Anda di halaman 1dari 57

Relativistic Quantum Fields 2

Mark Hindmarsh
University of Sussex
m.b.hindmarsh@sussex.ac.uk
http://www.pact.cpes.susx.ac.uk/users/markh/Teaching/

Spring Term 2002


Monday 4pm Arundel 1A
Tuesday 11:30pm Arundel 1A
Friday 11:30am Arundel 1A
PDF version of the lecture notes available:
http://www.pact.cpes.susx.ac.uk/users/markh/RQF2/rqf2.pdf

Aims and Learning Outcomes

Aim: to enable 1st year particle physics graduate students to understand


quantum field theory as used in current research in particle physics.
Learning outcomes: to be able to give an account of the quantisation pro-
cedures for fermions and gauge fields; to understand the gauge principle in
particle physics; to be able to compute 1-loop Feynman diagrams in renor-
malisable field theories; to understand the theortical basis of the Standard
Model of particle physics.
Skills: students will improve their problem-solving; and gain necessary
subject-specific concepts (gauge field theory, renormalisation, symmetry and
symmetry-breaking).

Syllabus

• Fermions (Ryder 4.3, 6.7). Dirac equation in Lagrangian formulation;


Canonical quantisation; Grassman variables; Fermionic path integral;
Feynman rules.

1
• Gauge field theory (Ryder 4.4, Cheng & Li Ch 8). Internal symme-
tries; Gauge symmetry 1: Abelian; The electromagnetic field; Gauge
symmetry 2: non-Abelian; Non-Abelian gauge field theory; Sponta-
neous symmetry breaking 1: Abelian; Spontaneous symmetry breaking
2: SU(2).
• Quantum gauge theory (Cheng & Li Ch 9, Ryder Ch7). Path-integral
quantisation; Fade’ev-Popov procedure, ghosts; Feynman rules in co-
variant gauge.
• Electroweak theory (Cheng & Li Ch 11). 1-family SU(2)×U(1) La-
grangian; Higgs mechanism; Mass spectrum; Family replication.
• QCD (Cheng & Li Ch 10). QCD Lagrangian and symmetries; Asymp-
totic freedom; Anomalies

Teaching methods
There will be 3 lectures a week throughout the term. Problem sheets will
be given every two weeks.

Assessment
Problem sheets will count for 100% of the total mark for the course, with a
25% weighting for a take-home exam at the end of the course.

Reading list

* Gauge theory of elementary particle physics, T.P. Cheng & L.F. Li


(O.U.P., Oxford, 1984).
* Quantum Field Theory, L. Ryder (C.U.P., Cambridge, 1984).
An introduction to quantum field theory, Michael E. Peskin, Daniel V.
Schroeder (Addison-Wesley, Reading, Mass; 1995).
Quantum field theory: a modern introduction, Michio Kaku (O.U.P.,
Oxford, 1993)
Quantum Field Theory, C. Itzykson and J.-B. Zuber (McGraw-Hill,
New York, 1980).
There are quite a few errors in Peskin and Schroeder which are supposed to
be corrected in the next printing of the book. Meanwhile, a list can be found
at

2
http://www.slac.stanford.edu/~mpeskin/QFT.html
The other books are complementary in some way. Kaku is very complete
but rather rushed. Itzykson & Zuber is compendious but lacks a coherent
development of the subject.

Prerequisites
Relativistic Quantum Fields 1, Further Quantum Mechanics.

Course Lecturer

Mark Hindmarsh, CPES, Arundel 213.


Telephone: 8934
E-mail: m.b.hindmarsh@sussex.ac.uk
Course web: http://www.pact.cpes.susx.ac.uk/users/markh/RQF2/
Office hour: Tuesday 2–3pm, or by arrangement (email is best).

3
Contents
1 Fermions 6
1.1 Plane wave solutions . . . . . . . . . . . . . . . . . . . . . . . 7
1.2 The interpretation of negative energy states . . . . . . . . . . 9
1.3 Quantising the spinor field . . . . . . . . . . . . . . . . . . . . 10
1.4 Conserved charge . . . . . . . . . . . . . . . . . . . . . . . . . 13

2 Gauge field theory 14


2.1 Internal symmetries . . . . . . . . . . . . . . . . . . . . . . . . 14
2.2 Gauge symmetry 1 . . . . . . . . . . . . . . . . . . . . . . . . 17
2.3 Maxwell’s equations in covariant form . . . . . . . . . . . . . . 19
2.4 Lagrangian formulation . . . . . . . . . . . . . . . . . . . . . . 21
2.5 Gauges . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.6 Coulomb gauge solutions . . . . . . . . . . . . . . . . . . . . . 23
2.7 Lorentz gauge solutions . . . . . . . . . . . . . . . . . . . . . . 24
2.8 Canonical quantisation in the Lorentz gauge . . . . . . . . . . 26
2.9 Feynman propagator in Lorentz gauge . . . . . . . . . . . . . 28

3 Quantum gauge theory 29

4 Electroweak theory 29

5 QCD 29

6 Renormalisation 29

A Pauli matrices 29

B Dirac matrices 30
B.1 Standard representation of Dirac matrices . . . . . . . . . . . 30

C Identities for Dirac matrices 31

D Feynman rules 31

E Real scalar field 33

F Complex scalar field 33

G U(1) gauge field 35

H Spinor field with U(1) gauge symmetry 35

4
I Non-Abelian gauge field 37

J Complex scalar field with non-abelian gauge symmetry 40

K Spinor field with non-abelian gauge symmetry 42

L Problem Sheets 44
L.1 Problem Sheet 1. . . . . . . . . . . . . . . . . . . . . . . . . . 44
L.2 Problem Sheet 2. . . . . . . . . . . . . . . . . . . . . . . . . . 46
L.3 Problem Sheet 3. . . . . . . . . . . . . . . . . . . . . . . . . . 49
L.4 Problem Sheet 4. . . . . . . . . . . . . . . . . . . . . . . . . . 52
L.5 Problem Sheet 5. . . . . . . . . . . . . . . . . . . . . . . . . . 54

5
1 Fermions
Dirac knew of the problem with the probability interpretation of the Klein-
Gordon equation, and traced it to the fact that it was second order in time
derivatives. He therefore set out to find a relativistic wave equation with
only one time derivative. The requirement that the form of the equation be
unchanged under Lorentz transformations, which mix up ∂/∂t and ∇, means
that the equation must be first order in spatial derivatives as well. Hence
Dirac proposed a relativistic free particle wave equation

ψ = −iα·∇ψ + βmψ.
i (1)
∂t
There clearly has to be something rather special about the objects α, β and
ψ in order that Lorentz covariance be preserved. In fact, α and β are 4×4
Hermitean matrices, and ψ is an object with four components called a spinor.
Now, we want a wave equation with plane wave solutions which satisfy
the relativistic energy-momentum relation E 2 = p2 + m2 . To see how this
property emerges from the Dirac equation, we act on both sides of (1) with
(i∂t − iα · ∇), to obtain
! ! !
∂ ∂ ∂
i − iα · ∇ i + iα · ∇ ψ = m i − iα · ∇ βψ. (2)
∂t ∂t ∂t
At this point, it is more convenient to switch to index notation for the 3-
vectors α and ∇, noting that α · ∇ = αi ∂i . We expand out the brackets on
the left hand side of (2), and add and subtract −iβαi ∂i on the right hand
side. The result is
∂2
! " ! #
i j ∂ i i i
− 2 + α α ∂i ∂j ψ = m i − iβα ∂i ψ + iβα ∂i ψ + iα β∂i ψ . (3)
∂t ∂t
It is now useful to define the anticommutator, represented by curly brackets:
{A, B} = AB + BA. (4)
(Sometimes you will also see the anticommutator written as [A, B]+ ). We
rewrite the left hand side as (−∂t2 + 21 αi αj ∂i ∂j + 12 αj αi ∂j ∂i )ψ, which we are
entitled to do by renaming the indices on α and ∂, to obtain
∂2
!
1
− 2 + {αi , αj }∂i ∂j ψ = m2 β 2 ψ + i{β, αi }∂i ψ. (5)
∂t 2
In order to reproduce the relativistic relation between energy and momentum,
we must end up with an equation like the Klein-Gordon equation. If αi and
β satisfy
{αi , αj } = δ ij , {β, αi } = 0, β 2 = 1, (6)

6
where 1 is the 4×4 identity matrix, ψ satisfies the equation

∂2
!
− 2 + ∇2 ψ = m2 ψ. (7)
∂t

Hence, each of the four components of ψ satisfy the Klein-Gordon equation.


We can make the Dirac equation more explicitly relativistic by defining
four new 4×4 matrices:

γ 0 = β, γ i = βαi . (8)

If we multiply both sides of the Dirac equation (1) by β, we obtain

(iγ µ ∂µ − m1)ψ = 0. (9)

The conditions (6) on αi and β are neatly unified into the matrix equation

{γ µ , γ ν } = 2η µν 1. (10)

1.1 Plane wave solutions


We know that plane wave solutions exist, but what are they? We therefore
look for solutions of the form

ψ(x) = ue−ip·x , (11)

with u a constant 4-component spinor. Although it appears from (7) that u


is arbitrary, this is not in fact true, as we shall see.
Substituting our plane wave ansatz (11) into the Dirac equation (9), we
find
(γ µ pµ − m)ue−ip·x = 0. (12)
At this point we shall introduce another piece of notation: we shall use the
“slash” symbol 6 to denote contraction of a 4-vector with the Dirac gamma
matrices:
aµ γ µ =6 a. (13)
Thus u satisfies the equation

(6 p − m)u = 0. (14)

Let us premultiply this equation by (6 p + m), and use an important identity,

6 a 6 b = a · b 1, (15)

7
to show that
(6 p + m)(6 p − m)u = (p2 − m2 )u = 0. (16)
This brings us back to the relation E 2 = p2 + m2 , although by a rather more
rapid route. It is conventional to choose p0 ≡ E always to be positive, in
which case we must also include solutions proportional to e+ip·x separately.
We write them
ψ(x) = veip·x , (17)
which satisfy
(6 p + m)v = 0. (18)
Solutions with time dependence e−iEt (eiEt ) are called positive (negative)
energy solutions.
We can use the first equality in Eq. (16) to show that the following spinors
are solutions to the Dirac equations for u and v:
! !
χ± 0
u± (p) = N (6 p + m) v± (p) = −N (6 p − m)
0 χ±

where N is a normalisation factor, and we have defined χ+ and χ− to span


the space of 2-component spinors (which means they are spinor spanners):
! !
1 0
χ+ = , χ− = . (19)
0 1

Any solution to the Dirac equation can be written as a superposition of the


four four-component basis spinors u± and v± .
To proceed further we need an explicit representation for the gamma
matrices, and we will use the so-called “standard representation” given in
Appendix B.1. It will turn out to be convenient to choose the normalisation
factor to be √
N = 1/ (E + m). (20)
The basis for the 4-component positive energy solutions can therefore be
written !
1 (E + m)χ±
u± = √ , (21)
E+m p · σχ±
and the negative energy basis is
!
1 p · σχ±
v± = √ . (22)
E+m (E + m)χ±

There are a couple of remarks that are worth making about the solutions (21)
and (22). Firstly, when the 3-momentum is small, that is when |p|  m,

8
two of the components are much smaller than the other two. For the positive
energy solutions, it is the lower two components that are negligible, while the
opposite is true for the negative energy solutions. Secondly, the appearance
of two degrees of freedom for each set of solutions (that is, χ± ) demands some
explanation. One can verify that χ± are eigenstates of the Pauli matrix σ 3 ,
with eigenvalues ±1:
σ 3 χ± = ±χ± . (23)
This degeneracy is actually a result of the fact that the particles described
by the Dirac equation (such as the electron) have an extra property, spin,
which is a type of angular momentum intrinsic to the particle itself, unlike
ordinary angular momentum which particles gain by virtue of rotation. Spin
is quantised in half-integer units, with the spin of the electron being s = 21 .
As with other types of quantised angular momentum, the value of the spin
projected along any axis (such as the z axis) takes the values s, s − 1, . . . , −s.
Thus Dirac-type particles have only two spin states, which we conventionally
call “up” and “down”.
With the given normalisation convention, the spinors have the properties

u†A uB = 2EδAB , vA† vB = 2EδAB (24)

We define the adjoint spinors ū and v̄ by

ūA = u†A γ 0 , v̄A = vA† γ 0 . (25)

The adjoint spinors have the further properties

ūA uB = 2mδAB , v̄A vB = −2mδAB . (26)

Note that this is a different normalisation convention than that used by


Greiner and Reinhardt.

1.2 The interpretation of negative energy states


In the last section we called half of the solutions “negative energy” solutions
without proper explanation. Of course, the explanation is that they appear
to represent quantum states with negative energy, for if one acts with the
quantum mechanical energy operator, one obtains a negative number:

i (veip·x ) = −E(veip·x ), (27)
∂t
1
with E = |(p2 + m2 ) 2 |. This raises an ugly possibility: that an electron with
positive energy E1 could decay into one with negative energy −E2 , radiating

9
a photon with energy E1 + E2 . There indeed seems to be nothing to stop the
energy of the electron from tumbling down towards minus infinity, releasing
an infinite amount of energy in the process. This is clearly nonsense. Dirac’s
way of circumventing this problem was to suppose that all the negative en-
ergy states are filled already, and the Pauli exclusion principle prevents a
positive energy electron from radiating a photon and occupying a filled nega-
tive energy state. One can still imagine exciting one of these negative energy
electrons into a postive energy state, whereupon it becomes a real electron,
leaving behind a hole. A hole is an absence of a negative electric charge,
which has positive charge. Thus Dirac realised that his theory predicted
the existence of a positively charged spin 12 particle with exactly the same
mass as the electron. Initially, Dirac identified this particle with the proton,
hoping that somehow Coulomb interactions would provide a mass difference,
but after others (including Weyl, Oppenheimer, and Tamm) showed that this
couldn’t be right, he eventually had to abandon this conservative position
and predict a new particle now know as the positron. Its discovery in 1932
by Carl Anderson in cosmic rays settled the issue.
However, this picture of a negative energy “sea” of electrons leaves much
to be desired. Bosons, particles with integer spin such as the photon (spin
1) and the pion (spin 0) do not obey the Pauli Exclusion Principle and
yet they still have negative energy states. Thus there can be no Dirac sea
picture and we are left where we were before. It is only if we abandon simple
wave mechanics and turn to quantum field theory that we find a satisfactory
resolution of this problem.

1.3 Quantising the spinor field


The Lagrangian density for a massive spinor field is
L = iψ̄ 6 ∂ψ − mψ̄ψ, (28)
where we recall that in the standard representation of the Dirac gamma
matrices the adjoint spinor ψ̄ = ψ † γ 0 . Rather like the complex scalar field,
one can obtain the field equation (which is the Dirac equation) by treating ψ̄
and ψ as independent quantities, and demand that the action be stationary
with respect to arbitrary variations in either. The variation of the action
with respect to ψ̄ gives
Z
∂L
δS = d4 x δ ψ̄ = 0, (29)
∂ ψ̄
(the Lagrangian is not a function of ∂µ ψ̄ in this formulation). Hence
Z
d4 x δ ψ̄(i 6 ∂ψ − mψ) = 0, (30)

10
from which we derive the Dirac equation i 6 ∂ψ − mψ = 0. Similarly, we can
vary with respect to ψ and obtain
Z
δS = d4 x (iψ̄ 6 ∂δψ − mψ̄δψ). (31)

Integrating by parts we find


Z Z
4 µ
δS = d x (−i∂µ ψ̄γ − mψ̄)δψ + dSµ iψ̄γ µ δψ, (32)

where the last term is an integral over the space-time surface at spatial
infinity (|x| → ∞) with end-caps at |t| → ∞. As usual, we suppose that the
variations die away at infinity so that we can drop the surface term, so we
recover the equation for the adjoint spinor i∂µ ψ̄γ µ + mψ̄ = 0.
The momentum conjugate to ψ is given by
∂L
π= = iψ̄γ 0 , (33)
∂ ψ̇
which shows that in the standard reresentation ψ and iψ † are canonically
conjugate variables. Thus we can find the Hamiltonian density:

H = π ψ̇ − L = −iψ̄γ k ∂k ψ + mψ̄ψ. (34)

The quantisation of the spinor field follows a familiar pattern. We first of


all suppose that ψ(x) is an operator satisfying some commutation relations,
acting on some space of quantum states. We specify the equal time com-
mutation relations, and then try to find the possible states. The problem
with the spinor field is that it is not obvious from the outset what commu-
tation relations to impose on the field operator, and the obvious relation
[ψ(t, x), iψ † (t, x0 )] = iδ(x − x0 ) is not in fact correct.
To see why this is so, let us compute the zero-point energy of the field.
Firstly, we expand the field in terms of its operator-valued Fourier coefficients:
X Z d3 p 1  −ip·x † ip·x

ψ(x) = c A (p)u A (p)e + dA (p)v A (p)e , (35)
A=± (2π)3 2E

where E 2 = p2 + m2 , and uA (p) and va (p) are 4-component spinors which


satisfy
(6 p − m)uA (p) = 0, (6 p + m)vA (p) = 0. (36)
Similarly, the conjugate spinor has the expansion
X Z d3 p 1  † +ip·x −ip·x

ψ̄(x) = c (p)ū A (p)e + dA (p)v̄ A (p)e . (37)
A=± (2π)3 2E A

11
We can now compute the Hamiltonian, which is the spatial integral of the
Hamiltonian density H, or
Z
H= d3 x (−iψ̄γ i ∂i ψ + mψ̄ψ) (38)

The end result is


XZ d3 p 1 
† †

H= E c A (p)c A (p) − dA (p)dA (p) . (39)
A (2π)3 2E

The vacuum state is traditionally defined as the state annihilated by all


annihilation operators cA (p) and dA (p):

cA (p)|0i = 0, dA (p)|0i = 0. (40)

We expect an infinite zero point energy again, removed by normal ordering.


However, H is not positive definite, and nor is :H: unless we define the
commutators of the operators such that

:dA (p)d†A (p): = −d†A (p)dA (p). (41)

This means that the fermionic creation and annihilation operators must an-
ticommute, and we quantise the spinor field by anticommutation relations

{dA (p), d†B (p0 )} = 2EδAB δ̄(p − p0 ), (42)


{cA (p), c†B (p0 )} = 2EδAB δ̄(p − p0 ). (43)

All other anticommutator brackets vanish:

{cA (p), cB (p0 )} = 0 = {dA (p), dB (p0 )} (44)


{c†A (p), c†B (p0 )} = 0 = {d†A (p), d†B (p0 )} (45)
{cA (p), d†B (p0 )} = 0 = {c†A (p), dB (p0 )}. (46)

There are four kinds of 1-particle states in this theory, created by the four
operators c†A (p) and d†B (p). We write them

|p, A; 0i = c†A (p)|0i, |0; p, Ai = d†A (p)|0i. (47)

As you can probably guess from the way the states are built up for the
complex scalar field, c†A (p) creates a particle with momentum p and spin
state A, while d†A (p) creates an antiparticle.
Let us examine the 2-particle state

|p1 , A1 , p2 , A2 ; 0i = c†A1 (p1 )c†A2 (p2 )|0i. (48)

12
If we interchange the particles, we must interchange the creation operators,
which gives the state a relative minus sign:

|p2 , A2 , p1 , A1 ; 0i = c†A2 (p2 )c†A1 (p1 )|0i, (49)


= −c†A1 (p1 )c†A2 (p2 )|0i, (50)
= −|p1 , A1 , p2 , A2 ; 0i. (51)

Thus we see something that has to be taken for granted in ordinary quan-
tum mechanics: the states of the spinor or Dirac field are antisymmetric
under particle interchange. We call these states fermionic, and the particles
fermions. Furthermore, suppose we take both momenta and both spin states
equal to p and A respectively: we then find

|p, A, p, A; 0i = −|p, A, p, A; 0i = 0. (52)

Hence two fermions can never be in the same state. This is precisely the
Pauli exclusion principle, which is fundamental to the understanding of the
physics of the atom. On the other hands, bosons, like the scalar particle we
considered previously, have commuting field operators, and so any number
of them can be in the same state.
Armed with the anticommutation relations for the creation and annihila-
tion operators, and the spinor completeness relations

uA A
vaA v̄bA = (γ · p)ab − mδab ,
X X
a ūb = (γ · p)ab + mδab , (53)
A A

the fields can be shown to satisfy the equal time anticommutation relations

{ψa (t, x), ψ̄b (t, x0 )} = ih̄δab δ(x − x0 ), (54)


{ψa (t, x), ψb (t, x0 )} = 0 = {ψ̄a (t, x), ψ̄b (t, x0 )} (55)

where the factor of h̄ has been revived, and labels a and b have been put in
to keep track of the rows and columns of the spinors ψ and ψ̄.

1.4 Conserved charge


One can show straightforwardly from the Dirac equation (and its adjoint)
that the 4-vector quantity

j µ (x) = ψ̄(x)γ µ ψ(x) (56)

satisfies a continuity relation ∂ · j = 0. Hence there is a conserved charge


Z Z
Q= 3 0
d x ψ̄γ ψ = d3 x ψ † ψ. (57)

13
By substituting the plane wave expansion for the Dirac field operator, one
can show that
XZ d3 p 1  † †

Q= c (p)c A (p) − d (p)dA (p) . (58)
A (2π)3 2E A A

Applying this operator to to the single particle states, one can show that

Q|p, A; 0i = +|p, A; 0i, Q|0, p, Ai = −|0; p, Ai. (59)

Hence the states created by c†A (p) have equal and opposite charge to those
created by d†A (p), which fits in with their interpretation as particles and
antiparticles.
One can also show that the existence of this conserved charge is a conse-
quence of a symmetry of the Lagrangian, as we should expect from Noether’s
theorem. This symmetry is

ψ(x) → ψ 0 (x) = eiθ ψ(x), ψ̄(x) → ψ̄ 0 (x) = e−iθ ψ̄(x). (60)

2 Gauge field theory


2.1 Internal symmetries
Recall Noether’s theorem for a set of N real fields:
For every transformation φa (x) → φ0a (x0 ) which leaves the action S[φa ]
invariant, there is a conserved current. Defining the total variation in the
field, which is assumed infinitesimal,

δφa (x) = φ0a (x0 ) − φa (x), (61)

one can find an (infinitesimal) 4-vector f µ which satisfies a continuity equa-


tion ∂µ f µ = 0, given by
!
∂L ∂L
fµ = δφa (x) − ∂ν φa (x) − Lδνµ δxν . (62)
∂(∂µ φa ) ∂(∂µ φa )

We studied in RQF1 the case of coordinate translation symmetry, x → x0 =


x+δx, with δxµ = µ , a constant but arbitrary infinitesimal 4-vector. This re-
sulted in four conserved currents, which were assembled into a rank-2 tensor,
the energy-momentum tensor
∂L
θµ ν = ∂ν φa (x) − Lδνµ . (63)
∂(∂µ φa )

14
In this section we will study internal symmetries, which transform fields
amongst themselves:
δφa = Tab φb , δxµ = 0 (64)
The non-infinitesimal matrix Tab is called the generator the symmetry.
With this transformation it is straightforward to show that the following
non-infinitesimal 4-vector is conserved:
∂L
jµ = Tab φb . (65)
∂(∂µ φa )
We can extend these considerations to N complex scalar fields. For this we
need to recall that when differentiating with respect to a complex variable
z, one treats z and z̄ as independent. Given this fact, it follows that the
infinitesimal conserved current which follows from a symmetry of the action
S[φa , φ̄a ] is
∂L ∂L
fµ = δφa (x) + δ φ̄a (x)
∂(∂µ φa ) ∂(∂µ φ̄a )
!
∂L ∂L
− ∂ν φa (x) + ∂ν φ̄a (x) − Lδνµ δxν . (66)
∂(∂µ φa ) ∂(∂µ φ̄a )
Internal symmetry transformations for complex fields take the form

δφa = Tab φb , δ φ̄a = Tab φ̄b , δxµ = 0. (67)

Example 1. Consider a theory with two real fields φ1 and φ2 , and an action
1
Z  
S= d4 x ∂µ φa ∂ µ φa − V (φ) , (68)
2
q
~ The
where φ = φ21 + φ22 is the magnitude of the 2-dimensional field vector φ.
action is clearly symmetric under rotations in the 2-dimensional “internal”
space of the field,
φ~→φ ~ 0 = Ωφ,
~ (69)
where Ω is a 2×2 orthogonal matrix
!
cos θ sin θ
Ω= . (70)
− sin θ cos θ

If the transformation is infinitesimal, we have


!
~'θ 0 1
δφ . (71)
−1 0

15
Using the antisymmetric symbol ab , defined by 12 = −21 = 1 and 11 =
22 = 0, we can rewrite the infinitesimal change in the field in component
form,
δφa = θab φb . (72)
In order to evaluate j µ we require
∂L
= ∂ µ φa . (73)
∂(∂µ φa )

Hence the conserved current in this theory is

j µ = ∂ µ φa ab φb = (∂ µ φ1 )φ2 − (∂ µ φ2 )φ1 . (74)

Note that the set of 2×2 orthogonal matrices form a group under matrix
multiplication. We have been considering those matrices which are contin-
uously connected to the identity, and therefore have unit determinant. The
group of 2×2 orthogonal matrices with unit determinant is called SO(2).

Example 2. Let us consider N = 1 complex scalar fields φ with action


Z  
S= d4 x ∂µ φ̄∂µ φ − V (|φ|) (75)

where |φ| is the modulus of φ. This is clearly invariant under the transforma-
tion φ → φ0 = e−iθ φ, i.e. under multiplication by complex numbers of unit
modulus. If θ is infinitesimal, we have

δφ = −iθφ, δ φ̄ = iθφ. (76)

In order to calculate the conserved current, we need the differentials


∂L ∂L
= ∂ µ φa , = ∂ µ φ̄a . (77)
∂(∂µ φa ) ∂(∂µ φ̄a )

It is now straightforward to show that

j µ = −i(∂ µ φ̄)φ + i(∂ µ φ)φ̄ (78)

is conserved. Note that the set of numbers of unit modulus form a group
under ordinary complex multiplication,
√ which is called U(1). Note also that
if one writes φ = (φ1 +iφ2 )/ 2, with φ1 and φ2 real, one can recover both the
action and the conserved current of the SO(2) scalar field theory. This is just
a consequence of the fact that the groups U(1) and SO(2) are isomorphic.

16
Example 3. Consider now the spinor field, whose action is
Z
S= d4 x ψ̄(i 6 ∂ − m)ψ. (79)

Again, this has the obvious symmetry ψ → ψ 0 = e−iθ ψ, which in infinitesimal


form is δψ = −iθψ and δ ψ̄ = iθψ̄. In order to evaluate the conserved current
we need
∂L ∂L
= iψ̄γ µ , = 0, (80)
∂(∂µ ψa ) ∂(∂µ φ̄a )
from which we can quickly infer that
j µ = ψ̄γ µ ψ (81)
is conserved. Note that one can show that the conserved charge Q = ψ̄γ 0 ψ
is equal to the difference in the numbers of particles and antiparticles:
Z
d¯3 p  † 
Q= cA (p)cA (p) − d†A (p)dA (p) , (82)
2E
which demonstrates explicitly that particles and antiparticles have opposite
charges.

Example 4. Finally, let us consider a massless spinor field, with action


Z
S= d4 x iψ̄ 6 ∂ψ. (83)

This also enjoys the symmetry ψ → ψ 0 = e−iθ ψ, but has an extra chiral
5
symmetry ψ → ψ 0 = e−iθ5 γ ψ, where γ 5 is defined in Appendix B. Note
5
that the adjoint spinor transforms as ψ̄ → ψ̄ 0 = ψ̄e−iθ5 γ , and so the mass
term ψ̄ψ is not invariant under this transformation. Similar considerations
to the previous examples show that the conserved current associated with
this symmetry is
j5µ = ψ̄γ µ γ 5 ψ. (84)

2.2 Gauge symmetry 1


If the parameters of a symmetry transformation are space-time dependent we
call it a local or gauge symmetry. Gauge theories have much richer structure
that their counterparts with global symmetries.
The prototypical gauge theory is electrodynamics. Let us consider “gaug-
ing” the massive spinor field, i.e. demanding that the U(1) global symmetry
discussed in the previous section becomes space-time dependent:
φ → φ0 = e−iα(x) φ. (85)

17
The simple spinor action is no longer invariant under this space-time depen-
dent transformation, as one can easily show:
Z
S → S0 = d4 x ψ̄eiα(x) (i 6 ∂ − m)e−iα(x) ψ,
Z
= S− d4 x ψ̄γ µ ψ∂µ α. (86)

In order to make the action symmetric under this transformation we can


introduce a 4-vector field Aµ with the transformation property
1
Aµ → A0µ = Aµ + ∂µ α, (87)
e
where e is a constant. Then one can show that
(∂µ + ieAµ )ψ → (∂µ + ieA0µ )ψ 0 = (∂µ + (ieAµ + i∂µ α)e−iα(x) ψ
= e−iα(x) (∂µ + ieAµ )ψ. (88)
Introducing the notation Dµ = (∂µ + ieAµ ), one sees that Dµ ψ transforms in
the same way as ψ under gauge transformations:
(Dµ ψ) → (Dµ ψ)0 = e−iα(x) (Dµ ψ). (89)
We call this special object the gauge covariant derivative.
Using the gauge covariant derivative, we can contruct a new action which
is invariant under gauge transformations:
Z
S= d4 x ψ̄(i 6 D − m)ψ. (90)

We say that this action is gauge invariant.


The dynamics of this theory is not very interesting as it stands. If we
want to treat Aµ as a dynamical field, and ask that the action be at an
extremum with respect to its variations:
δS δ Z h i
= d4 x ψ̄(i 6 ∂ − m)ψ − eψ̄γ µ ψAµ ψ
δAµ (x) δAµ (x)
Z
δAν (y)
= −e d4 y ψ̄γ µ ψ
δAµ (x)
µ
= −eψ̄(x)γ ψ(x) = 0. (91)
Thus we see that the current of the global symmetry must vanish if this
action is at an extremum. We need to supplement the action to obtain a
field equation for Aµ , in which case ψ̄(x)γ µ ψ(x) will act as a source.
In order to realise this project, we first need to study the formulation of
the field equations for vector fields – otherwise known as Maxwell’s equa-
tions.

18
2.3 Maxwell’s equations in covariant form
This section recaps some important results, and introduces the formulation
of the theory in an explicitly Lorentz covariant manner.
Firstly, we recall Maxwell’s equations in free space, writing them down
in natural units, in which the permittivity and permeability of free space µ0
and 0 are both unity:

Homogeneous Inhomogeneous
∇·B = 0 ∇·E = ρ (92)
∂ ∂
B+∇∧E=0 − E+∇∧B=j
∂t ∂t
The homogeneous and inhomogeneous (i.e. having a source term on the right
hand side) equations have differing status. The homogeneous equations imply
the existence of potentials φ and A from which the physically measurable
quantities E and B can be calculated. These potentials are specified only
up to a gauge transformation, φ → φ − Λ̇ and A → A + ∇Λ, where Λ is an
arbitrary function of space and time. The inhomogeneous equations imply
that the source terms must obey a current conservation equation, ρ̇+∇·j = 0.
To summarise:
Homogeneous Inhomogeneous
potentials φ, A current conservation
B = ∇∧A
(93)
E = −Ȧ − ∇φ
ρ̇ + ∇·j = 0
φ → φ − Λ̇
A → A + ∇Λ

All these quantities can be assembled into explicitly Lorentz covariant ob-
jects. The gauge potentials belong together in a 4-vector potential Aµ =
(φ, A), while the charge density ρ and the current density j can be put to-
gether into a 4-vector current density j µ = (ρ, j). Recall that putting quanti-
ties together into 4-vectors is not just a matter of notation: it means that the
quantities transform just like the space-time coordinates xµ under a Lorentz
transformation.
The electric and magnetic fields E and B also belong together in a Lorentz
covariant object, as they are mixed up by Lorentz transformations (an ob-
server moving through a magnetic field also sees an electric field). However,
this object cannot be a 4-vector as there are a total of 6 components of the
electric and magnetic fields, when they are taken together. In fact, the object
is an antisymmetric tensor, the field strength tensor F µν , whose entries are

19
as follows:
0 −E 1 −E 2 −E 3
 
 E1 0 −B 3 B 2 
F µν =  . (94)
 
2 3 1 
 E B 0 −B 
E 3 −B 2 B 1 0
Note that E 1 represents the first component of the electric field vector, which
in Cartesian coordinates is the x component. The reverse relations may be
written
1
E i = F i0 , B i = − ijk F jk , (95)
2
which introduces the Levi-Civita symbol ijk . The Levi-Civita symbol is
defined by 
 +1 if i 6= j 6= k cyclic,

ijk = −1 if i 6= j 6= k anticyclic, (96)


0 otherwise.
The field strength tensor has a neat expression in terms of the gauge potential
4-vector Aµ :
F µν = ∂ µ Aν − ∂ ν Aµ . (97)
Let us check this explicitly for the components corresponding to the electric
field:
F i0 = ∂ i A0 − ∂ 0 Ai = ∂ i A0 − Ȧi = −∂i φ − Ȧi = E i , (98)
where we have used the expression for the electric field in terms of the gauge
potentials in equation (93).
When expressed in terms of 4-vectors and tensors, electromagnetism looks
very simple and beautiful. For example, the covariant expression of the
current conservation equation is simply

∂µ j µ = 0. (99)

Maxwell’s equations become:


Homogeneous Inhomogeneous
λ µν µ νλ ν λµ
(100)
∂ F + ∂ F + ∂ F = 0, ∂µ F µν = j ν .

The first of these expressions looks as if it contains many more equations


than the 4 of the original homogeneous Maxwell equations. However, the
antisymmetry of the field strength tensor (F µν = −F νµ ) means that the
expression is trivial if any of the two indices are equal. Thus all the indices
λ, µ and ν must take different values, and the number of ways of choosing
three different numbers from a set of four is 4 C3 , which is equal to 4, precisely
the number of equations we started with.

20
We can use the four-dimensional Levi-Civita tensor to re-express the ho-
mogeneous equations more compactly. This tensor has four indices, and is
defined by

if µ 6= ν 6= ρ 6= σ, symmetric,
 +1

µνρσ = −1 if µ 6= ν 6= ρ 6= σ, antisymmetric, (101)


0 otherwise.

A symmetric permutation of the indices in one in which an even number of


indices are exchanged, while an antisymmetric permutation is one for which
an odd number of indices are exchanged. Thus, for example, 0123 = 1032 =
+1, 1023 = 0132 = −1, but 0012 = 0.
There is also a version with the indices raised:

µνρσ = η µα η νβ η ργ η σδ αβγδ , (102)

whose symmetric permutations take the value −1 and antisymmetric ones


+1.
Using this totally antisymmetric (i.e. antisymmetric on the exchange of
any two indices) tensor, the homogeneous Maxwell equations become

µνρσ ∂ ρ F µν = 0. (103)

2.4 Lagrangian formulation


We recall that the correct action to reproduce the inhomogeneous Maxwell
equations is
1
Z  
4 µν µ
S = d x − Fµν F − jµ A , (104)
4
where Fµν = ∂µ Aν − ∂ν Aµ . We firstly recall that the functional derivative of
a 4-vector field with respect to itself is

δAν (y)
= δνµ δ 4 (y − x). (105)
δAµ (x)

To obtain the field equations we extremise the action, finding


δS
= ∂ν F νµ (x) − j µ (x) = ∂ 2 Aµ (x) − ∂ µ ∂ · A(x) − j µ (x) = 0. (106)
δAµ (x)

21
2.5 Gauges
The physical quantities, the electric and magnetic fields, contained in the
field strength tensor Fµν , are invariant under a gauge transformation

Aµ (x) → A0µ (x) = Aµ (x) − ∂µ Λ(x), (107)

where Λ is an abitrary function of the spacetime coordinates xµ . Indeed,


using (97) one finds

F µν → ∂ µ (Aν − ∂ ν Λ) − ∂ ν (Aµ − ∂ µ Λ) = F µν − (∂ µ ∂ ν − ∂ ν ∂ µ )Λ = F µν . (108)

The last step follows because we can take the partial derivatives in any order.
There are two related problems with the theory as formulated above.

1. Gauge invariance means that there is no unique solution to the field


equations for a given source j µ .

2. If we wish to find the Hamiltonian of the theory, as is necessary for


canonical quantisation, we must first find the canonical momentum,
∂L
πµ = = Fµ0 . (109)
∂ Ȧµ
This means that π0 = 0, i.e. the canonical momentum associated with
the timelike component of the gauge field A0 is identically zero.

The solution to these problems begins with fixing the gauge: imposing a
constraint on the field to remove the freedom to make a gauge transformation.
For each constraint, the field equations may be solved.
There are in principle an infinite number of constraints one could apply,
but in practice three turn out to be convenient.

1. Temporal gauge A0 = 0.

2. Coulomb gauge ∇·A = 0.

3. Lorentz gauge ∂ · A = 0.

Only the last is obviously Lorentz invariant, and so is most useful for
relativistic field theory. The other two gauges are useful in other circum-
stances: for example, temporal gauge is the easiest to use for solving the
field equations numerically on a lattice, while Coulomb gauge is useful for
non-relativistic applications in atomic physics.

22
2.6 Coulomb gauge solutions
This is also called radiation gauge, and is defined by

∇·A = 0, (110)

which implies that the first of the inhomogeneous equations in (92) becomes

∇2 φ = −ρ. (111)

This gauge is therefore very useful in solving electrostatics problems, and we


shall use it later on when studying the Casimir effect. However, it is not so
often used in relativistic applications, as the gauge condition does not respect
Lorentz invariance.
The second of the inhomogeneous Maxwell equations becomes, on using
the gauge condition,
∂  
− −Ȧ − ∇φ − ∇2 A = j, (112)
∂t
which we may write as
∂2
!
− ∇2 A = jT , (113)
∂t2
where jT = j+∇φ̇. One can show, using the equation of current conservation
and Eq. (111), that
∇·jT = 0. (114)
This of course is necessary for the consistency of Eq. (113), because if one
takes the divergence of the left hand side one gets zero as well.
In free space, that is, in the absence of charges and currents, we may take
φ = 0, and the equation for the vector potential becomes
∂2
!

2
− ∇2 A = 0. (115)
∂t
One of the important discoveries of the last century was that this equation has
plane wave solutions, which carry energy and momentum: electromagnetic
waves. A plane wave solution with a particular wavenumber k may be written

A(t, x) = ae−iωt+ik·x , (116)

where a is a constant 3-vectors, and ω = |k|.


The gauge condition (110) imposes a condition on the vector a, for

∇·A = ik·a e−iωt+ik·x = 0, (117)

23
which implies
k·a = 0. (118)
Hence the gauge potential A is orthogonal to the wave vector k. This means
that both the electric and magnetic fields are also orthogonal: we say that
the waves are transverse.

2.7 Lorentz gauge solutions


This is very often the gauge to choose if one is interested in wave propagation
in electromagnetism. It is defined by

∂µ Aµ = 0, (119)

and is manifestly Lorentz invariant. In this gauge the equation of motion for
the gauge field is simplified, for

∂µ F µν = ∂µ ∂ µ Aν − ∂µ ∂ ν Aµ = ∂ 2 Aν = j ν . (120)

However, an irritating feature of this gauge is that it does not quite specify
Aµ fully. One can still make a gauge transformation Aµ → Aµ − ∂ µ Λ which
satisfies the Lorentz gauge condition (119), as long as the function Λ satisfies
∂ 2 Λ = 0. Such functions are called harmonic. In classical field theory this is
not too much of a problem, but in setting up the quantum theory of gauge
fields care must be taken.
In free space, the field equation is (120)

∂2
!
2
∂ A ≡µ
− ∇2 Aµ = 0. (121)
∂t2
A particular solution is
Aµ (t, x) = aµ e−ik·x , (122)
where k 0 = ωk = |k|, and aµ is a constant 4-vector.
The Lorentz gauge condition (119) gives ∂ · A = −ik · ae−ik·x = 0, which
implies
k · a = 0. (123)
Once again the field is orthogonal to the wavenumber k, but this time in the
4-vector sense.
The general solution may be constructed from a superposition of plane
wave solutions
Z
d¯3 k  µ 
Aµ (x) = a (k)e−ik·x + a∗ µ (k)eik·x , (124)
2ωk

24
where the amplitudes aµ (k) must all satisfy k · a(k) = 0.
It would be more convenient to be able to choose the amplitudes of the
components of the gauge field independently, and to this end we introduce a
set of four basis 4-vectors, or polarisation vectors, A µ (k), with A = 0, 1, 2, 3,
for each wavevector k. They must satisfy a completeness condition in order
for them to be considered as basis vectors, and it is also convenient to make
them orthonormal, in a 4-vector sense.

A ∗µ (k)B ν (k)ηµν = ηAB (Orthonormality), (125)


A ∗µ (k)B ν (k)η AB = η µν (Completeness). (126)

A good choice is

0 = (1; 0),
1 = (0, n̂ × k)/|n̂ × k|,
2 = (0, k × (n̂ × k))/|k × (n̂ × k)|,
3 = (0; k̂),
(127)

where n̂ = (0, 0, 1) is a unit 3-vector. We refer to these polarisation vec-


tors as timelike (A = 0), transverse (A = 1, 2), and longitudinal (A = 3)
respectively.
Using this basis we can write aµ = aA A µ , and find that the gauge condi-
tion implies k · a = k 0 a0 − |k|a3 = 0, or

a0 (k) = a3 (k). (128)

Thus the general solution is now more easily written down as


Z
d¯3 k  A 
Aµ (x) = a (k)A µ (k)e−ik·x + aA (k)∗A µ (k)eik·x . (129)
2ωk
There now appears to be, for each k, three independent degrees of freedom,
down from the original four. However, one of the amplitudes, a3 , does not
affect the physical value of the electric and magnetic fields. This is a result of
the remaining freedom in the Lorentz gauge to make gauge transformations
with a harmonic function Λ(x), which may be decomposed into plane waves,
according to
Z
d¯3 k  µ 
Λ(x) = λ (k)e−ik·x + λµ (k)eik·x . (130)
2ωk
Hence under such a gauge transformation,
µ
aµ (k) → a0 (k) = aµ (k) − ik µ λ(k). (131)

25
By multiplying with 3µ we see that this is equivalent to shifting a3 (k),
3
a3 (k) → a0 (k) = a3 (k) − i|k|λ(k). (132)

Thus although the Lorentz gauge does not entirely remove the freedom to
make gauge transformations, it does confine the ambiguity in the solution to
one and only one of the sets of constants in the solution. There are only two
physical amplitudes to choose, a1 and a2 .
Let us end this section by calculating the Hamiltonian in the Lorentz
gauge.
Z  
HLg = d3 x πµ Ȧµ − LLg
1Z 3
= − d x(Ȧµ Ȧµ + ∇Aµ ·∇Aµ ). (133)
2
We already notice from this expression that the Hamiltonian is not positive
definite, as the timelike component of the gauge field A0 appears with a
negative sign. It turns out that we have already cured this problem in the
classical theory. To see this, let us substitute the plane wave expansion of
the field operator, obtaining
Z
d¯3 k
HLg = − ωk ηAB a∗A (k)aB (k). (134)
2ωk

We saw earlier that the Lorentz gauge condition Eq. 119 implies a0 = a3 ,
and so we find that
X Z d¯3 k
HLg = ωk a∗A (k)aA (k), (135)
A=1,2 2ωk

which is clearly positive definite, and depends only on the physical tranversely
polarised components.

2.8 Canonical quantisation in the Lorentz gauge


In order to quantise this theory in the canonical manner, we must first iden-
tify the canonical momentum conjugate to the degrees of freedom, impose
canonical commuation relations, identify the ladder operators and find their
commutation relations, and construct a Fock space of states. We should also
calculate the Hamiltonian, and verify that it is positive definite.
As one might imagine, gauge invariance poses problems for the quantum
theory as well. Firstly, we have the problem mentioned in Section 2.5, that

26
the canonical momentum vanishes. This turns out to be cured by fixing the
gauge, as we can verify in the Lorentz gauge:
LLg
πµ = = −Ȧµ . (136)
∂ Ȧµ
Hence the equal time canonical commutation relations can be guessed to be
those for four independent real fields

[Aµ (t, x), πν (t, x0 )] = iδνµ δ 3 (x − x0 ), (137)


[Aµ (t, x), Aν (t, x0 )] = 0, [πµ (t, x), πν (t, x0 )] = 0. (138)

By substituting the plane wave expansion for the field operator Aµ (x), one
can find (after some algebra) that

[aA (k), a∗B (k0 )] = −η AB 2ωk δ̄ 3 (k − k0 ), (139)


[aA (k), aB (k0 )] = 0, [a∗A (k), a∗B (k0 )] = 0. (140)

Hence we have four pairs of ladder operators, one pair for each polarisation.
We define the vacuum state |0i by

aA (k)|0i = 0 ∀k, A. (141)

and we can construct excited states by acting with the raising operators
a∗A (k). For example,
|k, Ai = a∗A (k)|0i. (142)
How do we impose the gauge constraint in the quantum theory? Firstly,
it is clear that demanding ∂ · A = 0 is too strong a condition on the operator
Aµ , as it is inconsistent with the equal time canonical commutation relations
Eq. 137. A weaker condition is to ask that matrix elements of the gauge
condition vanish. This we cannot do for every possible state, but we only
accept as physical states for which this is true: i.e.

hφ0 |∂ · A|φi = 0 (143)

for any two physical states |φi and |φ0 i. Any equivalent condition is to
demand that the positive frequency part of the gauge condition gives zero
when acting on a physical state:

∂ · A(+) |φi = 0, (144)

where Z
d¯3 k
∂ · A(+) = − ik · ae−ik·x . (145)
2ωk

27
Hence physical states must satisfy

(a0 (k) − a3 (k))|φi = 0, (146)

which is the equivalent of Eq. (128) in the quantum theory.


Finally, we shall check that the Hamiltonian is positive definite for phys-
ical states. We must deal with the infinity in the vacuum energy by normal
ordering, and study the matrix elements of the operator
Z
d¯3 k
:HLg : = − ωk ηAB a∗A (k)aB (k). (147)
2ωk
The quantum version of the gauge condition Eq. (146) implies that for phys-
ical states |φi and |φ0 i,

X Z d¯3 k
hφ0 |:HLg :|φi = ωk hφ0 |a∗A (k)aA (k)|φi. (148)
A=1,2 2ωk

In particular, the expectation value of the energy is clearly positive definite


for physical states.

2.9 Feynman propagator in Lorentz gauge


By now we should be accustomed to thinking of the Feynman propagator as
a Greens function for solving the field equations with a small imaginary mass
term,
(∂ 2 − i)Aµ (x) = j µ (x). (149)
As there are four fields, the Green’s function is a 4 × 4 object, and is defined
by
(∂ 2 − i)DF µ ν (x − x0 ) = −δνµ δ 4 (x − x0 ). (150)
Hence the Feynman propagator for a gauge field in the Lorentz gauge is just
four copies of a massless scalar propagator. The Fourier space representation
is
−ik·(x−x0 )
4 e
Z
0
DF µν (x − x ) = ηµν d¯ k 2 . (151)
k + i
Strictly speaking, one should also show that this is related to the two point
function, i.e. the vacuum expectation value of the time-ordered product of
two fields:

iDF µν (x − x0 ) = Gµν (x, x0 ) = h0|T [Aµ (x)Aν (x0 )]|0i. (152)

28
However, proving this relation will have to wait until we have studied the
path integral quantisation of the gauge field. Meanwhile, we will jump the
gun and write down the Feynman rules for the electromagnetic field
k,  r
Ingoing µ (k)
k, 
Outgoing r ∗µ (k)
µ νr Z
iηµν
Internal r d¯4 k 2
k + i

3 Quantum gauge theory


4 Electroweak theory
5 QCD
6 Renormalisation
A Pauli matrices
The Pauli matrices σ i (i = 1, 2, 3) are defined to be
! ! !
1 0 1 2 0 −i 3 1 0
σ = , σ = , σ = . (153)
1 0 i 0 0 −1

They are Hermitean, i.e.


(σ i )† = σ i , (154)
and they square to unity:

(σ 1 )2 = (σ 2 )2 = (σ 3 )2 = 1. (155)

One can verify that


σ 1 σ 2 = iσ 3 , (and cyclic). (156)
Lastly, any two different Pauli matrices anticommute, or

σ 1 σ 2 + σ 2 σ 1 = 0, (and cyclic). (157)

This set of relations can be neatly expressed using the 3 dimensional Levi-
Civita symbol εijk :
[σ i , σ j ] = 2iεijk σ k , (158)

29
and with the anticommutation relations
{σ i , σ j } = 2δ ij . (159)
One may thus write the product of any two Pauli matrices as
1 1
σ i σ j = {σ i , σ j } + [σ i , σ j ] = δ ij + iεijk σ k . (160)
2 2
The Pauli matrices form a representation of the angular momentum al-
gebra with (spin) angular momentum 12 .

B Dirac matrices
Dirac matrices (in four space-time dimensions) are 4×4 matrices defined by
the anitcommutation relations
{γ µ , γ ν } = 2η µν 14 , (161)
where η µν is the Minkowski metric, and 14 is the 4×4 identity matrix (which
is often dropped and left implicit in the equation).
Properties of the Dirac γ-matrices include:

(i). (γ µ )† = γ 0 γ µ γ 0 .

One can also define another, linearly independent, γ-matrix


γ 5 = iγ 0 γ 1 γ 2 γ 3 , (162)
which has the easily derivable properties
(γ 5 )2 = 1, (γ 5 )† = γ 0 γ 5 γ 0 , {γ µ , γ 5 } = 0. (163)
The eigenvector of this matrix are known as chirality eigenvectors. Eigenvec-
tors with eigenvalue +1 are termed right-handed, with -1 left-handed. One
can define projectors onto these eigenvectors:
1 1
PR = (1 + γ 5 ), PL = (1 − γ 5 ) (164)
2 2

B.1 Standard representation of Dirac matrices


! ! !
0 1 0 i 0 σi 5 0 1
γ = , γ = , γ = . (165)
0 −1 −σ i 0 1 0
Here, σ i are the Pauli matrices, which are defined in Appendix A, and 0 and
1 are the 2×2 zero and identity matrices respectively.

30
C Identities for Dirac matrices
(i). tr(γ µ γ ν ) = 4η µν

(ii). tr(γ µ1 . . . γ µ2n−1 ) = 0, n a positive integer.

(iii). tr(γ µ γ ν γ ρ γ σ ) = 4(η µν η ρσ − η µρ η νσ + η µσ η νρ )

(iv). tr(6 a1 . . . 6 a2n ) = 2n i


i=2 (−1) a1 · ai tr(6 a1 . . . 6 ai . . . 6 a2n ) (the hat denotes
P c
that 6 ai is removed from the trace)

(v). γ µ γµ = 41

(vi). γ µ 6 aγµ = −2 6 a

(vii). γ µ 6 a 6 bγµ = 4(a · b)1

(viii). γ µ 6 a 6 b 6 cγµ = −2 6 c 6 b 6 a

(ix). γ µ 6 a 6 b 6 c 6 dγµ = 2(6 d 6 a 6 b 6 c+ 6 c 6 b 6 a 6 d)

D Feynman rules
The rules which one derives from the LSZ reduction formula are for calcu-
lating S-matrix elements Sαβ = hα|Ŝ|βi. In momentum space one finds, for
a real scalar field, that the following rules are useful:
Z
i
Internal line (propagator) d¯4 k
k2 − m2 + i

k2 k3

k1 k4

Vertex −iλ δ̄ 4 (k1 + k2 + k3 + k4 )


Momentum is conserved for particle excitations in the vacuum, so when
one performs as many of the integrations using the vertex δ-functions as
one can, one always finds a factor of δ̄ 4 (K 0 − K), where K is the total 4-
momentum in the initial state, and K 0 the total 4-momentum in the final

31
state. Furthermore, there are also uninteresting contributions from graphs
with no vertices.
It is therefore convenient to define matrix elements Mαβ through

iMαβ δ̄ 4 (K 0 − K) = hα|(1 − Ŝ)|βi.

We can derived slightly simpler Feynman rules for these matrix elements,
in which the propagators lose their associated integrations and the vertices
their δ-functions.
Consider a diagram with E external lines, V vertices, and I internal lines.
For the simple φ4 theory we know that external lines end on a vertex, internal
lines have vertices at each end, and vertices must be attached to four lines.
Hence
E + 2I = 4V.
There are I integrations and V δ-functions, and we know that there is always
one δ-function left over. Hence we can trivially perform V − 1 integrations,
leaving
L=I −V +1
integrations to perform.
We call the number of free integrations the number of loops, as for simple
graphs it does indeed correspond to the number of closed loops in the graph.
The following sections give the Feynman rules for working out the matrix
elements Mαβ for several theories of interest.

32
E Real scalar field
1 1 1
Z  
4
S= dx ∂µ φ∂ µ φ − m2 φ2 − φ4
2 2 4!

k

Ingoing line 1

k

Outgoing line 1


k i
Internal line (propagator)
k2 − m2 + i

Vertex −iλ

F Complex scalar field


1
Z  
S= d x ∂µ φ̄∂ φ − m φ̄φ − λ(φ̄φ)2
4 µ 2
4

33
k

Ingoing particle line 1


k

Ingoing antiparticle line 1

k

Outgoing particle line 1


k

Outgoing antiparticle line 1


k i
Internal line (propagator)
k2 − m2 + i

Vertex −iλ

34
G U(1) gauge field
!
Z
1 1
S= d x − Fµν F µν − (∂ · A)2
4
4 2ξ

k, 

Ingoing line µ

k, 

Outgoing line ∗µ

µ ν

k ηµν + (ξ −
Internal gauge line (propagator) i
k2

H Spinor field with U(1) gauge symmetry


Z
S= d4 x ψ̄(i 6 D − m)ψ.

Here Dµ ψ = (∂µ + ieAµ )Ψ.

35
k, A

Ingoing particle line uA (p)


k, A

Ingoing antiparticle line v̄A (p)

k, A

Outgoing particle line ūA (p)


k, A

Outgoing antiparticle line vA (p)


p i
Internal line (propagator)
6 p − m + i

k1

k2

3-point vertex −ieγ µ

• Closed fermion loops are associated with a factor (−1).

36
I Non-Abelian gauge field
!
Z
4 1 a aµν 1
S= d x − Fµν F − (∂ · Aa )(∂ · Aa ) − c̄a ∂ · (Dc)a
4 2ξ

37
k, , a

Ingoing line

k, , a

Outgoing line

µ, a ν, b

k ηµν
Internal gauge line (propagator) iδab

a b

k
Internal ghost line (propagator)

k2 , ν, b

k1 , µ, a

k3 , ρ, c ig[(
3-point gauge vertex (

ν, b ρ, c

−ig 2 [feab
4-point gauge vertex +feac
µ, a σ, d +fead f

38
µ, a

k, c
• Closed ghost loops are associated with a factor (−1).

39
J Complex scalar field with non-abelian gauge
symmetry
Z  
S= d4 x (Dµ Φ)† (Dµ Φ) − m2 Φ† Φ − Vint (Φ, Φ† ) .

Here Φ is a column vector of Nf scalar fields φm , and Dµ Φ = (∂µ −igAaµ T a )Φ.


Equivalently, (Dµ φ)m = (δmn ∂µ −igAaµ Tmn
a
)φn . Rules are given for Vint (Φ, Φ† ) =
1
4
λ(Φ† Φ)2 .

40
k

Ingoing particle line 1


k

Ingoing antiparticle line 1

k

Outgoing particle line 1


k

Outgoing antiparticle line 1


k iδmn
Internal line (propagator)
k2 − m2 + i

k1 , m

µ, a

k2 , n

a
3-point gauge-scalar vertex igTmn (k1 + k2 )µ

ν, b n

41
K Spinor field with non-abelian gauge sym-
metry
Z
S= d4 x Ψ̄(i 6 D − m)Ψ.

Here Φ is a column vector of Nf spinor fields ψm , and Dµ Ψ = (∂µ −igAaµ T a )Ψ.


Equivalently, (Dµ ψ)m = (δmn ∂µ − igAaµ Tmna
)ψn .

42
k, A, m

Ingoing particle line umA (p)


k, A, m

Ingoing antiparticle line v̄mA (p)

k, A, m

Outgoing particle line ūmA (p)


k, A, m

Outgoing antiparticle line vmA (p)

m n

p iδmn
Internal line (propagator)
6 p − m + i

k1 , m

µ, a

k2 , n

a
3-point vertex igTmn γµ

• Closed fermion loops are associated with a factor (−1).

43
L Problem Sheets
L.1 Problem Sheet 1
1. (a) If we write
ψ(x) = ue−ip·x + veip·x ,
1
with p0 = E ≡ |(p2 + m2 ) 2 |, show that the the following expres-
sions for u and v give solutions to the Dirac equation (i 6 ∂ −m)ψ =
0:
! !
− 21 χ± − 12 0
u± (p) = (E+m) (6 p+m) v± (p) = −(E+m) (6 p−m)
0 χ±
! !
1 0
where χ+ = , and χ− = .
0 1
(b) Show also that
u†A uB = 2EδAB , vA† vB = 2EδAB
where the indices A and B range over the spin states + and −,
and that
ūA uB = 2mδAB , v̄A vB = −2mδAB ,
where ūA = u†A γ 0 , and v̄A = vA† γ 0 are adjoint spinors.
2. The Lagrangian density for a spinor field may be written
L = ψ̄(i 6 ∂ − m)ψ.

(a) Given that the general expression for the canonical energy-momentum
tensor is
∂L ∂L
θµ ν = ∂ν ψ + ∂µ ψ̄ − Lδνµ ,
∂(∂µ ψ) ∂(∂µ ψ̄)
evaluate the momentum density θi 0 and the energy density θ0 0 ,
showing that the latter is equal to the Hamiltonian density H.
(b) Given the plane wave expansion of the spinor field
X Z d¯3 p  
ψ(x) = cA (p)uA (p)e−ip·x + d†A (p)vA (p)eip·x ,
A=± 2E
show that
X Z d¯3 p  
H= E c†A (p)cA (p) − dA (p)d†A (p) .
A=± 2E

44
3. We define the matrix γ 5 = iγ 0 γ 1 γ 2 γ 3 .

(a) Show that γ 5 has the following properties:

(γ 5 )† = γ 5 , (γ 5 )2 = 1, {γ 5 , γ µ } = 0,

where 1 is the 4 × 4 identity matrix, and evaluate γ 5 in the stan-


dard representation. (You may find it useful to recall the relation
(γ µ )† = γ 0 γ µ γ 0 .)
(b) Eigenvectors of γ 5 are called eigenvectors of chirality, and eigen-
vectors with eigenvalue +1 are termed right-handed, with -1 left-
handed. Show that ψR = 21 (1 + γ 5 )ψ is right-handed, and that
ψL = 12 (1−γ 5 )ψ is left-handed. Show also that ψ̄ψ = ψ¯L ψR +ψ¯R ψL
and ψ̄γ 5 ψ = ψ¯L ψR − ψ¯R ψL . Note that ψ¯L = (ψL )† γ 0 .
(c) Consider the current density

j µ = ψ̄γ µ ψ.

Show that this current is conserved if ψ is a solution to the Dirac


equation, and show that ψ̄γ µ ψ = ψ¯R γ µ ψR + ψ¯L γ µ ψL
(d) Using γ 5 , we may also define another current carried by spinors,
called the axial-vector current:

j 5µ (x) = ψ̄(x)γ µ γ 5 ψ(x).

Assuming that ψ(x) satisfies the Dirac equation, calculate ∂µ j 5µ ,


and exhibit j 5µ (x) in terms of chirality eigenstates ψR and ψL .
Under what circumstances is the axial-vector current conserved?

45
4. (a) Consider a function of a pair complex Grassmann variables f (θ, θ̄).

f (θ, θ̄) = a + bθ + cθ̄ + dθθ̄,

and define the following operators:


∂ ∂
D= + iθ̄, D̄ = − + iθ.
∂θ ∂ θ̄
i. Show that {D, D̄} = 2i
ii. Show that if Df (θ, θ̄) = 0, then f (θ, θ̄) = a + cθ̄ − iaθθ̄.
(b) Now consider two pairs of complex Grassman variables, θa and θ̄b ,
with a and b taking the values 1, 2. Show that
Z Z
dθ2 dθ1 θa θb = ab , and dθ̄2 dθ̄1 θ̄a θ̄b = ab ,

where 12 = −21 = 1 and 00 = 11 = 0.


(c) Given the function of two pairs of complex Grassman variables

f (θ̄a , θb ) = exp(−θ̄a Mab θb ),

where M is a 2 × 2 matrix, show that


Z
dθ2 dθ̄2 dθ1 dθ̄1 f (θ̄a , θb ) = det M.

L.2 Problem Sheet 2


1. A complex scalar field theory has action
Z
S= d4 x (∂ µ φ̄)(∂µ φ) − V (|φ|).

The field may be expanded


Z
d¯3 k  −ik·x 
φ(x) = ak e + b∗k eik·x ,
2ωk
1
where ωk = (k2 + m2 ) 2 . In the quantum field theory, the ladder oper-
ators ak , a∗k , bk , and b∗k satisfy the relations [ak , a∗k0 ] = 2ωk δ̄ 3 (k − k0 ),
[bk , b∗k0 ] = 2ωk δ̄ 3 (k − k0 ), with all other pairs of ladder operators com-
muting.

46
(a) As shown in the lectures, there is a U(1) symmetry leading to a
conserved current j µ = iφ̄(∂ µ φ) − i(∂ µ φ̄)φ. Show that
Z
d¯3 k ∗
:Q: = (a ak − b∗k bk ).
2ωk k

where Q = d3 xj 0 (t, x) is the charge operator.


R

(b) Show that [:Q:, ak ] = −ak , [:Q:, a∗k ] = a∗k , [:Q:, bk ] = bk , and
[:Q:, b∗k ] = −b∗k ,
(c) Hence or otherwise show that the states a∗ (k)|0i and b∗ (k)|0i
(where |0i is the vacuum state) are eigenstates of the charge op-
erator, and give their charges.
(d) Show that [:Q:, φ(x)] = −φ(x), and prove that of Q|0i = 0 then
the vacuum expectation value of the field operator, h0|φ(x)|0i,
must vanish.

47
2. Consider a massless spinor with action
Z
S= d4 x ψ̄i 6 ∂ψ,

which is invariant under the two-parameter global symmetry transfor-


5
mation ψ → ψ 0 = e−iα−iβγ ψ.

(a) Suppose we wish to gauge these symmetries by allowing space-time


dependent α and β. Find the transformation laws for the vector
fields Vµ and Aµ which are required to construct the covariant
derivative
Dµ ψ = (∂µ − ieV Vµ − ieA Aµ γ 5 )ψ.
(b) Write down the gauge invariant action constructed from this co-
variant derivative. Show that it can be re-expressed in terms of
the chirality eigenspinors ψL and ψR as
Z  
Sgi = d4 x ψL i 6 D1 ψL + ψR i 6 D2 ψR ,

where Dµ1 = ∂µ − ie1 A1µ and Dµ2 = ∂µ − ie2 A2µ , giving the coupling
constants e1 and e2 in terms of eA and eB , and the vector fields
A1µ and A2µ in terms of Vµ and Aµ .

3. The action for a free electromagnetic field is


1
Z  
S= 4
d x − Fµν F µν .
4
(a) Show that

δFρσ (x) ∂ 4 ∂
= ρ
δ (x − y)δσµ − σ δ 4 (x − y)δρµ ,
δAµ (y) ∂x ∂x

and hence that δAδS


µ (x)
= ∂ν F νµ (x). Write down the field equation
for Aµ in the Lorentz gauge.
(b) Show also that the two actions
1 1 1
Z   Z  
S1 = d x − Fµν F µν − (∂ · A)2 ,
4
S2 = d x − ∂µ Aν ∂ µ Aν
4
4 2 2
both give the field equation you wrote down in part (a). Explain
briefly how it is that apparently different actions give the same
equation.

48
4. Consider an action for an SU(2) doublet spinor coupled to an SU(2)
gauge field
1 a aµν
Z  
4
S= d x − Fµν F + Ψ̄i 6 DΨ − mΨ̄Ψ .
4
where Dµ = ∂µ − igAaµ τ a , Fµν
a
= ∂µ Aaν − ∂ν Aaµ + gabc Abµ Acν , and τ a =
a a
σ /2, where σ are the Pauli matrices.

(a) Show that the resulting field equations are

∂ν F aνµ + gabc Abν F cνµ = −g Ψ̄γ µ τ a Ψ,


i 6 DΨ − mΨ = 0.

a
(b) Show that F̃µν = 12 µνρσ F aρσ satisfies the identity

∂ ν F̃µν
a
+ gabc Abν F̃µν
c
≡ 0.

You may find the identity abc cde = δad δbe − δae δbd useful.
(c) (Optional) Defining a 4-vector “current” Kµ = Aaν F̃µν
a
, show that

1
∂ µ Kµ = F aµν F̃µν
a
2
.

L.3 Problem Sheet 3


1. An action for a massive vector field is
1 1
Z  
S= d4 x − Fµν F µν + m2 Aµ Aµ .
4 2
(a) Show that the field equations resulting from extremising the action
is
[ηµν (∂ 2 + m2 ) − ∂µ ∂ν ]Aν = 0,
and show that ∂ · A = 0 follows. Contrast this equation for ∂ · A
with the status of ∂ · A = 0 in the theory for a massless vector
field. Justify the assertion that the massive vector field has three
polarisation states.
(b) By writing Aν (x) = aν e−ik·x , with aν a constant, show that the
momentum space version of the operator in square brackets in
part (a) is [ηµν (−k 2 + m2 ) + kµ kν ].

49
(c) The Feynman propagator for the massive vector field is defined in
the usual way by

[ηµν (∂ 2 + m2 ) − ∂µ ∂ν ]Dν ρ (x) = ηµρ δ 4 (x).

Give an argument which shows that the propagator must have the
form Z
Dµν (x) = δ̄ 4 k [A(k 2 )ηµν + B(k 2 )kµ kν ]e−ik·x

and show that A(k 2 ) = −(k 2 − m2 )−1 , B(k 2 ) = [m2 (k 2 − m2 )]−1 .


2. The action for a complex scalar field is
Z h i
S= d4 x (∂ µ φ̄)(∂µ φ) − V (|φ|) ,

with V (|φ|) = −µ2 φ̄φ + λ(φ̄φ)2 .

(a) By making the change of field variables φ = √12 (v + h(x) + ia(x)),


with v 2 = µ2 /λ, show that the quadratic (free) part of the action
is
1 1 2 1
Z  
4 µ µ
S0 = d x ∂µ h∂ h − m + ∂µ a∂ a ,
2 2 2
giving an expression for m in terms of the original parameters of
the action.
(b) Suppose we introduce a spinor field ψ, so that the action gains an
extra piece
Z  
Sψ = d4 x ψ̄i 6 ∂ψ − gφψL ψR − g φ̄ψR ψL .

Show that the spinor field has a mass gv, and interacts with the
fields h and a through the terms −ghψ̄ψ and −igaψ̄γ 5 ψ

3. The Lagrangian for a complex doublet scalar Φ with an SU(2) gauge


symmetry is
1
L = (Dµ Φ)† (Dµ Φ) − V (Φ) − tr(Fµν F µν ),
2
where Dµ = ∂µ − igAµ and −igFµν = [Dµ , Dν ].

(a) If we make the SU(2) tranformations Φ → Φ0 = U Φ and Aµ →


A0µ = U Aµ U −1 − gi ∂µ U U −1 , verify that the covariant derivative
transforms as
Dµ → Dµ0 = U Dµ U −1 .

50
(b) If the potential has the form V (Φ) = −µ2 Φ† Φ + λ(Φ† Φ)2 , show
that the Lagrangian has the U(1) global symmetry Φ → Φe−iβ .
Suppose that we wish to gauge this symmetry: how should we
amend the Lagragian? (You should display the transformation
properties of any extra fields you introduce).
(c) Suppose now that we have two SU(2) scalar doublets Φ1 and Φ2 ,
transforming under the combined SU(2)×U(1) gauge symmetry
as
Φ1 → Φ01 = U Φ1 e−iβ , Φ1 → Φ02 = U Φ2 eiβ .
Verify that the following potential is gauge invariant:

V (Φ1 , Φ2 ) =
−µ21 Φ†1 Φ1 − µ22 Φ†2 Φ2 + λ1 (Φ†1 Φ1 )2 + λ2 (Φ†2 Φ2 )2 + λ3 (Φ†1 Φ2 )(Φ†2 Φ1 ),

and show that the potential has an extremum at


! !
0√ 0√
Φ1 = , Φ2 = .
v1 / 2 v2 / 2

Does this theory have any other internal symmetries?



δf
4. (a) It follows from the identity 1 = Dφδ[f (φ) − ω] det δφ
R
that

−1
δf
δ[f (φ) − ω] = δ[φ − φ̃] det .

δφ

where φ̃ is the solution to the equation f (φ)−ω = 0 (here assumed


to be unique). Using the latter form of the identity, show that
Z R R
1
d4 x φ2 (x) d4 x (ln ω(x)+ 12 ln2 ω(x))
Dφδ[eφ − ω]e− 2 = e− .

(b) The measure for the functional integration over a complex scalar
field is Dφ̄Dφ = x dφ̄(x)dφ(x). Introducing “polar” fields ρ(x)
Q

and θ(x), where φ = ρiθ , changes the measure according to Dφ̄Dφ =


DρDθ| det M |, where
 
δφ(x) δ φ̄(x)
δρ(y) δρ(y)
M (x, y) =  δφ(x) δ φ̄(x)
.
δθ(y) δθ(y)

Show that R
d4 x ln ρ(x)
Dφ̄Dφ = DρDθe .

51
L.4 Problem Sheet 4
1. The axial gauge is defined by

Ga (A) = tµ Aaµ = 0,

where tµ is a constant spacelike 4-vector.

(a) If αa are the parameters of an infinitesimal gauge transformation


on Aaµ , find the operator

δGa (A(x))
Mab (x, y) = ,
δαb (y)
and hence write down the complete action for a non-Abelian gauge
field Aaµ in the axial gauge, include gauge-fixing and ghost terms.
(b) Show that the momentum space propagator in the axial gauge is

tµ k ν + k µ tν (ξk 2 − t2 ) µ ν
" #
µν δab
Dab (k) = 2 −η µν + + k k .
k + i k·t (t · k)2

2. (a) Show that

tr(γ µ γ ν ) = 4η µν , tr(γ µ γ ν γ ρ γ σ ) = 4(η µν η ρσ − η µρ η νσ + η µσ η νρ )

(b) Consider the following diagram in Quantum Electrodynamics (QED),


which is a theory of a spinor field of mass m and charge e, coupled
to an Abelian gauge field, in Feynman gauge.

µ ν

Show that applying the Feynman rules to this diagram results in


the expression
−iηµρ ρσ −iησν
iΠ (k) ,
k 2 + i k 2 + i

52
where
Z
[q µ (q + k)ν + (q + k)µ q ν − η µν (q · (q + k) − m2 )]
iΠµν (k) = −4e2 d¯4 q .
[(q + k)2 − m2 + i][q 2 − m2 + i]

NB Πµν (k) is called the polarization tensor.


(c) It can be shown that kµ Πµν (k) = kν Πµν (k) = 0. What does this
imply about the dependence of Πµν on k?

3. (a) Prove the following γ-matrix identities:

γ µ γµ = 41, γ µ γ ν γµ = −2γ ν .

(b) Consider the following diagram in QED:

Show that applying the Feynman rules (in Feynman gauge) to this
diagram results in the expression
i i
iΣ(p)
6 p − m + i 6 p − m + i
where
Z
4m − 2 6 q
iΣ(p) = −e2 d¯4 q .
[(p − q)2 + i](q 2 − m2 + i)

4. Consider the following correction to the propagator of a non-Abelian


gauge field:

53
µ, a ν, b

(a) Find the expression for the contribution of this diagram to the
polarisation tensor Πµν
ab resulting from the application of the Feyn-
man rules (in Feynman gauge) to this diagram.
(b) In a non-Abelian gauge theory, with no other fields, what other
diagrams contribute to Πµν
ab ?

L.5 Problem Sheet 5


1. Consider a theory of a real scalar field φ and an N -component complex
scalar field χa (a = 1, . . . , N ), whose action is
1 1 1
Z  
S= d4 x ∂φ · ∂φ + ∂ χ̄a · ∂χa − m2 φ2 − M 2 χ̄a χa − gφ2 χ̄a χa .
2 2 2
(a) Write down the generating functional for the n-point functions of
this theory, and show that in the absence of external currents one
can perform the functional integral over χa to obtain
Z R
d4 x ( 12 ∂φ·∂φ− 12 m2 φ2 −V1 (φ))
Z= Dφei ,

where V1 (φ) = −iN d¯4 k ln(k 2 − M 2 − 12 gφ2 + i).


R

(b) Define what is meant by a 1-particle irreducible (1PI) Feynman


diagram. Write down the Feynman diagram and its associated
integral for the the 1PI function for 2n ingoing φ fields with zero
4-momentum, Γ2n (0, . . . , 0), at one-loop order.
(c) Show that, up to a φ-independent (and possibly infinite) constant,

X 1 2n 2n
V1 (φ) = i φ Γ (0, . . . , 0).
1 (2n!)
Why can we ignore the constant?

54
(d) Define m2 (φ) = m2 + 21 gφ2 . Show that, using an upper cut-off
Λ  m(φ) on the momentum integration,

Λ2 2 1 1
   
4 2 2
V1 (φ) = N 2
m (φ) + N 2
m (φ) ln m (φ)/Λ − .
16π 16π 2

55
2. (a) Write down the covariant derivative terms for the fermion fields
in the Glashow-Weinberg-Salam (GWS) model of electroweak in-
teractions, in terms of their weak eigenstates, SU(2)×U(1) gauge
fields Aaµ and A0µ , and gauge couplings g and g 0 . Show that the
interaction terms can be written
Ls,i = g(Wµ J µ + Wµ† J †µ + Zµ JZµ ) + eAµ Jem
µ
,

where Wµ = (A1µ − iA2µ )/ 2, Zµ = cos θW A3µ − sin θW A0 µ, Aµ =
sin θW A3µ + cos θW A0µ , and θW is the weak mixing angle. Be sure
to give full expressions for the currents J µ , JZµ , and Jem
µ
.
(b) Show that JZµ and Jem µ
maintain their form in the mass eigenstate
basis, i.e. that (unlike the charged current) these currents do not
couple fermions with different flavour indices.
(c) Let Φ be the scalar doublet of the GWS model, with Φ† Φ = (v +
h(x))2 /2 in the symmetry-broken phase, where h is the Higgs field.
Show that the couplings of the Higgs to fermion mass eigenstates
are proportional to the masses of the fermions.
(d) Show also that there are Higgs-W interaction terms of the forms
hW W and hhW W , giving the coupling constants.
3. The bare interaction Lagrangian of QED is Li = e0 ψ̄0 γµ ψ0 Aµ0 .
(a) Show that if we demand that the action be dimensionless in d
dimensions, the mass dimension of the bare coupling constant e0
is /2, where  = 4 − d.
(b) Show that the photon self energy of Problem 4.2(b) becomes
iΠµν (k) =
Z 1 Z
(1 − d2 )η µν l2 − 2x(1 − x)pµ pν + η µν (m20 + x(1 − x)k 2 )
−4e20 dx d¯d l ,
0 (l2 + α)2
where α = m20 −x(1−x)k 2 . Note that in d dimensions, ηµν η µν = d.
(c) Perform the integrals to show that iΠµν (k) = (η µν k 2 −k µ k ν )iΠ(k 2 ),
where
e20 Z 1 2
 
2
Π(k ) = − dxx(1 − x) − ln α − γ .
2π 0 
(d) Given that the bare photon propagator can be written
−iηµν
iDµν (k) = ,
k 2 (1
− Π(k 2 ))

56
show how the divergence in Π(k 2 ) can be dealt with using the
photon wavefunction renormalisation constant Z3 .
(e) Introducing an arbitrary renormalisation parameter µ to define a
dimensionless renormalised coupling e = Z3 e0 µ−/2 , find the QED
β-function,
de
β=µ

in the limit  → 0. What is peculiar about the relation between
the renormalised and bare couplings?

Formulae for Problem 3:


Z 1
1 1
= dx
A 1 A2 0 [xA1 + (1 − x)A2 ]2
 n−d/2
Z
d 1 1 Γ(n − d/2) 1
d¯ l 2 n
= .
(l + α) (4π)d/2 Γ(n) α
1
zΓ(z) = Γ(1 + z), lim Γ(z) = −γ
z→0 z
where γ ' 0.5772.

57

Anda mungkin juga menyukai