y))T (j (x)j (y))|e e+ >
2

e2
= d4 x d4 y < 2|T (A (x)A (y))|0 >< 0|T (j (x)j (y))|e e+ > (8.6)
2
where in the last step we have used the fact that |e e+ > |0 > |e e+ >e and similarly for
|2 >.
8.1 e+ e 2: the electron propagator 85

Let us first evaluate the part. Defining (0, k,i ), we can rewrite equation 6.24 as:
k,i


2 ( )
A (x) = ak,i ek (x) + a ek (x) (8.7)
k,i k,i
k i=1

thus

< 2|A (x)A (y)|0 >=


( )( )

= < 0|aka a akb b k (ak ek (x) + ak ek (x)) k (ak ek (y) + ak ek (y)) |0 >
k k

=
ka a eka (x)kb b ekb (y) + kb b ekb (x)ka a eka (y) =< 2|A (y)A (x)|0 > (8.8)

and the term we are interested in is:

< 2|T (A (x)A (y))|0 > = < 2|A (x)A (y)|0 > (x0 y 0 )+ < 2|A (y)A (x)|0 > (y 0 x0 )
= < 2|A (x)A (y)|0 > ((x0 y 0 ) + (y 0 x0 ))
= ka a eka (x)kb b ekb (y) + kb b ekb (x)ka a eka (y) (8.9)

Notice that it is invariant under the changes , x y


Let us now evaluate the electron term < 0|T (j (x)j (y))|e e+ >.

< 0|T (j (x)j (y))|e e+ >=< 0|j (x)j (y)|e e+ > (x0 y 0 )+ < 0|j (y)j (x)|e e+ > (y 0 x0 )
= < 0| : (x) (x) : : (y) (y) : |e e+ > (x0 y 0 )
+ < 0| : (y) (y) : : (x) (x) : |e e+ > (y 0 x0 ) (8.10)

Notice that the second term is obtained from the first by changing , x y. As the photon
part is invariant under these changes, eectively we can write:

< 0|T (j (x)j (y))|e e+ >= 2 < 0| : (x) (x) : : (y) (y) : |e e+ > (x0 y 0 ) (8.11)

To simplify the notation, we can write:



(x) = (x) + 0 (x), ((x) aps fps (x), 0 (x) bps gps (x))
ps ps

(x) = 0 (x) + (x), (0 (x) , (x) 0 ) (8.12)

The superscript 0 indicates that it contains creation operations only. Thus, the term in
equation 8.11 can be given as:

: (x) (x) : : (y) (y) :=



: (0 (x) + (x))i ij ((x) + 0 (x))j :: (0 (y) + (y))k kl

((y) + 0 (y))l :=
[ ]

ij : i0 (x)j (x) + i (x)(x)j + i0 (x)j0 (x) + i (x)j0 (x) :
[ ]

kl : k0 (y)l (y) + k (y)(y)l + k0 (y)l0 (y) + k (y)l0 (y) :=
[ ]

ij i0 (x)j (x) + i (x)(x)j + i0 (x)j0 (x) j0 (x)i (x)
[ ]

kl k0 (y)l (y) + k (y)(y)l + k0 (y)l0 (y) l0 (y)k (y) (8.13)
86 Chapter 8. QED

but this term must be sandwiched by < 0| and |e+ e >, which implies that only the terms in the
first bracket with annihilation operators on the left survives (only the second term i (x)(x)j ).
In addition, the total eect must be to annihilate 1e and 1e+ , which is already done by this
surviving term in the first bracket. Thus, only terms in the second bracket that create and
annihilate the same particle will survive. Finally:

< 0|T (j (x)j (y))|e e+ >= 2 < 0| : (x) (x) : : (y) (y) : |e+ e > (x0 y 0 )
[ ]

= 2ij kl < 0|i (x)(x)j k0 (y)l (y) l0 (y)k (y) |e+ e > (x0 y 0 ) (8.14)

Let us work some more on this last expression, taking into account that it must be multiplied
by the photon term that is invariant under the change , x y:

< 0|T (j (x)j (y))|e e+ >=


[ ]

= 2 < 0| ij
kl i (x)(x)j k0 (y)l (y)(x0 y 0 ) ij kl i (x)(x)j l0 (y)k (y)(x0 y 0 ) |e+ e >
[ ]

= 2 < 0| ij
kl (x)j k0 (y)i (x)l (y)(x0 y 0 ) ij kl i (y)(y)j l0 (x)k (x)(y 0 x0 ) |e+ e >
[ ]

= 2 < 0| ij
kl (x)j k0 (y)i (x)l (y)(x0 y 0 ) ij kl i (y)l0 (x)k (x)(y)j (y 0 x0 ) |e+ e >
[ ]

= 2 < 0| ij
kl (x)j k0 (y)i (x)l (y)(x0 y 0 ) kl ij k (y)j0 (x)i (x)(y)l (y 0 x0 ) |e+ e >
[ ]

= 2ij kl < 0| (x)j k0 (y)(x0 y 0 ) k (y)j0 (x)(y 0 x0 ) i (x)(y)l |e+ e >
[ ]

= 2ij kl < 0| (x)j k0 (y)(x0 y 0 ) k (y)j0 (x)(y 0 x0 ) ( |n >< n|)i (x)(y)l |e+ e >
n
[ ]
=
2ij kl < 0| (x)j k0 (y)(x0 y )
0
k (y)j0 (x)(y 0 x ) |0 >< 0|i (x)(y)l |e+ e >
0
(8.15)

where we have used the fact that I = n |n >< n|, the index running over all the bases, but
only the term |0 >< 0| survives.
Thus, the matrix element:

e2
Sif = d4 x d4 y < 2|T (A (x)A (y))|0 >< 0|T (j (x)j (y))|e e+ >
2

e2
= d4 x d4 y < 2|A (x)A (y)|0 >
2 [ ]

2ij kl < 0| (x)j k0 (y)(x0 y 0 ) k (y)j0 (x)(y 0 x0 ) |0 >< 0|i (x)(y)l |e+ e >
(8.16)

has a very remarkable interpretation. When studying the Klein-Gordon field, we have seen that
(0, x) applied to |0 > creates a spin-0 particle at x at t = 0. Similarly, i (x)|0 >= i0 (x)|0 >
(i (x)|0 >= i0 (x)|0 >) can be interpreted as creating a positron (electron), the i component
of which is at x = (x0 , x). Also, (x) ((x)) can be interpreted as the annihilation of a positron
(electron) at x = (x0 , x).
So:

< 0|i (x)(y)l |e+ e > is the amplitude to kill the e+ at x and the e at y.

< 0|(x)j k0 (y)|0 > is the amplitude to create an e at y and kill it at x. The (x0 y 0 )
that multiplies this term indicates that the time ordering this happens is such that x is
later that y.
8.1 e+ e 2: the electron propagator 87

< 0|k (y)j0 (x)|0 > is the amplitude to create a e+ at x and kill it at y. The (y 0 x0 )
that multiplies this term indicates y happens later that x.

< 2|T (A (x)A (y))|0 > is the amplitude to create one at x and other at y.

All this is summarized in Figure 8.2.

Figure 8.2: Diagrammatic representation of the creation/annihilation processes in e+ e 2

Notice that:
[ ]
< 0| (x)j k0 (y)(x0 y 0 ) k (y)j0 (x)(y 0 x0 ) |0 >
[ ]
= < 0| (x)j k (y)(x0 y 0 ) k (y)j (x)(y 0 x0 ) |0 >= SFjk (x y) (8.17)

is just the Dirac propagator discussed in section 5.4.1. This, together with the use of the relation:

< 0|i (x)(y)l |e+ e >=< 0| bps (gps (x))i ap s (fp s (y))l ap1 s1 bp2 s2 |0 >= (gp2 s2 (x))i (fp1 s1 (y))l
ps p s
(8.18)
allows us to express equation 8.16 as:

e2
Sif = d x 4
d4 y < 2|T (A (x)A (y))|0 > 2ij kl SF (x y)jk (gp2 s2 (x))i (fp1 s1 (y))l
2
( )
e2
= d4 x d4 y ka a eka (x)kb b ekb (y) + kb b ekb (x)ka a eka (y) 2gp2 s2 (x) SF (x y) fp1 s1 (y)
2
= e 2
d x d4 ygp2 s2 (x) ka a SF (x y) kb b fp1 s1 (y)eka (x)ekb (y) + (a b)
4


vp2 s2 eip2 x d4 p ieip(xy) up1 s1 eip1 y eika x eikb y
= e2 d4 x d4 y ka a k + (a b)
2p02 V (2)4 p m + i b b 2p01 V 2ka0 V 2k 0 V
b

d4 p ei(p2 +pka )x ei(p1 pkb )y i
= e 2
d x4
d4 y vp2 s2 ka a k up s + (a b)
(2)4 (2p02 V )(2p01 V )(2ka0 V )(2kb0 V ) p m + i b b 1 1

d4 p (2)4 4 (p2 + p ka )(2)4 4 (p1 p kb ) i
= e2 vp2 s2 ka a k up s
(2) 4
(2p0 V )(2p0 V )(2k 0 V )(2k 0 V ) p m + i b b 1 1
2 1 a b
+(a b) (8.19)
88 Chapter 8. QED

integrating d4 p and using the second Dirac delta ( p = p1 kb ), we obtain:


(2)4 4 (p2 + p1 kb ka ) i
Sif = e2 vp2 s2 ka a k up s + (a b)
(2p02 V )(2p01 V )(2ka0 V )(2kb0 V ) p1 kb m + i b b 1 1

(2)4 4 (p2 + p1 kb ka )
M (8.20)
(2p02 V )(2p01 V )(2ka0 V )(2kb0 V )
where we have defined the invariant matrix element M as:
i
M = e2 vp2 s2 ka a k up s + (a b) (8.21)
p1 kb m + i b b 1 1

Figure 8.3: Lowest-order Feynman diagrams for e+ e 2


Notice that this invariant matrix element can be obtained from Figure 8.3 by applying the
following rules.

Draw the diagrams connecting e+ e to 2, using the only possible vertex ee up to the
desired order in the coupling constant.

Then assign the following factors to construct M:


(1) For an outgoing fermion particle with momentum p write u(p).
(2) For an outgoing fermion antiparticle with momentum p write v(p).
(3) For an incoming fermion particle with momentum p write u(p).
(4) For an incoming fermion antiparticle with momentum p write v(p).
(5) For an outgoing photon with momentum p write (p). When considering the helicity
basis, use (p).
(6) For an incoming photon with momentum p write (p).
(7) For the interaction vertex write ie and take into account that momentum is con-
served.
(8) For any internal fermion line with momentum p write the Dirac fermion propagator,
i
SF (p) = pm+i
8.1 e+ e 2: the electron propagator 89

These are called the Feynman rules for QED. But one important rule is still missing: which
is the term we should apply for an internal photon line: that is, the photon propagator. We
will deduce it in the following sections and it turns to be ig /k 2 , but before doing so, let
us compute the probability of the process we are studying from the amplitude we have just
obtained.

8.1.2 The e+ e 2 cross section


The study of phenomena where a beam (of energy or particles) hits a target made of a single
entity (such as a single atom) or several of those entities, is a common experimental method
to study interactions. Very often the eects of these phenomena (such as the scattering of the
beam) are proportional to the intensity of the beam. In this case we can define an eective area
of the target as:
energy(particles) scattered per unit time
target = (8.22)
energy(particles) incident per unit surface and unit time
Thus, the significance of this eective area is that all the (energy)particles from the incident
beam that impinges on this area will be scattered. Moreover, assuming that there is no collective
eect between the targeted entities (such as screening), this area of the target will be just the
eective area of a single entity times the number of entities targeted. This allows us to define
the eective area of a single entity of the target as:
energy(particles) scattered per second
=
energy(particles) incident per unit surface and unit time number of targeted entities
(8.23)
A typical example of this phenomena is a pencil beam of flux and section S, made of
particles with speed v, hitting a thin target made of particles with particle density and length
L. Assuming a probability prob of interaction between a particle of the beam and a particle of
the target, the number of scattered particles per unit time will be:

particles scattered per unit time = (S) (LS) prob (8.24)

thus, the cross section defined in equation 8.23 is:


(S) (LS) prob V
= = S prob = prob (8.25)
(LS) vT
If we are interested in only studying the scattering around a given region of the phase space

of the final products, the probability can be computed by f |Sf i |2 , where the sum runs over
these phase space regions. All the previous relations are clearly maintained and we should talk
about the dierential cross section:
V
d = |Sf i |2 (8.26)
vT f

In the process we are studying, we are interested in the dierential cross section to find the
final state in a particular range of momenta, k1 dk1 /2 and k2 dk2 /2 for the two photons. As:
V
= d3 k (8.27)
kdk/2
(2)3
90 Chapter 8. QED

the dierential cross section will be:


V (2)4 4 (p2 + p1 kb ka )
d = | M|2
vT f 0 0 0
(2p2 V )(2p1 V )(2ka0 V )(2kb V )
3 (2) (p2 + p1 kb ka )V T
V V V 4 4
= d3
k a d kb |M|2
vT (2)3 (2)3 (2p02 V )(2p01 V )(2ka0 V )(2kb0 V )
(2)4 4 (p2 + p1 kb ka ) 3 3
2 d ka d kb
= |M|
(2p02 )(2p01 )v(2ka0 )(2kb0 ) (2)3 (2)3
(2)4
= |M|2 d2 (8.28)
4p02 p01 v
where d2 is the two body Lorentz-invariant phase space introduced above:
d 3 ka d 3 kb
d2 4 (p2 + p1 kb ka ) (8.29)
(2)3 2ka0 (2)3 2kb0
which in the C.M. system gives:

pcm |
|
d2 = d (8.30)
all except (2)6 4 s

where pcm (pcm ) are the momenta of the final photons (initial leptons) in the C.M., s
the C.M. energy and d the dierential solid angle around the final particle of momentum
p , as shown in Figure 8.4. This, plus the fact that in the C.M. p02 p01 v = E1 E2 (v1 + v2 ) =

E1 | p2 | + E2 |
p1 | = s| pcm | allows us to write:

d pcm |
|M|2 |
= (8.31)
d C.M. 64 s |
2 pcm |

Figure 8.4: C.M. collision a + b c + d

They are many useful variants of the previous formula. If the cross section does not depend
on the azimuthal angle, the scattering cross section can be written in a Lorentz-invariant form:
d |M|2
= (8.32)
dt 16(s, m21 , m22 )
8.1 e+ e 2: the electron propagator 91

with t (p p)2 and (x, y, z) = x2 + y 2 + z 2 2xy 2yz 2zx.


Let us go back to the M calculation from equation 8.21, dropping i since it is well behaved:

p 1 kb + m
M = ie2 vp2 s2 ka a k up s + (a b) (8.33)
(p1 kb )2 m2 b b 1 1

as it is Lorentz invariant, we will calculate it in the lab frame where the electron momentum is
p1 = (m, 0). The polarization vectors are perpendicular to k and of the type = (0, ). Thus,
in the lab frame, p1 = 0 and taking into account that a b = 2a b b a, this implies that p1
and anti-commute. So:

p1 kb b up1 s1 = kb b p1 up1 s1 = kb b mup1 s1 (8.34)

since ( p m)u = 0. This together with (p1 kb )2 = m2 + 0 2p1 kb , allows us to write

kb
M = ie2 vp2 s2 ka a k up s + (a b)
2p1 kb b b 1 1
( )
e2 k b b k a k a a k a a k b k b b
= i vp2 s2 + up1 s1
2 p 1 ka p1 kb
e2
i vp2 s2 (A)up1 s1 (8.35)
2
We assume that the initial particles are not polarized. Thus, in the calculation of the square of
M we will average over the 4 possible spin configurations of the electron and positron:

1 e4 e4
|M|2 = up1 s1 Avp2 s2 vp2 s2 Aup1 s1 = up s A( p2 m)Aup1 s1
4 s1 ,s2 16 s1 s2 16 s1 1 1
e4 ( ) e4 ( )
= (up1 s1 ) A( p2 m)A (up1 s1 ) = ( p1 + m) A( p2 m)A
16 s1 16
e4 ( )
= T r ( p1 + m)A( p2 m)A
16
( ( )
e4 k a a k a k b b k b b k b k a a
= T r ( p1 + m) +
16 p 1 ka p 1 kb
( ))
k b b k a k a a ka a kb kb b
( p2 m) + (8.36)
p 1 ka p 1 kb

since a b c = c b a = c b a, ( = ).
From the expression of equation 8.36 we can see that there are 4 terms similar to this one:

e4 1
T1 = T r (( p1 + m) ka a ka kb b ( p2 m) kb b ka ka a ) (8.37)
16 (p1 ka )2

that we are going to consider in detail. The other 3 can be manipulated in the same way. To
proceed, we will use, among others, the following rules:

a a = a2

a b = 2a b b a

T r( a b) = 4a b
92 Chapter 8. QED

T r( a b c d) = 4(a b c d a c b d + a d b c)

The trace of an odd number of s is zero.

In particular ka ka = ka2 = 0, a a = 2a = 2a = 1. Then:

e4 1 ( )
T1 = T r p 1 k k a k p 2 k k a k m 2
T r k k a k k k a k
16 (p1 ka )2 a a b b b b a a a a b b b b a a

e4 1 ( )
= T r ka a p1 ka kb b p2 kb b ka ka a + m2 T r ka a ka ka ka a
16 (p1 ka )2

e4 1
= (+T r p1 ka kb b p2 kb b ka + 0)
16 (p1 ka )2
e4 1 e4 1
= T r ((2p 1 k a k a p 1 ) k p 2 k k a ) = T r ( kb b p2 kb b ka ) + 0
16 (p1 ka )2 b b b b
8 p 1 ka
e4 1
= (kb b p2 kb b ka kb b kb b p2 ka + kb b ka p2 kb b ) (8.38)
2 p 1 ka
Now, using the relations:

p1 + p2 = ka + kb (p1 kb )2 = (p2 ka )2 p1 kb = p2 ka
kb b (p1 + p2 ) = kb b (ka + kb ) kb b p2 = kb b ka (8.39)

allows us to write:
e4 1 e4 1 ( )
T1 = (2p2 kb b kb b ka + p2 ka ) = 2(ka kb b )2 + p1 kb (8.40)
2 p1 ka 2 p1 ka
Proceeding with the other 3 terms in equation 8.36 in the same manner, we finally obtain:
( )
1 e4 p 1 kb p 1 ka
|M|2 = + 4(ka a kb b )2 + 2 (8.41)
4 s1 ,s2 2 p 1 ka p 1 kb

which does not depend on the azimuthal angle. This allows us to use equation 8.32 to express
the cross section in the lab frame where p1 = (m, 0) and p1 ka = mwa . Finally,
( )
d 2 a wb wa e2 1
= + 4(ka a kb b )2 + 2 , = (8.42)
d lab 8mp2 (m + E2 p2 cos a ) wa wb 4 137
In the low energy limit, where both electron and positron are almost at rest (wa = wb = m, p2
m2 , so

d 2 ( )
= 1 ( )2
(8.43)
d lab 4m2 2
k
a a k
b b

which shows that the two photons are favorably produced with perpendicular polarization. This
occurs because QED conserves parity, and the final state should have negative parity since the
initial parity of e+ e at rest is negative.
In case we do not measure the polarization of the final photons, the dierential cross section
should be the sum of all 4 of the contributions:
2 ( )
d 2 2
= 1 ( )2
= (4 1 (k k ) 2
) = (8.44)
d lab 4m2 2
k
a a k a b
b b
4m2 2 2m2 2
a b
8.2 The gamma propagator 93

and
1 d r2
labT OT = d = e , re (8.45)
2 d lab 2 m
The factor 1/2 is introduced to avoid double counting, as we have computed the dierential
cross section that one photon goes to d no matter which one of the two.

8.2 The gamma propagator


One important ingredient for the systematic calculation of the invariant matrix element M
using Feynman diagrams is the rule associated with a photon internal line. As we have seen,
this corresponds to the photon propagator.
We have two methods to obtain the expression for the photon propagator. The first one is
to compute the corresponding expected value of ordered fields in the vacuum, and the second
one is to solve the corresponding Green function.
To be more general, we will obtain it for massive photons. Knowing that a massive spin-1
particle is essentially a set of a Klein-Gordon fields,
its Lagrangian density is:
1 1
L = F F m2 A A (8.46)
4 2
Notice the change of sign in the mass term with respect to the Klein-Gordon field. The justifi-
cation for this is, as we will see, that this Lagrangian gives the correct Klein-Gordon equation
for each of the A fields that we take to be Hermitian.
Applying the principle of minimum action, we get the so-called Proca equation:

F + m2 A = 0 (8.47)

and taking the derivative of it: 0 = F + m2 A = m2 A , which shows that the


Lorentz condition is automatically required. Using this condition on the Proca equation we get:

0 = ( A A ) + m2 A = (2 + m2 )A (8.48)

which is the Klein-Gordon equation as expected. Now the reason for choosing the above La-
grangian density becomes clear.
Again, as in the massless photon field, the conjugate field of the first component is zero,
meaning that A0 is not independent. In fact:
1
A0 = (8.49)
m2
With a massive spin-1 particle there are 3 degrees of freedom. In this case, it is equivalent
to removing the extra degree from the 4fields A by eliminating 0 A0 from the Lagrangian or
just requiring the transverse condition A = 0. Thus, the plane-wave solutions at rest are of
the type:
eipx
A (x) = p (8.50)
2wV
94 Chapter 8. QED

with p = (m, 0) and the time component of the polarization vectors has to be zero to satisfy the
transverse condition: (0 = A = ip A p p = 0). This allows us to use as polarization
vectors at rest the two orthonormal vectors used for the massless photon field, plus the unit
vector in the momentum direction. When p = (m, 0) is boosted to become p = (p0 , p), these 3
polarization vectors transform thus:
( ) ( )( )
0 0
= , with = p0 /m, = |
p|/m

= (8.51)

showing that the transverse polarization vectors stay the same, while the longitudinal one (0 =
p|/m, p0 p/m). Finally, we can write the 3 polarization
0, = 1, = 0) is transformed to: (|
vectors corresponding to p:

p,1 = (0, p,1 )


p,2 = (0, p,2 )
p0
p,3 = (|
p|/m, p) (8.52)
m
Thus, the momentum expansion is:
3 (
)

A (x) = ap, p, e(x) + h.c. (8.53)
=1
p

where e(x) = eipx / 2EV . To proceed to its quantization we have just to identify a, a as the
annihilation and creation operators with the usual commutator rules for bosons.
Similar to the case of the massless spin-1 particle, we can define the new polarization vectors:
1
+ = (1 + i2 )
2
0 = 3
1
= (1 i2 ) (8.54)
2
and new creation/annihilation operators. This allows us to work with helicity instead of linear
polarization.
Let us proceed now to compute the Feynman propagator using the vacuum expectation of
the ordered fields:

DF (x y) =< 0|A (x)A (y)|0 > (x0 y 0 )+ < 0|A (y)A (x)|0 > (y 0 x0 ) (8.55)

Using the expression for the A field given above it is straightforward to prove:
( )
d4 p eipz
3
DF (z) =i p, p, (8.56)
(2)4 p2 m2 + i =1
( )
3
The sum =1 p
, p
, is a Lorentz tensor that obeys the relation:


3
p p
= g (8.57)
=1
mm
8.3 The scattering cross section a + b 1 + 2 + ...n 95

This allows us to write the propagator as:


p p
d4 p g + p2 ip(xy)

DF (x y) = i e (8.58)
(2)4 p2 m2 + i
and its Fourier transform
p p
g + p2
DF (p) =i (8.59)
p2 m2 + i
This last expression is just the factor we must use to calculate the invariant matrix element
M for an internal line of a massive spin-1 particle.
To obtain the propagator of the photon, we just take m = 0 in equation 8.59. In addition,
taking into account that in general a photon is attached to a fermion current, the second term
in the numerator does not contribute and the photon propagator can be written:
g
DF (k) = i (8.60)
k2 + i
As we have seen for the Klein-Gordon and Dirac field, this propagator must be related with
the corresponding Greens equation of 8.47. Working in an arbitrary gauge, the Greens equation
is
1
(2g (1 ) + m2 )D (x y) = i (4) (x y), (8.61)

whose solution, in momentum space, is:

g (1 ) p pp2
(i) , (8.62)
p2 m2 + i
Again, taking the limit m 0, and fixing the gauge as = 1 (called the Feynman gauge) we
obtain the same equation as 8.60.

8.3 The scattering cross section a + b 1 + 2 + ...n


We now want to find a general expression for the cross section of a+b 1+2+...n. Independently
of the types of particles and interaction, following on from what we have done in the previous
section, it is clear that the matrix element can be written as:

(2)4 4 (pa + pb ( ni=1 pi ))
Sif = M (8.63)
(2p0b V )(2p0a V ) ni=1 (2p0i V ))

since each particle in the final state will contribute with a term proportional to eipi x / 2p0i V

and the initial particle with eiPa x / 2Pa0 V , and the d4 x will produce the 4 of the above
expression. The expression is valid to any order of the coupling constant. The matrix element
M depends on the type of particle and the interaction and can be computed as we have done
in the previous sections or systematically following the Feynmam rules.
Thus, the cross section we are looking for is:

(2)4 4 (pa + pb ( ni=1 pi )) n
d3 pi (2)4
d = n |M|2
= 0 p0 v |M| dn
2
(8.64)
(2p0b )(2p0a )v i=1 (2p0i )) i=1
(2) 3 4p b a

where dn is the Lorentz-invariant phase space of n particles introduced above.


96 Chapter 8. QED

8.4 QED Feynman rules


The invariant matrix M can be obtained systematically from diagrams by applying the following
rules:

Draw the diagrams connecting the initial state with the final state using the only possible
vertex ee up to the desired order in the coupling constant.

Then assign the following factors to construct M.

(1) For an outgoing fermion particle with momentum p write u(p).


(2) For an outgoing fermion antiparticle with momentum p write v(p).
(3) For an incoming fermion particle with momentum p write u(p).
(4) For an incoming fermion antiparticle with momentum p write v(p).
(5) For an outgoing photon with momentum p write (p). When considering the helicity
basis, use (p).
(6) For an incoming photon with momentum p write (p).
(7) For the interaction vertex write ie and take into account that momentum is conserved.
(8) For any internal fermion line with momentum p write the Dirac fermion propagator:

i
SF (p) = . (8.65)
p m + i

(9) For any internal photon line with momentum k and free indices write:
g
DF, (k) = i . (8.66)
k2 + i
In general more than one Feynman diagram contributes to a process and sometimes identical
particles appear in the initial and final states.
(10) The overall sign of a given diagram is not observable, but diagrams that dier in the
exchange of two identical particles in the initial or final state, or a particle-antiparticle in the
initial or final state, respectively, should come with opposite signs on account of Fermi statistics.
(11) To compute the cross section, divide by n! if there are n identical particles in the final
state.
At higher orders in perturbation theory loops appear.
(12) For any internal fermionic loop include a (1) sign and take the Dirac trace.
(13) Fermionic loops with an odd number of photon insertions need not be included (Furrys
theorem).
d4 p
(14) For any closed loop: (2) 4.
Chapter 9

Elementary processes in QED

In this chapter we will compute some elementary QED processes of the type a + b c + d. First,
we will compute the scattering of electrons by a central potential and study the kinematics of
that type of process.

9.1 e scattering by a central potential


Let us now discuss the scattering of an electron by a Coulomb potential, which corresponds to
the process e (p) e (p ). The change in momentum is due to the interaction with an external
fixed static potential given by:
Ze
= 0,
A A0 = , r = |x|. (9.1)
4r
From equation 7.29, the corresponding S-matrix element will be:
( )

Sf i < e (p )|i
d xLint |e (p) >=< 0|aps i
4
d xe : (x) (x) : A (x) aps |0 >
4

( [ ] )
= ie < 0|a s
p
4
d
xij i0 (x)j (x) + i (x)(x)j + i0 (x)j0 (x) j0 (x)i (x) A (x) aps |0 >
(9.2)

where we have used the notation introduced in 8.12. As the last expression is sandwiched
within the vacuum, and there is one creation and one annihilation electron operator outside
the brackets, this implies that only the first term inside the brackets will survive (creates and
annihilates an electron). Thus
( )
Sf i ie < 0|aps d x (x) (x)A (x) aps |0 >
4 0

( )
= ie < 0|aps d 4
xaps fp s (x) aps fps (x)A (x) aps |0 >

up s ups
= ie d4 xei(pp )x A (x) (9.3)
2E V 2EV
Defining k = p p , for this particular potential we have:

4 ikx 0 ik0 x0 3 ik
x 0 Ze
d xe A0 (x) = dx e d xe A0 (x) = 2(k ) d3 xeikx . (9.4)
4r
98 Chapter 9. Elementary processes in QED

Taking into account that:


1 4
d3 xeikx = , (9.5)
r |k|2
for the amplitude for this process we get:
up s 0 ups Ze 2(E E )
Sf i ie 2(E E ) = M (9.6)
2E V 2EV |k|2 (2EV )(2E V )
where, similarly to what happened in the previous chapter, we have defined the invariant matrix
element M, which in the present case is:
Ze ig
M = ieu(p s ) 0 u(ps) = u(p s )(ie )u(ps) jZe (9.7)

|k|2 |k|2
(iZe,
where jZe 0). Notice that this last expression is the one we would obtain by applying the
Feynman rules given in the previous chapter to Figure 9.1, where instead of using the current
in momentum space u(p s )(ie )u(ps) as is done for the electron, we use the current iZe for
the Coulomb potential source. This is just i-times the Fourier transform of the charge density
of a static point source Ze:


jZe
(x) = (Ze(x)), 0) jZe (q) =
d3 xjZe (x)eiqx = (Ze, 0) (9.8)

Figure 9.1: e scattering


As in the previous chapter, we are interested in studying the scattering around a given region
of the phase space of the outgoing electron. In this case, the dierential cross section is:
V V V V
3 2(E E )
d = prob = |Sf i |2 = d p | M|2
vT vT 3
p
d p
vT (2)3 (2EV )(2E V )
d3 p
V V 3 2(E E )T (2)
= d p |M|2
= |M|2
(E E ) (9.9)
vT (2)3 (2EV )(2E V ) 2Ev (2)3 2E
If we are not interested on the spin of the electron, we must add the two possible final
configurations and average over the two initial ones:
1 1 (Ze2 )2
|M |2 = |M |2 = T r( 0 ( p + m) 0 ( p + m)), (9.10)
i f
2 i f 2 |k|4
9.2 Kinematics of the process a + b 1 + 2 99

and after well-known manipulations:


(Ze2 )2
|M |2 = (4EE 2pp + 2m2 ). (9.11)
i f |k|4

Since the external potential is time independent, the process conserves energy (notice the (E
E ) on equation 9.6 ), E = E, thus | p | and pp = E 2 |
p| = | p|2 cos , where is the deflection
angle. The modulus of the change in 3-momentum k = p p will be:


p|2 sin2 .
k 2 = 4| (9.12)
2
All we have to do is to compute the phase space integral. Elementary considerations lead to:

1 1 1 E
d = p |2 d|
d(cos )| p | (|
p | |
p|) |M |2 . (9.13)
p| 2
2| 2E |
p| i f

and, taking into account that in the system of units we use: e2 = 4

d Z 2 2
= 2 4 (2|p|2 (1 + cos ) + 4m2 ). (9.14)
d(cos ) k

This is the Mott cross section. In the non-relativistic limit it reduces to the well-known Ruther-
ford cross-section.

9.2 Kinematics of the process a + b 1 + 2



In the process a+b 1+2+3+...n the Lorentz-invariant element i f |M |2 , where the sums
run over the possible spin states of the particles, have (n + 2)3 4 6 degrees of freedom (d.o.f).
As that element can only depend on the 3-momentum of the particles, initially one could think
that we have (n + 2)3 d.o.f, but from this number we have to subtract 4 restrictions coming from
the 4-momentum conservation and 6 additional restrictions coming from the Lorentz invariance
(3 from boost and 3 from rotations, or what is equivalent: we can fix the frame where particle b
is at restthis removes 3 d.o.fwe can chose one of the axes in the direction of particle 1this
removes 2 d.o.fand finally we can chose the second axis perpendicular to the previous one in
such a way that particle 2 is contained in the plane defined by this new axis and the previous
onethis removes 1 d.o.f).

In the case a + b 1 + 2, its Lorentz-invariant element i f |M |2 will have only 2 d.o.f
and can be expressed as a function of only two independent Lorentz-invariant quantitiesscalar
products of the 4 momenta. For this reason we introduce the Mandelstam variables:

s = (pa + pb )2 = (p1 + p2 )2 = 4ECM


2
, t = (pa p1 )2 = (pb p2 )2 ,
u = (pa p2 )2 = (pb p1 )2 .
(9.15)

These quantities are not independent since s + t + u = ma + mb + m1 + m2 , thus i f |M |2
2 2 2 2

can be always written as a combination of two of them. As particles (antiparticles) going in/out
with momentum p and spin s can be interpreted as antiparticles(particles) going out/in with
100 Chapter 9. Elementary processes in QED

momentum p and spin s, this will allow to related the process a + b 1 + 2 with others like
a + 1 b + 2, just finding out the relation between their Mandelstam variables.
In the CM if m = ma = mb = m1 = m2 , we have:

(a,1)
t = (pa p1 )2 = 2m2 2pa p1 = 2m2 2(Ea,CM
2
|
pa,CM |2 cos CM ) = (9.16)
(a,1)
= 2|
pa,CM |2 (1 cos CM )) (9.17)
(a,2) (a,1)
u = (pa p2 )2 = 2|
pa,CM |2 (1 cos CM )) = 2|
pa,CM |2 (1 + cos CM )) (9.18)

We state that the cross section for a process a+b something can be written as (eq. 8.64):

(2)4
d = |M|2 dn (9.19)
4p0b p0a v

It is easy to see that in the C.M. system and in the lab frame: p0b p0a (va +vb ) = (pa pb )2 m2a m2b .
Defining (x, y, z) x2 + y 2 + z 2 2xy 2yz 2zx we can write:

(2)4 (2)4
d = |M|2 dn = 1/2 2 , m2 )
|M|2 dn (9.20)
4 (pa pb ) ma mb
2 2 2 2 (s, m a b

If we want to study a + b 1 + 2 to obtain the dependence of the cross section as, for
example, a function of the Mandelstam variable t, we must perform:


d (2)4
(s, t) = 1/2 2 2
(t (pa p1 )2 )d2 |M|2 (9.21)
dt 2 (s, ma , mb ) i f all

This expression is general when going to n particles instead of two, but in that case the expression

will depend on more than the two invariants t and s (or u). As i f M is only a function of
two variables, s and t, we can perform the integral over the phase space at fixed s and t:

1 1
I (t (pa p1 )2 )d2 = ... = = (9.22)
all 4(2) |
6 pa |(E1 + E2 ) 2(2) (s, m2a , m2b )
6 1/2

which allows us to write:


d 1
(s, t) = 2 2 |M|2 (9.23)
dt 16(s, ma , mb ) i f

As variable t is related to the scattering angle between particles a and 1 in the C.M.
frame, this allows us to obtain the cross section in the C.M. frame as a function of that scattering
angle in a simple manner:

d 1 d dt 1 1 1/2 (s, m21 , m22 )


|CM = = |M|2 (9.24)
d 2 dt d cos 64 2 s 1/2 (s, m2a , m2b ) i f

Notice that if we do not sum over the spin of the particles, M will have additional d.o.f. (for
example 2 for each spin-1/2 particle) and some of the simplified expressions given above do not
apply.
9.3 e (p1 )e+ (p2 ) (q1 )+ (q2 ) 101

Figure 9.2: Lowest-order Feynman diagrams for e (p1 )e+ (p2 ) (q1 )+ (q2 )

9.3 e (p1 )e+ (p2 ) (q1 )+ (q2 )


We will first discuss the process e (p1 )e+ (p2 ) (q1 )+ (q2 ). Direct application of the Feyn-
man rules over the lowest-order diagram in Figure 9.2 gives:

g
M = ie2 v(p2 ) u(p1 ) u (q1 ) v (q2 ) (9.25)
(p1 + p2 )2 + i

where is used for muons. If we are not interested in the spin states:

1 e4 1
|M |2 = |M |2 = |v(p2 ) u(p1 ) u (q1 ) v (q2 )|2
i f
4 spins 4 spins (p1 + p2 )2
e4
= (v(p2 ) u(p1 ))(v(p2 ) u(p1 )) (u (q1 ) v (q2 ))(u (q1 ) v (q2 ))
4(p1 + p2 )4 spins
e4
= T r(( p2 me ) ( p1 + me ) )T r(( q1 + m ) ( q2 m ) )
4(p1 + p2 )4
e4 ( )
= 32 2m 2 2
m
e + m 2
e (q 1 q 2 ) + m 2
(p1 p2 ) + (p 1 q 1 )(p 2 q 2 ) + (p 1 q 2 )(p 2 q 1 )
4(p1 + p2 )4
(9.26)

To keep the discussion simple, we will neglect masses (this is usually justified for electrons, not
so much so for muons, except at very high energies). In this case:

s = (p1 + p2 )2 = 2p1 p2 = 2q1 q2 (9.27)


s
t = (p1 q1 )2 = 2p1 q1 = 2p2 q2 = (1 cos ) = s sin2 (/2) (9.28)
2
s
u = (p1 q2 ) = 2p1 q2 = 2p2 q1 = (1 + cos ) = s cos2 (/2)
2
(9.29)
2
and
e4 2 2
4t + u
|M |2 = 8 ((p 1 q 1 )(p 2 q 2 ) + (p 1 q 2 )(p2 q 1 )) = 2e (9.30)
i f
(p1 + p2 )4 s2
102 Chapter 9. Elementary processes in QED

Finally the cross section in C.M. is is:

d 1 1 1/2 (s, m21 , m22 ) 1 1 4 t2 + u 2 2 t2 + u 2 2


|CM = |M|2
= 2e = = (1+cos2 )
d 64 2 s 1/2 (s, m2a , m2b ) i f 64 2 s s2 2s s2 4s
(9.31)
being the angle between the incoming e
and the outgoing
in the C.M. frame. This result
has been confirmed experimentally at high energies, except for additional contributions coming
from week interactions. The total cross section is:

d 42
= d = (9.32)
d 3s

9.3.1 Polarizations in e (p1 )e+ (p2 ) (q1 )+ (q2 )


If we are interested in the spin states, we should not sum over them, thus:
1
|M |2 = e4 |v(p2 ) u(p1 ) u (q1 ) v (q2 )|2
(p1 + p2 )2
e4
= (v(p2 ) u(p1 ))(v(p2 ) u(p1 )) (u (q1 ) v (q2 ))(u (q1 ) v (q2 ))
(p1 + p2 )4
e4 (1 + 5 s2 ) (1 + 5 s1 )
= T r( ( p 2 m )
e ( p1 + me ) )
(p1 + p2 )4 2 2
(1 + 5 s1 ) (1 + 5 s2 )
T r( ( q1 + m ) ( q2 m ) ) (9.33)
2 2
where we have used the relations given in 2.92. Let us recall that the polarization 4-vector
appearing in the last expression corresponds to the boost of the polarization unit vectors at rest
to the frame of the corresponding particle. As we saw in chapter 2, this can be greatly simplified
in the high-energy limit when considering helicity projections. In that case, the two helicity
projection operators that filter out the two eigenstates of the the component of physical spin
along p are:
1 5
(p) = (9.34)
2
where = 1 for energy solutions.
Using chirality operators:
1
PR (1 5 ) (9.35)
L 2
we see that PR filters out right-handed helicity (+) when applied to a fermion, and left-handed
helicity (-) when applied to an anti-fermion. PL filters out left-handed fermions and right-handed
anti-fermions.
These operators have the following properties:
1 1
PR = (1 5 ) = (1 5 ) = P L
L 2 2 R
1 1
PR = (1 5 ) = (1 5 ) = P L
L 2 2 R
PR2 = PR
L L
PR PL = 0 (9.36)
9.4 e (p1 )+ (p2 ) e (q1 )+ (q2 ) 103

We can now consider the cross section when a left-handed electron collides with a left-handed
positron. In this case, the matrix element can be written as:
1
|M |2 = e4 |PR v(p2 ) PL u(p1 ) u (q1 ) v (q2 )|2 (9.37)
(p1 + p2 )2

since PL u(p1 ) (PR v(p2 )) is a left-handed electron (positron). Then

PR v(p2 ) PL u(p1 ) = v PR PL u = vPL PL u = v PR PL u = 0 (9.38)

Thus e+ (lef t) + e (lef t) + at least to first order of perturbation theory. Similarly for
e+ (right) + e (right) + . Therefore, at high energies this process can only occur via
e+ (lef t) + e (right) or e+ (right) + e (lef t) where the total angular momentum is 1 (states
|1, 1 >).

9.4 e (p1 )+ (p2 ) e (q1 )+ (q2 )

Figure 9.3: Lowest-order Feynman diagram for e (p1 )+ (p2 ) e (q1 )+ (q2 ) scattering

From Figure 9.3 the matrix element is:


( ) 1
M = ie2 vq 2 vp 2 (uq1 up1 ) (9.39)
(p2 q2 )2

Taking the masses as equal to zero and summing over the spins we obtain:
1 s2 + u2
|M |2 = |M |2 = 2e4 (9.40)
i f
4 spins t2

as expected by comparison with the the related process e (p1 )e+ (p2 ) (q1 )+ (q2 ), inter-
changing particles and antiparticles in/out and

p2 q1

p2
q1
104 Chapter 9. Elementary processes in QED

Thus
t = (p1 q1 )2 (p1 + p2 )2 = s
s = (p1 + p2 )2 (p1 q1 )2 = t
u = (p1 q2 )2 (p1 q2 )2 = u
and we obtain 9.40 from 9.30 just interchanging t s.
Finally, the cross section is:
( ) ( )
d 2 s2 + u2 2 1 + cos4 (/2)
|CM = = (9.41)
d cos s t2 s sin4 (/2)
where is the angle between the incoming e and the outgoing e in the C.M. frame.

9.5 Bhabha scattering e (p1 )e+ (p2 ) e (q1 )e+ (q2 )


In the process e (p1 )e+ (p2 ) e (q1 )e+ (q2 ), called Bhabha scattering, we have two Feynmam
diagrams (Figure 9.4) contributing to the lowest order in perturbation theory. Comparing with
the process e+ e + (Figure 9.2) the new diagram is called the T channel and it is
dominant at a low scattering angle.

Figure 9.4: Lowest-order Feynman diagrams for Bhabha scattering

The matrix element is


1 1
M = ie2 (vp2 up1 ) (uq1 vq2 ) ie2 (vp2 vq2 ) (uq1 up1 ) M1 + M2
(p2 + p1 )2 (p2 q2 )2
(9.42)
Notice the relative sign dierence between the two diagrams. As the two diagrams dier in the
exchange of a particle-antiparticle in the initial or final state, the opposite sign is necessary to
account for Fermi statistics.
To simplify the calculations, we will set me = 0 and we will sum over the spins. In this case:
1 1 1 1 1
|M |2 = |M |2 = |M1 +M2 |2 = |M1 |2 |M2 |2 + (M M2 +M1 M2 )
i f
4 spins 4 spins 4 spins 4 spins 4 spins 1
(9.43)
9.6 Moller scattering e (p1 )e (p2 ) e (q1 )e (q2 ) 105

The first term, corresponding to the annihilation diagram, gives the same as in the e e+ +
case (9.30):
1 t2 + u 2
|M1 |2 = 2e4 1 + cos2 (9.44)
4 spins s2

The second term, corresponding to the T channel, gives the same as in the e + e +
case (9.40):
1 s2 + u2 1 + cos4 (/2)
|M2 |2 = 2e4 (9.45)
4 spins t2 sin4 (/2)

and the interference term gives:

1 2u2 cos4 (/2)


(M1 M2 + M1 M2 ) = 2e4 (9.46)
4 spins st sin2 (/2)

where is the angle between the incoming e and the outgoing e in the C.M.frame.
Finally:
( )
d 2 1 1/2 (s, m21 , m22 ) 2 1 4 t2 + u2 s2 + u2 2u2
|CM = 2 2
|M|2 = 2e + +
d cos 1/2 2
64 s (s, ma , mb ) i f 64 2 s s2 t2 st
( )
2 1 1 t s
= u2 ( + )2 + ( )2 + ( )2 (9.47)
s s t s t

which at very low angle goes as 1/4 .

9.6 Moller scattering e (p1 )e (p2 ) e (q1 )e (q2 )

In the process e (p1 )e (p2 ) e (q1 )e (q2 ), called Moller scattering, we have two Feynmam
diagrams (Figure 9.5) contributing to the lowest order in perturbation theory. Comparing with
the process e e (Figure 9.3) the new additional diagram is called the U channel.

Figure 9.5: Lowest-order Feynman diagrams for Moller scattering


106 Chapter 9. Elementary processes in QED

The matrix element is:


1 1
M = ie2 (uq2 up2 ) (uq1 up1 ) ie2 (uq1 up2 ) (uq2 up1 ) M1 + M2
(p2 q2 ) 2 (p2 q1 )2
(9.48)
Notice the relative sign dierence between the two diagrams. As the two diagrams dier in the
exchange of two identical particles in the initial or final state, the opposite sign is necessary to
account for Fermi statistics.
To simplify the calculations, we will set me = 0 and we will sum over the spins. In this case:
1 1 1 1 1
|M |2 = |M |2 = |M1 +M2 |2 = |M1 |2 |M2 |2 + (M M2 +M1 M2 )
i f
4 spins 4 spins 4 spins 4 spins 4 spins 1
(9.49)
The first term, corresponding to the T channel, gives the same as in the e + e + case
(9.40):
1 s2 + u2
|M1 |2 = 2e4 (9.50)
4 spins t2

The second term, corresponding to the so called U channel, is very easy to obtain since it is
the same as the T channel but interchanging q1 q2 , which is equivalent to interchange t u.
Thus:
1 s 2 + t2
|M2 |2 = 2e4 (9.51)
4 spins u2

and the interference term gives:

1 2s2
(M1 M2 + M1 M2 ) = 2e4 (9.52)
4 spins ut

Finally:
( )
d 2 1 1/2 (s, m21 , m22 ) 2 s2 + u2 s2 + t2 2s2
|CM = |M|2
= + + (9.53)
d cos 64 2 s 1/2 (s, m2a , m2b ) i f s t2 u2 ut
Appendix A

Lorentz invariance

Introducing the metric tensor g by:



1 0 0 0

0 1 0 0
g
= (A.1)
0 0 1 0
0 0 0 1

a Lorentz transformation must satisfy the relation:

g = g (A.2)

which ensures that the dot product of 4-vectors is invariant:

A B = A B = g A B = g A B = g A B = A B (A.3)

In matrices, the previous relation A.2 can be written as LT GL = G where: L ( ) and


G (g ). This equation is the sucient and necessary condition for a dot-product of any
4-vector to be invariant under the transformations. It is easy to show that the set of L that
satisfy LT GL = G form a group, called the Lorentz group.
The relation A.2 can be rewritten as: = g = Comparing it with the definition
of the inverse matrix of , (1 ) = , we see that:

(1 ) = (A.4)

This means that starting from the original , the inverse can be obtained by first up-down
interchanging the two indices ( ) and then transposing it .
A 4-vector x transforms as x = x and the inverse transformation is

x = (1 ) x = x (A.5)

Examples of physical 4-vectors that transform under the Lorentz group are:

X = (X 0 , X 1 , X 2 , X 3 ) = (ct, x, y, z) = (ct, x)
P = (P 0 , P 1 , P 2 , P 3 ) = (E/c, px , py , pz ) = (E/c, p) (A.6)
108 Apendix A. Lorentz invariance

If the frame K is moving with velocity in the frame K in the x-direction, the above vectors
transform by the Lorentz boost transformation:

E 0 0 E

px 0 0 p
x
p = 0 0
(A.7)
y 0 1 py
pz 0 0 0 1 pz

where = 1/ 1 2 and = . Notice that 2 2 = 1. As P P = P 0 P 0 P 1 P 1 P 2 P 2
P 3 P 3 = E 2 p2 does not change its value under a Lorentz transformation, since: P P = P P
Two consecutive Lorentz boosts do not commute in general and are not another Lorentz
boost. In order to form a group, a rotation and a boost have to be combined. Actually, the
Lorentz group defined by LT GL = G is even larger than rotation + boost. It contains also T
(time reversal) and P (parity transformations)

1 0 0 0 1 0 0 0

0 1 0 0 0 1 0 0
T =
0
, P = (A.8)
0 1 0
0
0 1 0
0 0 0 1 0 0 0 1
Suppose f (x) is a scalar field, then how does f /x = (f /t, f /x, f /y, f /z) trans-
form?. To find out, take the relation f (x1 ) f (x2 ) = f (x1 ) f (x2 ) when x1 and x2 are very
close:
f f

dx = dx (A.9)
x x
f is Lorentz invariant, and since dx is clearly a superscript 4-vector
this tells us that x dx
f
(contra-variant), x should transform as a subscript 4-vector (covariant). To make this point

clear, the operator x is often written as




= ( , ) (A.10)
x t
which transforms with the inverse transformation:
x
= = = = (A.11)
x x x x
Then an expression such as j could be a Lorentz-invariant quantity, provided that j =
(, j) transforms as a 4-vector.
The Lorentz transformations that belong to the rotation+boost part (called the proper
Lorentz transformations) can be written as:

= ei Ki +i Li , i = 1, 2, 3 (A.12)

where i and i are real numbers and Ki , Li are 4x4 matrices called the generators of the group.
Let us review the rotations and the boost. A small rotation around the z-axis by a infinites-
imal angle will transform (x, y) to:
( ) ( ) ( )( ) ( ( )) ( ) ( )
x x y 1 x 0 1 x x
= = = 1 + = (1 + Lz )
y y + x 1 y 1 0 y y
(A.13)
109

where Lz is the generator of the rotations around the z-axes. Any finite rotation through angle
can be obtained by n consecutive rotations of angle /n
( ) ( ) ( ) ( )( )
x n x Lz x cos sin x
= limn (1 + /nLz ) =e = (A.14)
y y y sin cos y

Similarly, we can obtain Lx and Ly and using 4 dimensional rotations, we obtain:



0 0 0 0 0 0 0 0 0 0 0 0

0 0 0 0 0 0 0 1 1 0
0 0
L1 Lx =
0 0 0 1 , L2 Ly = 0
, L3 Lz =


0 1
0 0 0 0 0

0 0 1 0 0 1 0 0 0 0 0 0
(A.15)
Then a general rotation can be written as = e L . This is not a rotation 3 around
z followed
i i

by a rotation 2 around y and a rotation 1 around x. This is a single rotation = 12 + 22 + 32


Prove it for an infinitesimal rotation.
around vector .
For a boost along the x-axis we have seen that it is represented by (only writing the t and
x coordinates): ( )

= (A.16)

when is small = 1 and
= ; thus
( ) ( )
1 0 1
= =1+ = 1 + Kx (A.17)
1 1 0

Suppose we apply n such boosts consecutively, the resulting transformation, when n is taken to
infinity while = n is fixed:
( )
cosh sinh
= limn (1 + /nKx )n = eKx = .... = (A.18)
sinh cosh

and comparing with the previous results = cosh, = sinh or = / = tanh. Notice
that the n-consecutive boosts by velocity did not result in a boost by velocity = n rather
in a boost by = tanh(n). This is a consequence of the addition of velocities rule. Similarly,
we can obtain the boost in y or z and finally we obtain the generators:

0 1 0 0 0 0 1 0 0 0 0 1

1 0 0 0 0 0 0 0 0 0 0 0
K1 Kx =
0
, K2 Ky =
1
, K3 Kz =
0

0 0 0 0 0 0 0 0 0
0 0 0 0 0 0 0 0 1 0 0 0
(A.19)

Now ei Ki represents a boost in the -direction by a velocity = tanh
Given the explicit forms of Ki , Li we can verify the following commutation relations:

[Li , Lj ] = ijk Lk , [Ki , Kj ] = ijk Lk , [Li , Kj ] = ijk Kk , (A.20)

this shows that rotations can never generate a boost, but a boost followed by another boost will
in general contain some rotation.
110 Apendix A. Lorentz invariance

A general proper Lorentz transformation can then be written as:

= ei Ki +i Li (A.21)

This, however, due to the commutation relations, does not have an interpretation as simple
as ei Ki or ei Li . Every proper orthochronous Lorentz transformation can be decomposed as:
= (R1 )(K1 )(R2 ), where (Ri ) are two spatial rotations and (K1 ) a boost in the x-
direction. It is clear that P and T cannot be generated by this transformation an they do not
belong to this subgroup.
Notice that the structure of the group is uniquely defined when the commutators among
all the generators are specified. In general, commutators of generators Gi , i = 1, ...n can be
expressed as linear combinations of Gi s: [Gi , Gj ] = cijk Gk ,, where cijk are constants called the
structure of the group. It is clear that if we have two sets of matrices in dierent dimensions, but
both sets satisfy the same commutation rules, their exponentiations are dierent representations
of the same group.
Appendix B

The harmonic oscillator

Consider the Hamiltonian:


P2 1
H= + m 2 X 2 . (B.1)
2m 2
Heisenberg gave an algebraic solution to this problem without giving a concrete representa-
tion. With the fundamental constant h and the two physical parameters m and present in
the Hamiltonian, we can define a quantity b with dimensions of length, b m h
. We define
the dimensionless operators:
1 b
X X , P P , (B.2)
b h
amb [X, P ] = i I. In terms of these operators, H = 12 h(X 2 + P 2 ).
We now introduce the relations,

1 1
a (X + iP ) , a (X iP ) , (B.3)
2 2

that perform [a, a ] = I, and the Hamiltonian can be expressed as:

1 1
H = h(a a + ) = h(N + ), N a a , (B.4)
2 2

which satisfies:

[H, a] = h[a a, a] = h[a , a] a = h a, [H, a ] = h a . (B.5)

Taking the base of the H eigenvalues (we will soon see that the energy eigenvalues can be labeled
with an index n, n = 0, 1, 2, ...; up to now n being any label)

H|En = En |En
Ha|En = (aH h a)|En = (En h)a|En
Ha |En = (a H + h a )|En = (En + h)a |En . (B.6)

Therefore a|En is an eigenvector with eigenvalue En h while a |En is an eigenvector with


eigenvalue En +h. For this reason we say that a is the creation operator and a the annihilation
112 Apendix B. The harmonic oscillator

operator. Since the expectation values of H cannot be negative, since |a a| = a|2 , we


take E0 as the minimum eigenvalues of H. Then:

a|E0 = 0 ,
1 1 h
0 = h a a|E0 = (H h)|E0 = (E0 h)|E0 E0 = . (B.7)
2 2 2
Now we can construct all the eigenvalues of H repeatedly using a a :
h
|E0 E0 =
2
h
|E1 a |E0 E1 = + h
2
h
|E2 a2 |E0 E2 = + 2 h
2
................
h 1
|En an |E0 En = + nh = (n + )h, n = 0, 1, 2, 3, . . . (B.8)
2 2
The spectrum has been completely determined, En = h 2 + nh, n = 0, 1, 2, 3, . An
energy such as E0,5 or any other positive non-integer numberis impossible because the state
a|E0,5 (which would have a non-zero norm, a|E0,5 2 = E0,5 |a a|E0,5 = E0,5 | h
H
12 |E0,5 =
2 E0,5 |E0,5 = 0) would have a lower energy than the ground state. In addition, again applying
1

a, we would have negative energies. At the end of this section we will see that the spectrum is
not degenerate.
The corresponding states should be normalized:

|En+1 = cn+1 a |En


H 1
1 = |cn+1 |2 En |aa |En = |cn+1 |2 En | + |En = |cn+1 |2 (n + 1) , (B.9)
h 2
and using the free phase, we can take the coecients cn to be real and positive
1 1
cn+1 = , |En = an+ |E0 . (B.10)
n+1 n!
It is easy to see now that:

a|En = n|En1 , a |En = n + 1|En+1 , N |En = n|En . (B.11)
Appendix C

Equivalent descriptions

In QM, the unitary operatorsgenerically denoted by O, to be distinguished from time evolution


operatorsdefine description changes that are physically equivalent. Its analogue in classical
mechanics (CM) and in the canonical formalism would be the canonical transformations: changes
of coordinates in the phase space that allow an equivalent reformulation of the dynamics of the
system under consideration.

QM allows us to calculate the probability of measuring a certain observable A(t) = i i (t) i (t)
(decomposition of the selfadjoint operator A in projectors i associated with its eigenvalues i )
over a certain state |(t), obtaining as a result the eigenvalue i (t), as:

P|(t) (A(t) :i (t)) = i (t)|(t)2 .

Notice that we allow the observable A to depend on time, so that both the eigenvalues, i , and
the projectors, i , can depend on time.
These results do not change if instead of these states and operators we use the transformed
ones under a unitary operator O(t)that can also depend on time t. The transformed states
|(t) and operators A(t) will be denoted as | (t), A (t). Clearly for states we have:

| (t) = O(t)|(t) (C.1)

while for the operators we request that the transformation is so, in order to complete the following
diagram at the same time

O(t)-
|1 (t) |1 (t)

A(t) A (t)

? ?
O(t)-
|2 (t) |2 (t)

from which it is clear that (recall that O (t) = O1 (t))

A (t) = O(t)A(t)O (t) (C.2)


114 Apendix C. Equivalent descriptions

To check that the results do not change, notice that if |(t) was an eigenvector A(t)|(t) =
a(t)|(t), then | (t) would also be a eigenvector of A with the same eigenvalue

A (t)| (t) = O(t)A(t)O (t)O(t)|(t) = a(t)O(t)|(t) = a(t)| (t)

and also with equal probability

P|(t) (A(t) :i (t)) = i (t)|(t)2 = (t)|i (t)|(t) = (t)|O (t)O(t)i (t)O (t)O(t)|(t)
= (t)|i (t)| (t) = P| (t) (A (t) :i (t)) , (C.3)

where we have used the fact that the transformed operator will be: A (t) = i i (t) i (t),
with i (t) = O(t)i (t)O (t). As the probabilities are the same, the calculation of expectation
values and dispersions will give identical results independent of whether we use the original
states/operators or the transformed ones.
Therefore, both the eigenvaluesthe possible outcomes of the measurementsand the pre-
dicted probabilities of observing themas a result of a measurementare invariant under any
unitary transformation defined by (C.2). In addition, as we will discuss in the next section, we
know the equations that determine the evolution of both the transformed states and the trans-
formed operators, so we can say conclusively that unitary operators provide us with description
changes that are physically equivalent: we can study the evolution of the system and make pre-
dictions using the initial states and operators and the usual dynamics or the transformed states
and operators and the new dynamics that we will deduce in the next section.

Time evolution of the transformed states and operators


Since O(t) depends on the time, even in the case of operators that are independent of time,
their transformed description will depend on it. In the transformed description, both states and
operators generally have a temporal evolution. Now we are going to determine it.
We will see that the transformed Hamiltonian,

H (t) = O(t)H(t)O (t)

is split:
H (t) = Hest

(t) + Hop (t)
to provide a dynamic time evolution for states, Hest (t) and another for operators H (t).
op

First we will determine Hest (t). The criterion to determine it is obviously that the states

| (t) also satisfy an equation of Schrodinger type ih t
(t)| (t):
| (t) = Hest

O(t)
ih | (t) = ih (O(t)|(t)) = ihO(t) |(t) + ih |(t)
t (
t t )
t
O(t)
= O(t)H(t)O (t) + ih O (t) | (t) Hest

| (t) . (C.4)
t
We have then found that:

O(t)
Hest = ih O (t) + O(t)H(t)O (t) (C.5)
t
115

Regarding the evolution of the transformed operators, from (C.2):


O(t) A(t) A (t) O(t)
ih A(t) + ihO(t) = ih O(t) + ihA (t) ,
t t t t
and therefore:
A (t) O(t) A(t) O(t)
ih = ih A(t)O (t) + ihO(t) O (t) ihA (t) O (t)
t t t t
O(t) A(t) O(t)
= ih O (t)A (t) + ihO(t) O (t) ihA (t) O (t)
t t t
O(t) A(t)
= [A (t), ih O (t)] + ihO(t) O (t)
t t
A(t)
= [A (t), Hop

(t)] + ihO(t) O (t) , (C.6)
t
where we have defined:
O(t)
Hop (t) ih O (t) (C.7)
t
as the operator that generates the time evolution of the operators due to the change of descrip-
tion. Notice that the ihO(t) A(t)
t O (t) in (C.6) cannot be absorbed into the definition of Hop
because we started from an observable A(t) with arbitrary dependence on t.

Thus it is clear that we have a complete description of the same physical system with new
states and new operators. Not only can we calculate probabilities of results with the transformed
states/operators but, as we see in (C.5) and (C.7), we can also follow the time evolution of the
system without the need to know the time evolution of the original state (before the change of
description) which was the only one we knew until now.

Temporal evolution pictures of Schrodinger, Heisenberg and Dirac


We now show several possibilities for the choice of the unitary operator O(t) .

Schrodinger picture (O(t) = I)


This is the description that we have adopted since the beginning for QM, where the time evo-
lution corresponds only to the states and not to the operators1 , it is the so-called Schrodinger
picture. If necessary, we will label states and operators as |(t)S and AS to show which picture
we are working with.

Heisenberg Picture (O(t) = U (t, t0 ))


In the Heisenberg picture the states do not undergo time evolution: this is achieved by changing
the description to
O(t) = U (t, t0 ) .
1
Although some operators may explicitly depend on time; for example in the case of a time varying magnetic
field, there will be a dependence on t, but not as an evolution given by the fifth postulate.
116 Apendix C. Equivalent descriptions

Notice that O(t0 ) = I, so the two pictures of Schrodinger and Heisenberg match at t0 . States
and operators in the Heisenberg picture are labeled as |H and AH (t). Notice that |(t)H =
U (t, t0 )|(t)S = |(t0 )S , which shows that |(t)H is independent of t.
From (D.3) it can also be deduced that:


ih U (t, t0 ) = U (t, t0 )H(t) ,
t
and
O(t) U (t, t0 )
ih O (t) = ih U (t, t0 ) = U (t, t0 )H(t)U (t, t0 ) ,
t t
so that, recalling (C.5) i (C.7), we see again that while the states do not evolve in time, Hest = 0,

the operators evolve according to HH Hop = H (t) H = H (t) = U (t, t )H(t)U (t, t ),
est 0 0

AH (t) A(t)
ih = [AH (t), HH (t)] + ihU (t, t0 ) U (t, t0 ),
t t
(Heisenberg equation).
Notice in particular that for the position and moment operators we have:

dXH 1
= [XH , HH ]
dt ih
dPH 1
= [PH , HH ] (C.8)
dt ih
equations analogous to the classical ones under the correspondence of the Poisson parentheses
to commutators, {, } ih1
[, ].
Also note that if neither H nor A depend on time, HH (t) = H and the Heisenberg equation
is
AH (t)
ih = [AH (t), H] .
t

Dirac picture (O(t) = U0 (t, t0 ))

In the Dirac picture, also called interaction picture, we consider a Hamiltonian of the type
H = H0 + H1 , where H0 dictates the free evolution of the system and H1 the interaction (for
P2 P2
example, in the case of two particles interacting with a potential we have: H0 = 2m11 + 2m22
and H1 = V (|x2 x2 |)). The states in the Dirac picture |I (I for interaction) only evolve
temporarily because of the presence of H1 . If there was no H1 , it would be the same case as the
Heisenberg picture, so in this case we take:

O(t) = U0 (t, t0 )

where U0 (t, t0 ) is the unitary time evolution operator generated by H0 . Proceeding in the same
manner as above, it is now

O(t)
ih O (t) = U0 (t, t0 )H0 (t)U0 (t, t0 ) ,
t
117

, (C.5) is now
obtaining that Hest

O(t)
HI (t) Hest (t) = ih O (t) + O(t)H(t)O (t)
t
= U0 (t, t0 )H0 (t)U0 (t, t0 ) + U0 (t, t0 )H(t)U0 (t, t0 )
= U0 (t, t0 )H1 (t)U0 (t, t0 ) (C.9)

and from (C.5),



Hop (t) = U0 (t, t0 )H0 (t)U0 (t, t0 ) .

So the evolution equations in the interaction picture are:



ih |(t)I = HI (t)|(t)I ,
t
for the states and, if we define as AI (t) the operators in the interaction picture (notice that
we make an exception in the notation in the definition of HI ) AI (t) = U0 (t, t0 )AS (t)U0 (t, t0 )
following (C.2),

AI (t) AS (t)
ih
= [AI (t), Hop (t)] + ihU0 (t, t0 ) U0 (t, t0 ) .
t t
(t) = H , and the last equation is:
If H0 does not depend on time, we can simplify: Hop 0

AI (t) AS (t)
ih = [AI (t), H0 ] + ihU0 (t, t0 ) U0 (t, t0 ) .
t t
In addition, if AS does not depend on t, the equation is even simpler:

AI (t)
ih = [AI (t), H0 ] .
t
118 Apendix C. Equivalent descriptions
Appendix D

The Time evolution operator

For a general quantum system, we define the time evolution operator as one that satisfies the
following relationship:
|(t) = U (t, t0 )|(t0 ) (D.1)

with the properties:

U (t3 , t1 ) = U (t3 , t2 ) U (t2 , t1 ) , t1 t2 t3 ,


1
U (t, t0 ) = U (t0 , t) ,
U (t, t) = I . (D.2)

This operator must satisfy the dynamic equation:

U (t, t0 )
ih = H U (t, t0 ) . (D.3)
t
If H does not depend on time, U can be written as:

U (t, t0 ) = exp [iH(t t0 )/h] . (D.4)

However, if H depends on time, U can be expressed formally as:


t
i
U (t, t0 ) = I dt HU (t , t0 ) (D.5)
h t0

or, equivalently:
[ ( t )]
i
U (t, t0 ) = T exp dt H(t ) , t > t0 (D.6)
h t0

where T is the Dyson time ordering operator, defined by:


{
H(t1 )H(t2 ) si t1 > t2
T [H(t1 )H(t2 )] = T [H(t2 )H(t1 )] = (D.7)
H(t2 )H(t1 ) si t2 > t1

Indeed, as:

deO() 1 dO O
= de(1)O e (D.8)
d 0 d
120 Apendix D. The Time evolution operator

(which we can see by the expansions of the exponential on both sides) we have:
[ ( )]
U (t, t0 ) i t
ih = ih T exp dt H(t )
t t h t0
[ ( )]
i t
= ihT exp dt H(t )
t h t0
1 t t
(1) hi dt H(t ) hi dt H(t )
= T de t0 H(t)e t0
0
t 1
hi dt H(t )
= H(t)T e t0 d
0
= HU (t, t0 ) (D.9)
145

References
J.L. Cardy. Scaling and renormalization in statistical physics. Cambridge Univ Pr, 1996.

F. Halzen and A.D. Martin. Quark & Leptons: An introductory Course in Modern Particle
Physics. Wiley-India, 2008.

M.E. Peskin and D.V. Schroeder. An introduction to quantum eld theory. Westview press,
1995.

K.G. Wilson. The renormalization group: Critical phenomena and the Kondo problem.
Reviews of Modern Physics, 47(4):773, 1975.

Potrebbero piacerti anche