You are on page 1of 77

Quantum Field Theory

Umut Gürsoy

ITP, Utrecht University - Fall 2019


Contents

1 Need for QFT 6

1.1 Quantum Mechanics . . . . . . . . . . . . . . . . . . . . . . . . . 7

1.2 Special Relativity . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

1.3 Free particle as an example . . . . . . . . . . . . . . . . . . . . . 9

1.3.1 Klein-Gordon equation . . . . . . . . . . . . . . . . . . . . 9

1.3.2 Dirac equation . . . . . . . . . . . . . . . . . . . . . . . . 10

1.4 Fundamental clash between QM and SR . . . . . . . . . . . . . . 12

2 Quantum Fields and canonical quantisation 16

2.0.1 Second quantisation . . . . . . . . . . . . . . . . . . . . . 16

2.1 Scalar field as a prototype . . . . . . . . . . . . . . . . . . . . . . 18

2.1.1 Lagrangian density, lagrangian and action . . . . . . . . . 19

2.1.2 Solution to the Klein-Gordon Equation . . . . . . . . . . 20

2.1.3 Classical Hamiltonian . . . . . . . . . . . . . . . . . . . . 21

2.2 Canonical Quantization . . . . . . . . . . . . . . . . . . . . . . . 22

2.2.1 Real scalar as an example . . . . . . . . . . . . . . . . . . 22

2.2.2 Quantum Hamiltonian . . . . . . . . . . . . . . . . . . . . 23

3 Path integral quantisation 25

3.1 Path integrals in quantum mechanics . . . . . . . . . . . . . . . . 25

3.1.1 Path integrals and time ordering . . . . . . . . . . . . . . 28

1
2 CONTENTS

3.1.2 Sources and n-point functions . . . . . . . . . . . . . . . . 29

3.1.3 Transition amplitude between arbitrary states . . . . . . . 30

3.2 Path integrals in QFT . . . . . . . . . . . . . . . . . . . . . . . . 32

3.2.1 Generating function: free case . . . . . . . . . . . . . . . . 33

3.2.2 Wick’s theorem . . . . . . . . . . . . . . . . . . . . . . . . 35

3.2.3 Effective potential . . . . . . . . . . . . . . . . . . . . . . 36

4 Interactions in QFT 39

4.1 Generating function: interacting case . . . . . . . . . . . . . . . . 39

4.2 Perturbative expansion and Feynman diagrams . . . . . . . . . . 40

4.2.1 Feynman rules . . . . . . . . . . . . . . . . . . . . . . . . 40

4.2.2 Symmetry factor of diagrams . . . . . . . . . . . . . . . . 41

4.2.3 Example: One point function . . . . . . . . . . . . . . . . 42

4.2.4 Connected vs. disconnected diagrams . . . . . . . . . . . 43

4.3 An introduction to renormalisation in QFT . . . . . . . . . . . . 45

4.3.1 Loop divergences . . . . . . . . . . . . . . . . . . . . . . . 45

4.3.2 Counterterms . . . . . . . . . . . . . . . . . . . . . . . . . 47

4.3.3 Renormalisation procedure . . . . . . . . . . . . . . . . . 48

4.3.4 Renormalisation conditions . . . . . . . . . . . . . . . . . 49

5 From theory to experiment: scattering in high energy physics 50

6 Representations of the Lorentz group and particles 51

6.0.1 Group representations . . . . . . . . . . . . . . . . . . . . 51

6.0.2 Lorentz generators and their algebra . . . . . . . . . . . . 53

6.0.3 Finite dimensional representations of the Lorentz group . 53

6.0.4 Representing the Poincare group in quantum mechanics . 56

6.0.5 Wigner’s construction of unitary representations . . . . . 58


CONTENTS 3

7 Spinors 61

7.1 Weyl and Dirac Lagrangians . . . . . . . . . . . . . . . . . . . . . 61

7.1.1 Dirac algebra and Lorentz transformations . . . . . . . . 63

7.1.2 Chirality . . . . . . . . . . . . . . . . . . . . . . . . . . . 64

7.2 Chirality and Helicity . . . . . . . . . . . . . . . . . . . . . . . . 65

7.2.1 Chirality . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

7.2.2 Helicity . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

7.3 Spin and Statistics . . . . . . . . . . . . . . . . . . . . . . . . . . 66

7.4 Canonical Quantization of Spinors . . . . . . . . . . . . . . . . . 66

7.5 The One-particle State . . . . . . . . . . . . . . . . . . . . . . . . 67

7.6 Quantization of Spinor Fields . . . . . . . . . . . . . . . . . . . . 68

8 Vector fields and the gauge symmetry 70

8.1 Proca Lagrangian: a = −1 . . . . . . . . . . . . . . . . . . . . . . 71

8.2 Vector Fields: Recap . . . . . . . . . . . . . . . . . . . . . . . . . 71

8.3 Quantum Vector Field . . . . . . . . . . . . . . . . . . . . . . . . 72

8.4 Gauge Invariance . . . . . . . . . . . . . . . . . . . . . . . . . . . 73

8.5 Quantum Gauge Field: Canonical Quantization . . . . . . . . . . 74

9 Quantum electrodynamics 76
Preliminaries

Lecturer: Umut Gürsoy, u.gursoy@uu.nl, BBG 7.16

Teaching assistants: Francisco Garcia Florez (f.garciaflorez@uu.nl), Chongchuo


Li (c.li@uu.nl), Natale Zinnato (natale.zinnato@gmail.com)

Grading: Bi-weekly exercises (exercises given in the odd numbered weeks will
be graded, starting from the first week): 20 %
Midterm exam on 8-11-2018: 30 %
Final exam on 31-1-2019: 50 %

Attendance: No attendance required to lectures but attendance is required to


the recitation classes.

Reading material: These notes are inspired by two QFT textbooks by

• Srednicki, M.

• Schwartz, M

and my own notes. There are other excellent material you can benefit from:
• Peskin, Schröder

• A. Zee

• Landau & Lifshitz

• Weinberg

• Brown, Ryder, Ramond

• D. Tong’s lectures

In general it is a good idea to use multiple sources and compare different ap-
proaches in studying an advanced subject like QFT. There are various different

4
CONTENTS 5

but equivalent ways to approach an idea in physics. For example you can study
QM using wavefunctions, operators, path integrals etc. and a concept such
as uncertainty in QM has different manifestations in different approaches, it is
essential to know all.

Content: I aim at a first introduction to QFT in these lectures and our plan
will be to cover the following topics:
• QFT basics (Clash b/w QM and SR)

• Scalar Field (Klein-Gordon Field) as a basic example


• Canonical Quantization
• Particles as unitary irreducible representations of the Lorentz algebra
• Spinors

• Vector fields
• Path integrals
• Correlation functions

• Scattering
• Gauge invariance and QED
Chapter 1

Need for QFT

It is a phenomenological fact that our world is described in terms of ele-


mentary particles at least at the highest energy scales available to us today.
This is for example what we confirm everyday in high energy experiments at
CERN. Two of the fundamental theories of the 20th century are supposed to
be applicable at these energy scales: quantum mechanics because we probe ex-
tremely short distance scales, and special relativity because these experiments
reach velocities close to the velocity of light.

The need to upgrade the quantum theory in terms of a relativistic field


as opposed to a Schrödinger wavefunction of a single particle stemmed from
a profound clash between quantum mechanics and special relativity. There
are various different manifestations of this clash. Let me first describe it in a
heuristic way1 . We know from QM that at extremely short time scales energy
can fluctuate wildly because of the Heisenberg’s uncertainty principle ∆E∆t ≥
~. Then, according to Einstein’s special relativity these fluctuations in energy
can be converted into creation of new particles essentially because energy and
mass are the same thing: E = mc2 . Therefore even if you start with a single
particle like an electron, when you probe it at very short distances you will end
up with seeing multiple electrons and photons. In other words, in the quantum
theory, the probability of an electron degenerating into another electron and
a photon, and then that photon degenerating into an electron and an anti-
electron, so on so forth is non-vanishing. If probability of a certain process is
non-vanishing in the quantum world and you choose to ignore this, you end up
in a terrible situation where the total probability of all processes at a given time
do not add up to 1, that is to say probability is not conserved. Mathematically
put, time evolution in quantum mechanics will be non-unitary. Loss of unitarity
is the end of the story, it is like non-conservation of energy. So we are forced
to include multi-particle states in our quantum theory in a manner consistent
with relativity.

As I said there are different manifestations of this clash. A more technical

1 You can read about this also in Landau & Lifshitz and A. Zee’s books.

6
1.1. QUANTUM MECHANICS 7

way to phrase it is to compare the way time and space are treated in QM.
Whereas the position of a particle correspond to the eigenvalues of the position
operator X̂i there is no such an operator that corresponds to time. Time enters
in QM merely as a label that labels the particular state of the system |ψ, ti in
the Hilbert space. This asymmetry in the way time and space are treated in QM
is in direct clash with special relativity where they form different components
of the same continuum, which is called the space-time. In other words Lorentz
transformations rotate space and time into one another.

1.1 Quantum Mechanics

To explain this conflict in mathematical terms we first state briefly what


we mean by quantum mechanics and special relativity. Axioms of quantum
mechanics are,

(i) State of the system at time t is represented by a vector in the Hilbert space
|ψ, ti ∈ H.

(ii) Observables are described by Hermitian operators Ô .

(iii) Measurement of an observable yields one of the (real) eigenvalues of the


operator Ô.

(iv) Time evolution of a state |ψ, ti is governed by the Schrödinger equation:

i~∂t |ψ, ti = H |ψ, ti (1.1)

In particular for a time-independent Hamiltonian the last equation is solved by


i
|ψ, ti = e− ~ Ht |ψ, 0i , (1.2)

showing that time evolution in QM is unitary. In other words hψ, t1 |ψ, t1 i =


hψ, t2 |ψ, t2 i for any t1 and t2 . Notice that this fact is a direct consequence of
having a single time derivative in the Schrödinger equation.

1.2 Special Relativity

Physics laws should be independent of the change in the inertial reference


¯). For simplicity, we define x0 = ct and xµ = (ct, ~x) where
frame: (ct, ~x) → (ct̄, ~x
µ = 0, 1, 2, 3 defines a space-time point in the Minkowski Spacetime. We have

x0 = −x0 , xi = xi .

More generally, we define the metric tensor to raise and lower indices:

xµ = ηµν xν , xµ = ηµν xν (1.3)


8 CHAPTER 1. NEED FOR QFT

where we employ the Einstein’s summation convention where repeated indices


are summed. From this equation you can infer that the metric with the upper
indices is inverse of the one with lower indices:
ηµν η νβ = δµβ → η µν = η −1 µν .

(1.4)

In Minkowski Spacetime, the metric tensor is given by:


 
−1 0 0 0
0 1 0 0
ηµν = 
0 0
 (1.5)
1 0
0 0 0 1
The inverse is defined by the equation:
xµ = η µν xν .
The inverse metric is defined as:
η µν ηνα = δαµ (1.6)
For an inertial frame, we have the transformation:
xµ → x̄µ
which is achieved through the Poincare transformation:
x̄µ = Λµν xν + aµ
where a is a constant translation in space-time and Λ is the Lorentz matrix.

Demanding that the space-time interval is invariant under Poincare trans-


formation
0 0
(x − x0 )2 = (xµ − x µ )ηµν (xν − x ν )
= −c2 (t − t0 )2 + |~x − x~0 |2
~
= (x̄ − x̄0 )2 = −c2 (t̄ − t¯0 )2 + |~x̄ − x¯0 |2
we arrive at the condition that defines the Lorentz transformation matrices.
Λµρ Λνσ ηµν = ηρσ (1.7)
A symmetric 4 by 4 matrix Λ can have 10 different entries. Satisfying the
condition above, it can only have 6 different parameters. These are
(i) 3 rotation angles θi
(ii) 3 boosts βi .

One can put equation (1.7) in the matrix form as follows:


ΛT · η · Λ = η , (1.8)
where T denotes transpose. This shows that the inverse Lorentz matrix is given
by
Λ−1 = η · ΛT · η . (1.9)
1.3. FREE PARTICLE AS AN EXAMPLE 9

1.3 Free particle as an example

To demonstrate the technical conflict between QM and SP let us consider


the hamiltonian for a free non-relativistic particle with mass m,
p̂2
Ĥ = , p̂ = −i~∇ .
2m
then the position space Schrödinger equation can be written as
~2 2
i~∂t ψ(~x, t) = − ∇ ψ(~x, t)
2m
where we define the position wavefunction as ψ(~x, t) = h~x|ψ, ti. This equation
is obviously not invariant under Lorentz transformations because there are two
space derivatives and only one time derivative. A Lorentz transformation would
mix space and time and necessitates treatment of time and space on equal
footing.

Well, you may say this is what you should expect because you considered a
non-relativistic Hamiltonian. Then, let us try a special relativistic Hamiltonian,
again for a free particle:
p p2
H= p2 c2 + m2 c4 ' mc2 + + ...
2m
which contains the correct physics in the non-relativistic limit. Then, we have
the equation in position space as:
p
i~∂t ψ(~x, t) = −~2 c2 ∇2 + m2 c4 ψ(~x, t) (1.10)
= mc ψ(~x, t) + O(∇2 ) ,
which can be Taylor expanded for a wavefunction varying over space slower than
the scale set by 1/m.

There are two problems with this equation:


(i) It is obviously still not Lorentz invariant as the number of time and space
derivatives are different.
(ii) This is a non-local equation of motion. This is because of the appearance
of infinitely many space derivatives obtained by expanding the square root. If
you are unfamiliar with what this means, think of a case you know better:
ei~a∇ ψ(~x, t) = ψ(~x + ~a, t)
where the infinite number of derivatives shifts the wavefunction in all space
non-locally.

1.3.1 Klein-Gordon equation

Now, let us square the Schrodinger equation to circumvent (ii) (at least):
−~2 ∂t2 ψ(~x, t) = (−~2 ∇2 + m2 c4 )ψ(~x, t) . (1.11)
10 CHAPTER 1. NEED FOR QFT

Let us denote a four-derivative as


 
∂ 1
∂µ = = ∂t , ∇ ,
∂xµ c

also define ∂ 2 = η µν ∂µ ∂ν . Then eq. (1.11) becomes:

m2 c4
 
2
−∂ + 2 ψ(~x, t) = 0 (1.12)
~

This is called the Klein-Gordon equation. We can easily see that it is Lorentz
invariant by changing the frame of reference by xµ → x̄µ = Λµν xν . The deriva-
tives ∂µ transform as ∂ µ → ∂¯µ = Λµν ∂ ν . Then using (1.7) ∂ 2 = ∂¯2 and the
Klein-Gordon equation remain the same if we identify

ψ̄(x̄) = ψ(x) . (1.13)

A field that transforms with this rule is called a scalar field. Note that this
transformation is trivial in the sense that the value of the field in any given
location remains the same. So equation (1.12) indeed seems to solve our problem
of finding a Lorentz invariant equation of motion.

It is not at all clear, however, whether the field ψ(x), when regarded as
a quantum mechanical wave-function ψ(x) = hx|ψi evolves unitarily in time.
Not clear because the Klein-Gordon equation, containing two time derivatives,
is certainly not of the Schrodinger form i~∂t |ψi = H |ψi which guarantees
unitary evolution for any hermitean H as the solution reads
i
|ψ(t0 )i = U (t0 − t) |ψ(t)i , U (t) ≡ e− ~ Ht . (1.14)

Thus, we still seem to find a clash between Lorentz invariance and axion (iv)
of QM reviewed above. Inspecting (1.11), you see that a promising solution
to this problem would be to somehow find a Hermitean differential operator
O, different than the Hamiltonian with the square-root in (1.10), that satisfies
O2 = −~2 ∇2 + m2 c4 so that a solution to i~∂t ψ = Oψ would also solve (1.10).
This would be both Lorentz invariant and unitary by construction. No such
O exists, as it should involve a ∇ not contracted with another vector which
results in LHS being a scalar but RHS being a vector. A more sophisticated
thing to do would be to write O = ai ∂i + · · · and impose the coefficients ai
satisfy ai aj = δij and so on. Of course no such collection of ordinary numbers
exist, but a collection of matrices does. This is what we discuss next.

1.3.2 Dirac equation

The problem of constructing such a first order in time equation was first
addressed successfully by Dirac for the spin 1/2 particles (spinors). Spinors
can be in two distinct quantum states ψ1 = |↑i and ψ2 = |↓i. This allows the
hamiltonian carry an additional spinor index Hab such that the Schrödinger
equation generalizes to i~∂t |ψa , ti = Hab |ψb , ti. Therefore H is now a matrix
and could solve the problem above. The most general Shrödinger equation first
1.3. FREE PARTICLE AS AN EXAMPLE 11

order in the derivatives and all vector indices contracted (to assure rotational
invariance) is then of the form (written in the position basis)

i~∂t ψa (~x, t) = −i~c(αj )ab ∂j + mc2 (β)ab ψb (~x, t) ,




where αi i = 1, 2, 3 and β are matrices with spinor indices. αi are much like
the Pauli matrices σ i but, as we see below, one should allow for a more general
possibility. The Hamiltonian written in a basis independent way is

Hab = cPj αj ab + mc2 (β)ab ,



(1.15)

and leads to unitary evolution by construction


 i 
i~∂t |ψa , ti = Hab |ψb , ti ⇒ |ψa , ti = e− ~ H |ψb , ti , (1.16)
ab

if we demand αi and β be Hermitean. To determine these matrices we require


square of this Hamiltonian to give the Klein-Gordon Hamiltonian, which we
know is Lorentz invariant2

(H 2 )ab = c2 Pj Pk (αj αk )ab + mc2 Pj (αj β + βαj )ab + m2 c4 (β 2 )ab


= (p2 c2 + m2 c4 )δab .

The solution is

{αj , αk }ab = 2δ jk δab (1.17a)


j
{α , β} = 0 (1.17b)
2
(β )ab = δab (1.17c)

where {A, B} = AB + BA. In Problem set I, exercise 2 you show that αj , β


should at least be 4 × 4. These are called Dirac γ−matrices.

Now the equation in the position basis can be written succinctly as,

i~γ µ ∂µ − mc2 ψ(~x, t) = 0 ,



(1.18)

where we suppress the spin indices a for simplicity. This is called the Dirac
equation. It describes propagation of a particle with two spin states, called a
spin-1/2 particle.

In exercise 2 you will also show that the gamma matrices are traceless.
Therefore Hab is a 4×4 matrix with the eigenvalues {E↑ (p), E↓ (p), −E↑ (p), −E↓ (p)}.
This means that along with the desired energy eigenstates for the up and down
spins we get two more negative energy eigenvalues. This leads to instability of
the system as follows. If you now couple the system to external medium, that
is to say, if the Dirac particles can interact with the surroundings e.g. with
electromagnetic waves, then they can lose energy indefinitely.

Dirac sea: To solve this issue Dirac supposed that all the negative eigen-
states are filled. Pauli exclusion principle allows this assumption because, as we
2 This is necessary but not sufficient for Lorentz invariance. We will see later when we

study the representations of Lorentz algebra, that indeed the equation before squaring it is
already Lorentz invariant.
12 CHAPTER 1. NEED FOR QFT

will see later, spin-1/2 particles are necessarily fermions and as Pauli showed
a state of a fermion can be occupied only once. Dirac argues that we do not
observe the Dirac sea, as we only observe the difference in energy in an experi-
ment. As the Dirac sea is always there, before and after experiment, we cannot
observe it. But, it has the following consequence.

If we send in a strong photon, then we can create a hole in the Dirac sea, as
some particles inside the sea will be excited. So, Dirac postulated that for every
fermion with charge e− there should be another fermion with charge −e− . For
every electron, there should be a positron. Dirac’s prediction of the positron in
1927 was confirmed experimentally in 1932 paving the way to his Nobel prize.

1.4 Fundamental clash between QM and SR

Now is time to recap. We started off with the observation that QM and
SR are incompatible with the single particle picture because many particles can
be created from a single one due to uncertainty in energy fluctuations in QM
and energy-mass equivalence in SR. What we managed to obtain above is an
example of a single particle propagation that is Lorentz invariant. But this
evolution is unitary only if you include the negative energy eigenstates! Indeed,
Dirac solved this issue by assuming an infinite sea of particles, falling back to
the same problem we started with: there is no way you can avoid the presence
of many particle states. In fact, as we will show below, the right solution to the
problem is to find a many particle representation of particle evolution in QM
and SR. That is to say the eigenstates should be combination of many particle
states. This will lead to the description in terms of quantum fields as we will
now show.

The easiest way to see why this clash is happening mathematically, and the
easiest explanation for the need for fields, is to recall the second axiom of QM:
Observables are represented by Hermitian operators: ~x, p~ etc. However, there
is no such operator for the time eigenvalue t which is only treated as a label of
the wave-function in QM. Thus, in QM, time (a label) and space (an operator)
is inherently separate and will necessarily transform differently under Lorentz
boosts making the entire formulation unsuitable.

There are two ways to solve this problem:

(i) Promote time to an operator


(ii) Demote ~x to a label

Let us start with the first and consider time in quantum mechanics also as
an eigenvalue of an operator. Now all space and time are treated democratically
and one can introduce a space-time basis: h~x, t|ψi = ψ(t, ~x). But then how do
we characterize the evolution in time, if time is a basis vector. The answer
is provided by special relativity: in terms of the proper time τ . Indeed the
history of a classical special relativistic particle is given by a curve xµ (τ ) in
1.4. FUNDAMENTAL CLASH BETWEEN QM AND SR 13

x y

Figure 1.1: The worldline of a particle parametrized by the proper time τ .


14 CHAPTER 1. NEED FOR QFT

x y
σ

Figure 2: “worldsheet” of an open string parametrized


by proper time τ and proper distance σ.

Figure 1.2: The worldsheet of an open string parametrized τ and σ.


1.4. FUNDAMENTAL CLASH BETWEEN QM AND SR 15

space-time, that is labelled by the proper time (see figure 1.4). This is called
the worldline of the particle. Thus the states in QM are labelled as |ψ, τ i and
the Schrodinger equation should be written with respect to τ . The interactions
in this description is given by splitting and joining of the worldlines. This
description is first introduced by Schwinger who used this idea to obtain the
first quantum correction to electron magnetic moment µ = 0.5 + .... This idea is
called the proper time formulation. It turns out to be hard to work with in the
presence of interactions. But this formulation had a great spin-off: by including
another variable σ to parametrize propagation one generalizes the worldline to
a worldsheet (see figure 1.4), that is, one arrives at the string theory!
Chapter 2

Quantum Fields and


canonical quantisation

In this course we will instead focus on the second method above, that is, we
will demote the position ~x from an opertator (or eigenvalue of an operator) to
a label of the quantum state of the system. This second method is literally the
quantum field theory where we demote the position ~x to a label. Hence both
position and the time become merely labels of the quantum state of the system:
|ψ, ~x, ti = |ψ(~x), ti (2.1)
It will be more useful for us to work in the Heisenberg picture where the opera-
tors become time dependent instead of the states. We then define the quantum
field operator in space-time as ψ̂(~x, t) and its time evolution is given by

ψ̂(~x, t) = eiHt/~ ψ̂(~x, 0)e−iHt/~ . (2.2)


I stress that the space dependence, either in (2.1) or in (2.2) is not the same as
the space dependence of the wave function a single particle in the position basis.
Here the quantum state of the system depends on the position irrespective of the
basis it is evaluated in. Equivalently the quantum operators themselves depends
on the position rather than their expectation value as in the single particle case:
h~x| Ô |~xi = Ô(~x). In other words, the Hilbert space of the system is extended
infinitely many times; it is now the infinite tensor product of the single particle
Hilbert space at every space point ~x. This point will become more clear in the
example below.

2.0.1 Second quantisation

In fact, this is not the first time you encounter the use of fields in quantum
mechanics. Non-relativistic quantum mechanics of N identical particles also
leads to a quantum field, and this formulation of many body quantum mechanics
is called the second quantisation.

16
17

The Schrödinger equation in the position basis for N particles with equal
mass, moving in potential U (~x) and with interparticle potential V (~x − ~y ) reads,
 
N  2
 X N N
X ~ X
i~∂t ψ =  − ∇2j + U (~xj ) + V (~xj − ~xk ) ψ
j=1
2m j=1 k=1,j6=k

where the wave function depends on particle positions and time ψ = ψ(~x1 , ..., ~xN , t).
We can write down an abstract Schrödinger equation, cf. eq. (1.1) by defining
creation and annihilation operators at every space point ~x. We introduce an
harmonic oscillator at every ~x: a(~x) annihilates a particle at position ~x, whereas
a† (~x) creates a particle at this position. These creation and annihilation oper-
ators are called quantum fields. Then the commutation relations between the
quantum fields follow from generalisation of the single harmonic oscillator to a
field of harmonic oscillators:
[a(~x), a(x~0 )] = [a† (~x), a† (x~0 )] = 0 (2.3a)
[a(~x), a† (x~0 )] = δ 3 (~x − x~0 ) (2.3b)
Using these creation and annihilation operators, the Hamiltonian can be written
as
~2 2
Z  
H = d3 xa† (~x) − ∇ + U (~x) a(~x) (2.4)
2m
Z
1
+ d3 x d3 yV (x − y)a† (x)a† (y)a(x)a(y) (2.5)
2
and the state as
Z
|ψ, ti = d3 x1 ... d3 xN ψ(x1 , ..., xN , t)a† (x1 )..a† (xN ) |0i . (2.6)

Note that the particle number N only enters in the state of system, eq. (2.6),
neither in the Hamiltonian (2.4) nor in the commutation relations (2.3b). This
means that this formalism is capable of describing any number of states — the
only thing you need to change is the number of creation operators (and the
number of integrals) in (2.6). Obviously the formalism can also be used when
the number N is not fixed but allowed to change in a process. We learn that the
use of quantum fields potentially solves the main problem in combining special
relativity with quantum mechanics, that I alluded to in the first page of chapter
2, namely the possibility to create many particles in sufficiently short increments
of time.

To understand the meaning of the expression (2.4) let us consider a simpler


system of a free non-relativistic scalar field in the second quantised picture:
∇2
Z  
H = d3 x a† (x) − a(x) .
2m
The Fourier transform of the quantum fields a(x) and a† (x) gives the annihila-
tion and creation operators in the momentum space:
d3 x −i~p·~x
Z
ã(~
p) = e a(~x) , (2.7)
(2π)3/2
18CHAPTER 2. QUANTUM FIELDS AND CANONICAL QUANTISATION

similarly for a† (x). Then the hamiltonian above can be written in the momen-
tum space as

p |2 †
|~
Z
H = d3 p ã (~
p)ã(~
p) . (2.8)
2m

The meaning is now clear. If we define the number operator N̂ (~ p) = ㆠ(~


p)ã(~
p)
that counts the number of particles carrying momentum p~, then, the hamiltonian
can be written as
Z
H = d3 pE(|~ p|)N̂ (~
p) ,

which is nothing but the operator that measures the total energy in the sys-
tem. To p
obtain a relativistic Hamiltonian, we can take the energy as relativistic
p| = |~
E(|~ p|2 c2 + m2 c4 ).

In particular, the state with no particles is called the vacuum state and
denoted by |0i. It is defined as

ã(p) |0i = 0, ∀~
p

Other states in the Hilbert space are obtained from the vacuum by acting with
creation operators:

ㆠ(~
p) |0i : one particle state
† †
ã (~ p2 ) |0i :
p1 )ã (~ two particle state ,

etc.

In passing we note that one can infer the statistics of the particle, whether
it is a fermion or a boson from the commutation relations. Commutators
a† (x)a† (x0 ) = a† (x0 )a† (x) result in Bose-Einstein statistics as they require
a symmetric wave-funtion in (2.6) whereas anti-commutators a† (x)a† (x0 ) =
−a† (x0 )a† (x)) lead to Fermi statistics.

Also, from now on I will be using the natural units in relativistic quantum
fields by setting ~ = c = 1.

2.1 Scalar field as a prototype

The simplest example of a relativistic quantum field is the scalar field we


discussed in section 1.3.1. We will derive most of the basic results of QFT using
this simple example and generalise to spin-1/2 and spin-1 fields later on. Also,
from here on I will be denoting the scalar field by φ(x). As discussed above it
transforms trivially under Lorentz transformations: φ̄(x̄) = φ(x) where x̄ = Λx
is the transformed frame. In the absence of interactions the relativistic scalar
field obeys the Klein-Gordon equation (1.12) which we reproduce here:

−∂ 2 + m2 φ(x) = 0 .

(2.9)
2.1. SCALAR FIELD AS A PROTOTYPE 19

So far there is nothing quantum about this equation. It just describes propa-
gation of a classical scalar field. One strategy to obtain a quantum relativistic
scalar field is to promote this classical field to a quantum field φ(x) → φ̂(x).
But how do we do this?

Actually, what we need is precisely the relativistic version of the quantum


fields a(x), a† (x) discussed in the previous section. Obviously, we should first
ask the question, what is the Hamiltonian associated with this field φ. To obtain
it we will first determine the lagrangian that leads to equation (2.9) and then
make a Legendre transformation to arrive at the hamiltonian.

2.1.1 Lagrangian density, lagrangian and action

In classical mechanics the lagrangian is the the function whose minimisation


leads to the equation of motion, and the action is defined as the time integral
of the lagrangian. Here we are dealing with a field φ(x) defined at every space
(and time) point, therefore the relevant quantity to minimise is the lagrangian
density L. The total lagrangian is given by the integral over L over the entire
space Z
L(t) = d3 x L(~x, t) , (2.10)

and the action is given by the time integral of the lagrangian


Z Z
S = dt L = d4 x L(~x, t) . (2.11)

As in classical mechanics the lagrangian depends on the field and its derivatives1 .
In addition, we need a theory that preserves Lorentz transformations, thus the
dependence on the derivatives should be in the relativistic form ∂µ φ(x).

More generally, the action should be invariant under Lorentz transforma-


tions if it leads to Lorentz invariant equations of motion. This means that
the lagrangian density should transform as a scalar under the Lorentz trans-
formations. You can easily see this: the space-time volume element d4 x is
itself invariant because the jacobian associated with Lorentz transformations is
|detΛ| = 1, which follows from (1.8) and (1.5). Then, for S to be invariant
under x → x̄ = Λ · x we need L̄(x̄) = L(x). This is the definition of a scalar.

We can now use the facts i) it is a scalar, and ii) its variation should lead
to (2.9) to write down the most general langangian density as follows:
1 1
LKG (x) = − ∂µ φ∂ µ φ − m2 φ2 + Ω0 , (2.12)
2 2
where Ω0 is a constant. The EOM can be obtained by the Euler-Lagrange
equations. The action is:
Z  
1 1
SKG = d4 x − ∂µ φ∂ µ φ − m2 φ2 + Ω0
2 2
1 There can be no explicit space and time dependence in a theory that is invariant under

space and time translations, that is, a theory that conserves energy and momentum.
20CHAPTER 2. QUANTUM FIELDS AND CANONICAL QUANTISATION

Changing φ(x) → φ(x) + δφ(x), the change in action is:


Z
δSKG = d4 x −∂µ (δφ)∂ µ φ − m2 φ(δφ)


Z Z
= − d4 x∂µ (δφ ∂ µ φ) + d4 x (∂µ ∂ µ − m2 )φ (δφ) = 0 .


The first total derivative term gives no contribution as we require the fields
and its derivatives go to zero sufficiently facts as |x| → ∞. The second term
then gives us the Klein-Gordon equation, since it should vanish for an arbitrary
variation δφ(x).

2.1.2 Solution to the Klein-Gordon Equation

Most general solution to the Klein-Gordon equation is of the superposition


of plane waves of the form
d3 k  ~ i~k·~x−iwt
Z 
~
φ(x) = a(k)e + b(~k)eik·~x+iwt
f (|~k|)
where a and b are, yet arbitrary functions that determine the magnitude of
each individual plane wave and f is a function that we will determine shortly
by demanding Lorentz invariance of the integral measure. This solves equation
(2.9) for the choice q
w(~k) = |~k|2 + m2 . (2.13)
This is called the “mass shell condition” or the “dispersion relation” depending
on the context. Note that there are who linearly independent solutions with ±w
in the exponent. The one that goes with −w would correspond to a negative
energy particle if we quantise this field by replacing the coefficients a and b by
creation operators. As we see below, the solution of QFT of this problem is to
replace a with an annihilation operator instead.

A specific case is a real scalar field which means that φ(x) ∈ R for all x. This
is what we will assume in the rest of this chapter. The more general, complex
scalar field case will be discussed in the exercises. Demanding φ(x) real, we find
hat b∗ (−~k) = a(~k) and f (~k) ∈ R. This yields

d3 k  ~ i~k·~x−iwt
Z 
~
φ(x) = a(k)e + a∗ (~k)e−ik·~x+iwt , (2.14)
f (|~k|)

after a simple change of variables. As mentioned above we determine f (|~k|)


by requiring that the integral measure in (2.14) is Lorentz invariant. chosen
properly such that φ(x) is a Lorentz scalar. We will determine the correct
measure by starting from a measure that we know to be Lorentz invariant:
kµ →k̄µ
d4 k = dk 0 d3 k −−−−−→ dk̄ 0 d3 k̄ = |detΛ| dk 0 d3 k = dk 0 d3 k,

because |detΛ| = 1 by the definition of Lorentz transformations (ΛηΛT = η =⇒


|detΛ|2 = 1). Here we defined k 0 = w for a compact notation. We should
2.1. SCALAR FIELD AS A PROTOTYPE 21

then also include a Heavyside theta function Θ(k 0 ) in the measure because
w is positive definite by definition (2.13). This theta function is invariant by
itself under the orthos=chronous subgroup of Lorentz transformations, see the
exercise. All in all we can change the integral into:
Z Z Z 3~
3 2 2 d k
d k dk0 δ(k + m )Θ(k0 ) = ,
2w
where k 2 + m2 = −k02 + |~k|2 + m2 = w2 − k02 . This means that the measure in
(2.14) will be invariant if we choose f to be proportional to w. Our convention
is to chose

˜ = d3 k
dk , (2.15)
(2π)3 2w
as the Lorentz invariant integral measure. The resulting scalar field becomes:
Z  
φ(x) = dk ˜ a(~k)ei~k·~x−iwt + a∗ (~k)e−i~k·~x+iwt
Z  
= dk ˜ a(~k)eik·x + a∗ (~k)e−ik·x . (2.16)

where k · x = k µ xµ = k 0 x0 + ~k · ~x = −wt + ~k · ~x.

2.1.3 Classical Hamiltonian

Having found the lagrangian, we now obtain the hamiltonian for the classical
scalar field via the Legendre transformation. Recall that in classical mechanics
the Legendre transform from the Lagrangian L(qi , q̇i ) of a collection
P of degree
of freedom labelled by i to the hamiltonian is H(qi , pi ) = i pi q̇i − L(qi , q̇i ).
The generalised momenta are defined as pi = δL/δ q̇i . In classical field theory,
these notions are immediately generalised to fields2 :
qi → φ(x) : Canonical Position
δL
pi → Π(x) = : Canonical Momentum
δ φ̇(x)
H(qi , pi ) → H[φ(x), Π(x)] = Π(x)φ̇(x) − L(x) : Hamiltonian density
Using (2.12) we find the canonical momentum for the Klein-Gordon field as
Π(x) = φ̇(x). Then the hamiltonian, that is the integral of the hamiltonian
density over space, is
Z
1 2 
H = d3 x φ̇ (x) + |∇φ(x)|2 + m2 φ2 (x) − 2Ω0 . (2.17)
2
We transform this Hamiltonian into the Fourier space as,
Z  Z
1 
˜ w a∗ (~k)a(~k) + a(~k)a∗ (~k) − d3 xΩ0
H= dk (2.18)
2
In the next section we use canonical methods to quantise this system.
2 You should think of discretising space by dividing it into little boxes and labelling these

boxes by i.
22CHAPTER 2. QUANTUM FIELDS AND CANONICAL QUANTISATION

2.2 Canonical Quantization

The basic idea of canonical quantisation of a field is the same as in quantum


mechanics. In quantum mechanics of a N particles, to quantise canonically
we start with the Hamiltonian formulation of the classical system where the
system is described in terms of the canonical degrees of freedom are the canonical
positions qi and canonical momenta pi . Then we promote these position and
momenta to operators q̂i and p̂i and promote their Poisson brackets

{q, qj } = 0 = {pi , pj }; {qi , pj } = δij , (2.19)

to commutation relations3 :

[q̂i , q̂j ] = 0 = [p̂i , p̂j ]; [q̂i , p̂j ] = i~ δij . (2.20)

Note that they are equal time commutators. That is, once they are defined at
a given time, say t = 0, their form do not change in any later time t. This is
obvious from the time evolution q̂(t) = exp(−iHt)q̂(0) exp(−iHt), etc.

Generalisation to canonical quantisation of fields is completely straightfor-


ward. Again, one first promotes the canonical degrees of freedom to quantum
operators and the Poisson brackets to (anti-)commutators.

2.2.1 Real scalar as an example

Easiest example is the real scalar field we discussed above. The canonical
position is given by the field φ(x) itself, see eq. (2.16). To promote this to
a quantum operator we need to promote the coefficients to operators acting
on a Hilbert space: a(~k) → â(~k) and a∗ (k) → a† (~k). These are precisely the
creation/annihilation operators defined in section 2.0.1, they destroy and create
a particle of momentum ~k out of the vacuum state. So the generalised position
is the quantum scalar field:
Z  
φ̂(x) = dk ˜ â(~k)eik·x + ↠(~k)e−ik·x . (2.21)

Then one imposes the equal-time canonical commutation relations that gener-
alise (2.20) to fields4 :

[φ(~x, t), φ(~x0 , t)] = 0 = [π(~x, t), π(~x0 , t)]


[φ(~x, t), π(~x0 , t)] = iδ 3 (~x − ~x0 ) . (2.22)

It is a straightforward exercise to substitute (2.21) in the commutators (2.22)


to obtain the commutators between the creation/annihilation operators:

[a(~k), a(~k 0 )] = 0 = [a† (~k), a† (~k 0 )] = 0


[a(~k), a† (~k 0 )] = (2π)3 2wδ 3 (~k − ~k 0 ) . (2.23)
3 orto anti-commutation relations if these degrees of freedom are fermionic.
4 From now on I will drop the hats on the fields as it will be clear from the context whether
they are classical or quantum fields.
2.2. CANONICAL QUANTIZATION 23

Note that if we had a spin 1/2 system, commutators would become anticommu-
tators when we apply the canonical quantization {a(~k), a† (~k 0 )} = i2w(2π)3 δ 3 (~k−
~k 0 ).

2.2.2 Quantum Hamiltonian

The Hamiltonian of the quantum field is now obtained by promoting the


coefficients a(~k) → â(~k) and a∗ (k) → a† (~k) in the classical Hamiltonian (2.18)
and using the commutation relations (2.23) as
Z
H = dk ˜ ω(~k) a† (~k)a(~k) + (0 − Ω0 )V (2.24)

where V is the volume of the space and


1 ∞ d3 k
Z q
0 = |~k|2 + m2 , (2.25)
2 −∞ (2π)3

is ~
q the energy density of the vacuum. Recalling the dispersion relation ω(k) =
|~k|2 + m2 , we see that 0 is just the sum over the ground states energy of
harmonic oscillators. The system describes a collection of harmonic oscillators
of every ~k at every space point. Translation invariance then gives rise to the
volume factor in (2.24).

Note that the additive constant in (2.24) depends on the order as and a∗ s we
choose in (2.18) before quantisation. This is called the ordering ambiguity and
it shows that the passage from classical to quantum mechanics is not unique.
However, this ambiguity does not show in energy differences which is what we
typically measure e.g. the energy of an excited state with respect to the vacuum
energy5 To fix the ordering ambiguity we will use the extra constant Ω0 and set
the ground state energy of the system to zero by choosing Ω0 = 0 . Then the
Hamiltonian becomes:
Z
H = dk ˜ ω(~k) a† (~k)a(~k). (2.26)

The creation/annihilation operators in this Hamiltonian satisfy the commuta-


tion relations (2.23). Defining the more conventional operators by
√ √
a(~k) = 2ω(2π)3/2 ã(~k) , a† (~k) = 2ω(2π)3/2 ㆠ(~k) ⇒ [ã(~k), ㆠ(~k 0 )] = δ 3 (~k−~k 0 ) ,
(2.27)
and recalling the definition dk ˜ = d3 k/2ω(2π)3 we discover that the Hamiltonian
of the Klein-Gordon field is
Z q
H = d k |~k|2 + m2 ㆠ(~k)ã(~k).
3
(2.28)

that is precisely the total energy of a free relativistic system. Compare this
with the non-relativistic counterpart (2.8). This is the answer we sought in the
5 It is important however then quantum fields are coupled to gravity, as the gravitational

force does depend on the energy of the fields they couple to.
24CHAPTER 2. QUANTUM FIELDS AND CANONICAL QUANTISATION

beginning of this chapter: the relativistic Klein-Gordon field upon quantisation


indeed describes a collection of relativistic free bosons. Furthermore, not only
the theory is Lorentz invariant but also it describes unitary evolution given by
the Hamiltonian (2.28). The issues associated with a relativistic Schrodinger
equation of a single particle that we discussed below equation (1.10) are all
resolved by enlarging the Hilbert space from a single particle to a field.
Chapter 3

Path integral quantisation

So far we have established two main points:

1. The need to pass from a single particles to fields to retain both Lorentz
invariance and unitarity.

2. The method of canonical quantisation of a field

In this chapter we will add another basic building block of QFT: description in
terms of path integrals. This is completely equivalent to the canonical methods,
yet has great advantages over the former which will become clear when we
discuss interactions in the next chapter. Below, we will first introduce the
concept of path integrals in quantum mechanics, then generalise to fields.

3.1 Path integrals in quantum mechanics

Consider a quantum mechanical hamiltonian of a non-relativistic single par-


ticle with mass m in a potential well V in one dimension (for simplicity) with
position q and momentum p. A generic Hamiltonian reads,

P2
H(P, Q) = + V (Q) , (3.1)
2m
where P and Q are the momentum and position operators respectively: P |pi =
p |pi, Q |qi = q |qi. A basic object we want to know is the probability amplitude
for this particle to travel a.k.a the propagator from point qi at time ti to qf at
time tf :
hqf , tf |qi , ti i = hqf | e−iH(tf −ti ) |qi i . (3.2)
Now divide the time interval T = tf − ti into N + 1 equal length pieces δt =
T /(N +1). In the end we will take the limit δt → 0, N → ∞ with T = (N +1)δt

25
26 CHAPTER 3. PATH INTEGRAL QUANTISATION

= constant. Then we can write


Z ∞ N
Y
hqf , tf |qi , ti i = dqj hqf | e−iHδt |qN i hqN | e−iHδt |qN −1 i · · · hq1 | e−iHδt |qi i ,
−∞ j=1
(3.3)
by inserting complete set of momentum eigenstates at N points. Take one of
these pieces now, hq2 | e−iHδt |q1 i and apply the Campell-Baker-HAusdorf for-
mula
1
eA+B = eA eB e− 2 [A,B]+··· .
We have
δt 2 2
e−iHδt = e−i 2m P e−iδtV (Q) eO(δt ) .
We can ignore the O(δt2 ) terms in the δt → 0 limit. Then we can evaluate this
piece by inserting a complete set of momentum eigenstates in between the two
exponentials:
Z ∞
δt 2
−iHδt
hq2 | e |q1 i = dp1 hq2 | e−i 2m p |p1 i hp1 | e−iδtV (Q) |q1 i
−∞
Z ∞
dp1 −i δt p2 −iδtV (Q) ip1 (q2 −q1 )
= e 2m e e
−∞ 2π
Z ∞
dp1 −iH(p1 ,q1 )δt ip1 (q2 −q1 )
= e e
−∞ 2π

where we used hq|pi = exp(ipq)/ 2π. Gathering all, we have,
Z ∞ N N
Y Y dpj ipj (qj+1 −qj ) −iH(pj ,qj )δt
hqf , tf |qi , ti i = dqj e e (3.4)
−∞ j=1 j=0

where we defined q0 = qi and qN +1 = qf . We now take the δt → 0 limit. First,


we can use the definition of the derivative
qj+1 − qj
q̇j = lim ,
δt→0 δt
in this limit. Similarly, the product of exponentials can be written as the expo-
nent of an integral
QN over t. The last crucial step is to realise that the product of
the q integrals j=1 dqj altogether can be understood as a sum over all possible
paths between points qi and qf , see figure 3.1. The same is true for the product
of p integrals. We call this limit as a path integral and denote it by a curly D
symbol:
YN Z ∞
Dq(t) ∝ lim dqj , (3.5)
δt→0,N →∞ −∞
j=1

where the proportionality factor will be fixed in the exercise. It can be rigorously
proven that this limit exists and results in a smooth integral over paths between
two points in the case of a single variable e.g. q(t). In the next section we will
generalise this to a path integral over fields, which can also be proven to exist
for many QFTs (certainly for the theories we consider in these lectures). For
more complicated QFTs e.g. the theory of strong interactions, QCD, nobody
3.1. PATH INTEGRALS IN QUANTUM MECHANICS 27

doubts that the path integral description exists but it is still not rigorously
proven mathematically.

Our final answer for the propagator is then,


Z q(tf )=qf Z R tf
hqf , tf |qi , ti i = Dq(t) Dp(t)ei ti dt{p(t)q̇(t)−H((p(t),q(t))} . (3.6)
q(ti )=qi

Some comments:

• The boundary conditions on the q integral means that we are summing


over all paths in space that starts at qi at time ti and end at qf at time
tf , see figure 3.1.
• There is no boundary condition on the p integral, which means that we
should also be integrating over the initial and final momenta. This, of
course, could not be otherwise, since the Heisenberg uncertainty principle
implies that the initial and final momenta are unfixed if the position are.
• Notice that all reference to operators, Hilbert space etc. all those notions
that we are familiar form canonical quantisation disappeared: there are no
operators inside the path integral (3.6). All terms are ordinary functions.

Equation (3.6) is our first path integral formula. We will now derive a simpler
expression by performing the p-path integral in the particular case when H is
at most quadratic in p. To do this we go back to the discrete expression (3.36)
and we perform each of the pj integrals using Gaussian integration. Denote the
exponent by F (p) = f2 p2 + f1 p + f0 and rewrite it as F (p) = f2 (p − p∗ )2 + F (p∗ )
where p∗ is the saddle point F 0 (p∗ ) = 0. Then the result of each p integral is
Z ∞
dp iF (p) 1
e =p eiF (p∗ ) .
−∞ 2π −2iπF 00 (p∗ )
We will be interested in the special case where the coefficient f2 = F 00 (p∗ ) is
independent of q just like in the case of non-relativistic QM where f2 = 1/2m.
In this case the multiplicative factor is just a constant and it will contribute
to the definition of the q path integral. We are therefore only interested in the
exponential, which is given by p∗ q̇ − H(p∗ , q) in the δt → 0 limit, where p∗ is
determined from the equation

∂ ∂
(pq̇ − H(p, q)) = 0 ⇒ q̇ = H(p∗ , q) .
∂p p=p∗ ∂p
This is precisely the Legendre transformation of the Hamiltonian, which means
that the exponential is precisely the Lagrangian of the theory L(qj , q̇j ). We
therefore arrive at our second expression for the path integral representation
which involves a single path integral over q:
Z q(tf )=qf R tf
hqf , tf |qi , ti i = Dq(t) ei ti dt L(q,q̇) . (3.7)
q(ti )=qi

This tells us that the probability amplitude of a particle at point qi at time ti to


arrive to point qf at time tf is given my summing over all paths between these
two points weighed exponential of the action.
28 CHAPTER 3. PATH INTEGRAL QUANTISATION

Figure 3.1: Definition of the path integral for a quantum mechanical particle
between initial position q 0 and final position q 00 and initial and final times t0 and
t00 .

3.1.1 Path integrals and time ordering

Path integral formulation is very useful to calculate scattering amplitudes


because it is automatically time ordered. Consider for example the expectation
value of the position operator between two position eigenstates at different times
ti and tf , hqf , tf | Q(t1 ) |qi , tt i where ti < t1 < tf . Here Q is the position operator
in the Heisenberg formalism

Q(ti ) = eiHt1 Qe−iHt1 .

The state |q, ti is defined as the instantaneous position eigenstate defined as


follows. Given an arbitrary state |ψ, ti, its position basis value is given by

ψ(q, t) = hq, t|ψi = hq|ψ, ti = hq| e−iHt |ψi ⇒ hq, t| = hq| e−iHt

hence
|q, ti = e+iHt |qi . (3.8)

Then we have

hqf , tf | Q(t1 ) |qi , tt i = hqf | e−iH(tf −t1 ) Qe−iH(t1 −ti ) |qi i .

This is now precisely the same form from which we derived the path integral
description. We can use the same trick to divide the intervals t1 − ti and tf − t1
into infinitesimal steps, inserting position and momentum eigenstates etc. This
leads to
Z q(tf )=qf R tf
hqf , tf | Q(t1 ) |qi , ti i = Dq(t) q(t1 ) ei ti dt L(q,q̇)
. (3.9)
q(ti )=qi
3.1. PATH INTEGRALS IN QUANTUM MECHANICS 29

Note that the operator is replaced by its eigenvalue in the path integral. Obvi-
ously this can be immediately generalised to an arbitrary operator
Z q(tf )=qf R tf
hqf , tf | F [Q(t1 ), P (tt )] |qi , ti i = Dq(t) F [q(t1 ), p(t1 )] ei ti dt L(q,q̇) ,
q(ti )=qi
(3.10)
following the same steps that lead to expression (3.6).

Now consider the situation where two operators inserted at different times
hqf , tf | Q(t1 )Q(t2 ) |qi , tt i. Since the path integral formalism is automatically
ordered in time, this expression can be represented in the path integral formalism
only for t1 > t2 . Introducing the time ordering symbol

A(t1 )B(t2 ) t1 > t2
T A(t1 )B(t2 ) = , (3.11)
B(t2 )A(t1 ) t2 > t1

we find that the path integral automatically yields the time-ordered two-point
or higher point functions:
Z q(tf )=qf R tf
hqf , tf | T Q(t1 )Q(t2 ) |qi , ti i = Dq(t) q(t1 )q(t2 ) ei ti dt L(q,q̇) . (3.12)
q(ti )=qi

This has the obvious generalisation to the product of n arbitrary operators:


Z qf
hqf , tf | T A1 [Q, P, t1 ] · · · An [Q, P, t2 ] |qi , tt i = Dq(t) A1 [q, p, t1 ] · · · An [q, p, tn ] eiS ,
qi
(3.13)
where we use the notation A[Q(t), P (t)] = A[P, Q, t]

3.1.2 Sources and n-point functions

Time ordered correlation functions measure how two or more points in time
are correlated with each other. For example how a measurement at time t1
affects another measurement at t2 etc. They constitute the basic observables
both in quantum mechanics and in quantum field theory. As we have discussed
above the 2-point correlators describe the probability amplitude of propagation
from one point in time to another. This idea can easily be generalised to n-
point correlators. As we discuss later in the course, the basic observables in
the high energy experiments, the scattering amplitudes are determined by these
correlation functions.

The n-point functions can be determined using the generating function. To


do this, we first define the functional derivative:
δ
j(t2 ) = δ(t1 − t2 ) . (3.14)
δj(t1 )
This is just a generalisation of the ordinary derivation δji /δjj = δij to continu-
ous variables. Now we modify the Hamiltonian by adding the source terms:

H(p, q) → H(p, q) − j(t)q(t) − h(t)p(t) ,


30 CHAPTER 3. PATH INTEGRAL QUANTISATION

and we generalise the propagator in the presence of sources as,


Z qf Z R tf
hqf , tf |qi , ti ij,h = Dq(t) Dp(t)ei ti dt{p(t)q̇(t)−(H((p(t),q(t))−j(t)q(t)−h(t)p(t)} .
qi
(3.15)
Now a general n-point correlator of a series of Q and P operators can be obtained
as
δ δ
hqf , tf | T Q(t1 ) · · · P (tn ) |qi , ti i = (−i)n
··· hqf , tf |qi , ti if,h=0 .
δj(t1 ) δh(tn )
(3.16)
So far we introduced the sources j(t) and h(t) as auxiliary functions to effec-
tuate calculation of n-point functions. In fact, a source, has a clear physical
meaning, as an external disturbance on the system. It is an external function
that measures how much the energy of the system is altered by adding a linear
deformation of position and momentum.

3.1.3 Transition amplitude between arbitrary states

In the previous sections we studied the time-ordered transition amplitudes


and n-point correlators between two position eigenstates. How about transition
amplitudes and n-point correlators between arbitrary states? These can be
obtained from inserting complete set of instantaneous position eigenstates in
hψf , tf |ψi , ti i as
Z
hψf , tf |ψi , ti ij,h = dqi dqf ψf∗ (qf )ψi (qi ) hqf , tf |qi , ti ij,h . (3.17)

In particular we will be interested in the vacuum-to-vacuum transition ampli-


tude. This is precisely the quantity we need to describe a scattering amplitude.
In a scattering experiment two beams of particles come from spatial infinity,
collide and produce new particles which then propagate toward the spatial in-
finity again. In the lab frame the situation state at infinite past is the vacuum,
then stuff happen in the intermediate times and the state in the infinite future
is again the vacuum. The corresponding generating function is given by
Z
h0|0ij,h = lim dqi dqf ψ0∗ (qf )ψ0 (qi ) hqf , tf |qi , ti ij,h (3.18)
ti →−∞,tf →+∞

This quantity can be obtained directly from the path integral by employing the
following trick. Suppose that the state at initial time is some generic state |ψi i:
Z
|ψi , ti i = dqi ψi (qi ) |qi , ti i .

The latter ket can be expanded in the energy eigenstates as


X
|qi , ti i = eiHti |qi i = eiEn ti |ni hn|qi i
n=0

where En denote the energy eigenvalues, we used (3.8) and assumed a discrete
energy spectrum for simplicity. Now, instead of taking the initial time to −∞
3.1. PATH INTEGRALS IN QUANTUM MECHANICS 31

as in (3.18) consider taking it to −∞ with a small imaginary component

lim |qi , ti i = eiHti |qi i = |0i h0|qi i ,


ti →−∞(1−i)

where  > 0 and infinitesimal. We managed to single out and project onto the
vacuum state because the energy of the vacuum state is defined to be 0 and all
the excited states have positive energies. We then find,
Z
lim |ψi , ti i = dqi ψi (qi )ψ0∗ (qi ) |0i = ci |0i ,
ti →−∞(1−i)

where ci = h0|ψi i. By the same token we have

lim hψf , tf | = c∗f h0| ,


ti →+∞(1−i)

where cf = h0|ψf i. Now the limits ti → −∞(1 − i) and tf → +∞(1 − i) in
the path integral can be effectuated by replacing the time integral on the real
axis in the action in (3.6) or (3.7) by a contour C that is slightly tilted toward
the lower half plane, see figure 3.1.3. All in all we arrived at the result1
Z R
h0|0ij = (ci c∗f )−1 Dq(t) ei C dt (L+jq) , (3.19)

where the contour C is given in figure 3.1.3.

Normalisation: There are two ambiguities in normalisation of this expres-


sion, first from the unknown coefficients ci and cf , second from the ambiguity
of including an overall constant in defining the path integral Dq itself. We can
fix them altogether by the following physical requirement: the probability am-
plitude for the vacuum state to remain vacuum in the absence of any source
perturbing it should be 1, that is h0|0i0 = 1:
Z R
∗ −1
(ci cf ) Dq(t) ei C dt L = 1 .

You may be confused that the LHS seems to depend on the choice of the initial
and the final states in the path integral but this is an illusion. There is also a
secret multiplicative dependence coming from the boundary conditions on the
path integral. In particular for the choice2 ψi and ψf vacuum then all the
multiplicative factors drop and we have Therefore we demand
Z R
h0|0ij = Dq(t) ei C dt (L+jq) , (3.20)

with normalisation Z R
Dq(t) ei C
dt L
= 1.

Our final result (3.20) is a great simplification because the entire dependence on
the initial and final states in the path integral disappeared in this expression.
1 From now on I present the result with only the position source j turned on. This is

immediately generalised to include the momentum source h using (3.6) instead of (3.7).
2 This is what I will assume from now on.
32 CHAPTER 3. PATH INTEGRAL QUANTISATION

Figure 3.2: Definition of the contour C that defines the vacuum-to-vacuum


amplitude.

They are all absorbed into the normalisation of the vacuum state. All of this
happens thanks to the “i trick” we used above by rotating the time integral
infinitesimally into the lower half plane. This projects the initial and final states
in the path integral to the vacuum state.

3.2 Path integrals in QFT

All the notions we introduced in the previous section, path integrals, gener-
ating function, n-point correlators, vacuum-to-vacuum amplitude etc carry over
to quantum field theory. Here, however we define the path integral over the
field configurations. This means that the propagation amplitude of a field from
a given field configuration φi (~x, ti ) at initial time ti to another field configura-
tion φf (~x, tf ) at final time tf is given by summing over all possible field paths
in the intermediate times which start and end on these configurations weighed
by exponential of the action.

Consider for example the free real scalar field. The Hamiltonian density for
this Klein-Gordon field is given by
1 2 1 1
H= Π + (∇φ)2 + m2 φ2 , (3.21)
2 2 2
where Π is the canonical momentum that is given by the variation Π = ∂L/∂ φ̇
of the Lagrangian density
1 1
L = − (∂φ)2 − m2 φ2 . (3.22)
2 2
To generalise the path integral of the previous section to this field we replace
q(t) → φ(~x, t), p(t) → Π(~x, t) and j(t) → J(~x, t)
3.2. PATH INTEGRALS IN QFT 33

With these replacements the propagation amplitudes (3.6) and (3.7) become
Z Z φf R tf 3
hφ(~x, tf )|φ(~x, ti )i = DΠ(x) Dφ(x)ei ti dtd x(Πφ̇−H) (3.23)
φi

for tf > ti , recalling that path integral automatically orders in time, and,
Z φf R tf 3
hφ(~x, tf )|φ(~x, ti )i = Dφ(x)ei ti dtd x L . (3.24)
φi

Similarly, the vacuum-to-vacuum amplitude in the presence of a source J be-


comes Z
4
R
Z0 [J] = h0|0iJ = Dφei C d x(L+Jφ) (3.25)
where the contour is the same as in figure3.1.3. Recall that in this final expres-
sion we do not have to specify the boundary conditions on the φ path integral,
as the infinite past and future time limit with the slight tilt of contour C (see
fig. 4) automatically projects onto the vacuum state regardless of the boundary
conditions in time.

3.2.1 Generating function: free case

We can calculate the generating function Z0 [J] in (3.25) directly. This is


easiest done by Fourier transforming to the 4-momentum space. Introduce the
Fourier transforms:
d4 k −ikx
Z
φ̃(k) = d4 xe−ikx φ(x) , φ(x) =
R
e φ̃(x) , (3.26)
(2π)4
d4 k −ikx ˜
Z
˜ = d4 xe−ikx J(x) , J(x) =
R
J(k) e J(x) . (3.27)
(2π)4
Substituting these in the action (3.22) we find
d4 k n
Z Z
4 1 2 2 ˜ φ̃(−k) + J(−k)
˜
o
S= d xL = − φ̃(k)(k + m )φ̃(−k) + J(k) φ̃(k) .
C 2 C̃ (2π)4
(3.28)
Note that we changed the contour C to C̃ in the integral when we pass from
spatial to momentum
R∞ integration.
 In order to preserve the form of the Fourier
transforms −∞ d0k exp −ik 0 t as one shifts from t → t exp(−i), one needs
to shift k 0 → k 0 exp{+i}. Therefore the contour in the momentum space is
slightly tilted toward the upper half plane as shown in figure 3.2.1. We now
Fourier transform the integration variable in the path integral (3.25) from φ(x)
to φ̃(k). This is a unitary transformation hence it has unit Jacobian. We can
now evaluate the path integral
Z R d4 k
i
{−φ̃(k)(k2 +m2 )φ̃(−k)+J(k)
˜ φ̃(−k)+J(−k)
˜ φ̃(k)}
Z0 [J] = Dφ̃(k)e 2 (2π)4 (3.29)

as a Gaussian integral, by making another change of variable3


˜
J(k)
φ̃(k) → φ̃(k) − .
k2 + m2
3 Again the Jacobian is unity as the RHS is linear in φ̃ with unit coefficient.
34 CHAPTER 3. PATH INTEGRAL QUANTISATION

Figure 3.3: Definition of the contour C̃ in the complex frequency space that
defines the vacuum-to-vacuum amplitude.

The result of the Gaussian integration is


− 21 ( Z
˜ J(−k)
˜
)
d4 k J(k)

i 2 2 i
Z0 [J] = N det[− (k + m ) exp , (3.30)
2 2 C̃ (2π)4 k 2 + m2

where N is a yet unknown multiplicative factor arising from multiplicative am-


biguities in defining the path integral. The determinant in front is independent
of J˜ hence it does not play an important role in calculation of the n-point cor-
relators. It is divergent hence should be defined carefully. We will do this later
in this course. Here we just note that all these ambiguities can be absorbed into
the normalisation of the generating function by writing:
( Z )
i ˜ J(−k)
d4 k J(k) ˜
Z0 [J] = Z0 [0] exp , (3.31)
2 C̃ (2π)4 k 2 + m2

where the contour of the k 0 integral is shown in fig. 5. We can now impose the
same physical requirement we imposed below equation (3.19), that the probabil-
ity amplitude for the system to remain in vacuum in the absence of any source
perturbing it is unity. This gives the normalisation condition

Z0 [0] = 1 . (3.32)

By the change of variables k 0 → k 0 ei we can put back the contour on the real
axis. Keeping track of the  we find the following expression for the exponent
Z ∞ ˜ J(−k)
˜
J(k)
dk 0 d3 k ,
−∞ k + m2 − i
2

up to O(). Then, using (3.27) the generating function is rewritten as


 Z 
i 4 4
Z0 [J] = exp d xd yJ(x)∆(x − y)J(y) (3.33)
2
3.2. PATH INTEGRALS IN QFT 35

Figure 3.4: Pole structure in the Feynman propagator.

where
d4 k eik(x−y)
Z
∆(x − y) = . (3.34)
(2π)4 k 2 + m2 − i
This is called the Feynman propagator. Poles structure in the Feynman propaga-
tor is shown in figure 3.2.1. The time ordered propagator for the Klein-Gordon
field obtained from the generating function is precisely the Feynman propagator:

1 ∂ 1 ∂
h0| T φ(x)φ(y) |0i = Z0 [J] = −i∆(x − y) . (3.35)
i ∂J(x) i ∂J(y) J=0

We can further perform the k 0 integral in (3.34) by using the Cauchy’s integral
theorem Z
1 f (z)
dz = f (z0 ) ,
2πi C0 z − z0
where the contour encloses the pole z0 . Using the pole structure shown in fig.
6 we find
Z Z
˜ ik(x−y) + iΘ(y 0 − x0 ) dke
∆(x − y) = iΘ(x0 − y 0 ) dke ˜ −ik(x−y) , (3.36)

where Θ is the Heaviside theta function. To arrive at this expression we closed


the contour C0 from above for y 0 > x0 and below for x0 > y 0 .

Let us define the

3.2.2 Wick’s theorem

Working out the same calculation that lead to (3.35), we find the following
result for the higher point correlators:

• If there are odd number of φs then the result vanishes when we set J = 0
at the end of the calculation:

h0| T φ(x1 )φ(x2 ) · · · φ(xn ) |0i = 0, for n odd . (3.37)


36 CHAPTER 3. PATH INTEGRAL QUANTISATION

• If there are even number of φs then the result is


1 X
h0| T φ(x1 )φ(x2 ) · · · φ(x2n ) |0i = n ∆(xi1 − xi2 ) · · · ∆(xi2n−1 − xi2n ) ,
i pairs
(3.38)

where the sum is over all different pairs. The result (3.37) and (3.38) is called
the Wick’s theorem.

3.2.3 Effective potential

The generating function in general can be written as exponential of an object


that is called the effective potential W [J]:
Z[J] = exp(iW [J]) . (3.39)
The meaning of the effective potential becomes clear when we recalling that
the generating function is, at the same time, the vacuum-to-vacuum transition
amplitude,
Z[J] = lim h0| e−iH[J]T |0i = lim exp(−iEvac T ) = exp(iW [J]) , (3.40)
T →∞ T →∞

R 4 H[J] is the hamiltonian corrected with the source term: H[J] = H −


where
d xJ[x]φ[x] and Evac is the energy of the vacuum state in the presence of
source. Therefore the effective potential measures how much the vacuum energy
is perturbed by adding an external source to the theory.

The source term itself has a clear physical meaning. The generating function
can be written as the expectation value of the source term,
Z R 4 R 4
Z[J] = DφeiS ei d xJ(x)φ(x) = h0| ei d xJ(x)φ̂(x) |0i . (3.41)

The operator in the last expression is reminiscent of coherent states in quantum


mechanics. Recall that a coherent state is a collective excitation in the quan-
tum system with the minimum uncertainty, hence a coherent addition to the
system that behave as a classical system as much as the quantum uncertainty
permits. The operator above is precisely the quantum field theory analog of this.
Recalling the expansion of φ in terms of creation and annihilation operators
Z  
φ̂(x) = dk ˜ â(~k)eik·x + ↠(~k)e−ik·x , (3.42)
R 
we see that exp i Jφ adds a collective excitation of particle-anti-particle pairs
to the vacuum. This is how we probe the quantum field theory, by adding a
classical like perturbation and see how the system reacts to it by computing the
n-point functions.

The function J(x) then determines the amplitude of this collective excita-
tion. For example we can choose this collective excitation to be localised around
two points in space by taking
J(x) = J1 (x) + J2 (x) ,
3.2. PATH INTEGRALS IN QFT 37

where the functions J1 and J2 has support around points x1 and x2 .

Let us determine how this perturbation modifies the energy of the vacuum
state by computing the effective potential W [J]. For a general source, by trans-
forming to Fourier space we have,
Z
1 1
W [J] = d4 k J˜∗ (k) 2 ˜ ,
J(k) (3.43)
2 k + m2 − i
where J˜ is the Fourier transform of J(x) and I used J˜∗ (k) = J(−k) that comes
from reality of J(x). For a source of the form above we have J(k)˜ = J˜1 (k)+ J˜2 (k)
˜
and there are four terms in (3.43) that arise from expanding J. These are W11
and W22 which correspond to self interaction of J1 with J1 , self interaction of
J2 with J2 and the cross interaction terms W12 and W21 between J1 and J2 . In
general we are interested in how the excitations interact with each other, hence
the term we are after is the latter. This is
Z
1 1
W12 [J] = d4 k J˜1∗ (k) 2 J˜2 (k) , (3.44)
2 k + m2 − i
This expression makes a significant contribution when k 2 +m2 = 0, that is when
the dispersion relation is satisfied. Thus, we learn that the interactions in QFT
are carried by particles that correspond to the excitations of the field! In addition
to this “on-shell” contribution there are contributions from k 2 6= m2 , which are
smaller. These are the “off-shell” contributions or “virtual excitations”.

To calculate the shift in the energy in the vacuum to this interaction let us
choose the sources as
J(x) = δ 3 (~x − ~x1 ) + δ 3 (~x − ~x2 ) .
This corresponds to test particles, that are infinitely massive as they are an-
chored to the space points ~x1 and ~x2 . Inserting the fourier transform
Z Z
0 0 ~
J˜1 (k) = d4 xe−ik·x δ 3 (~x − ~x1 ) = dx0 e+ik x e−ik·~x1

, etc. in (3.44) and carrying out the x02 integral which sets k 0 = 0, one easily
finds
~
d3 k e−ik·(~x1 −~x2 )
Z
1
W12 [J] = T , (3.45)
2 (2π)3 ~k 2 + m2
where I ignored the i in the denominator as the real part of the denominator is
positive definite. The multiplicative factor T is the total time of the interaction
and it comes from the dx01 integral which factored out exactly.
R

Comparison to (3.40) shows that the shift in the energy of the vacuum due
to the interaction between the two test particles is (W21 contributes equally)
~
d3 k e−ik·(~x1 −~x2 )
Z
E12 [J] = − . (3.46)
(2π)3 ~k 2 + m2

The ~k integral can be done exactly and one finds


1
E12 [J] = − e−m|~x1 −~x2 | . (3.47)
|~x1 − ~x2 |
38 CHAPTER 3. PATH INTEGRAL QUANTISATION

This is the potential energy of two test particles that are interacting by ex-
changing φ field excitations. Two remarks: We see that in the massless limit
this is the very similar to the Coulomb interaction in electrostatics. In fact, as
we will see later in the course, in the quantum theory of electromagnetism, we
find precisely the coulomb interaction which arise due to exchange of a mass-
less photon between the test particles. This type of potentials, including the
mass term, is called Yukawa potential. Secondly, we find that the interaction
potential is negative definite! This shows that, in this particular case of the real
scalar field the presence of delta-function sources lowers the total energy of the
system rather than increasing, hence the interaction, which arise from coupling
the test particles to the quantum field φ turns out to be attractive! This last
point i.e. the sign of the potential, depends on the theory under consideration.
Chapter 4

Interactions in QFT

Interaction terms in quantum field theory are given by the terms higher
than quadratic order in the Lagrangian (or Hamiltonian). This can be un-
derstood as follows: As discussed in the previous section the quadratic terms
in the Lagrangian determine the propagator which describe propagation of a
point-like particle from one space time point to another. Particle interactions
on the other hand are described by exchange of another particle between two
particles. In order to exchange a particle the initial particle should emit one.
This emission process necessitates presence of at least 3 particles at the same
space-time point:incoming particle, emitted particle, outgoing particle. This
spacetime point which includes at least three particles is called an interaction
vertex. As the single-particle states are created and annihilated by the quantum
field φ(x), see e.g. (2.21), presence of an interaction vertex requires a cubic or
higher order term in the Lagrangian. These terms are called the interaction
terms.

4.1 Generating function: interacting case

We will now calculate the generating function of an interacting QFT. We


will develop the necessary techniques in the path integral formalism using a
simple example: the real scalar field with a cubic interaction term with the
lagrangian density

1 1 g
L = − (∂φ(x))2 − m2 φ(x)2 + φ(x)3 . (4.1)
2 2 3!

Here g is a constant called the coupling constant and we inserted 3! for simplicity,
as will become clear below. To calculate the generating function of this theory
we will employ the following trick.

One can obtain a formal expression for the generating function with inter-

39
40 CHAPTER 4. INTERACTIONS IN QFT

actions using the following trick,


∞ ∞  l !n
(gq l )n jq X ∂ l
gq l +jq
ejq = eg( ∂j ) ejq .
X ∂
e = e = g
n=0
n! n=0
∂j

This trick can easily be generalised to functional derivatives (3.14) in place of


ordinary derivatives. Replacing q with φ(x) and j with J(x) we find
l
d4 x(gφ(x)l +J(x)φ(x)) d4 xg ( δJ(x)
δ
)
R
d4 xJ(x)φ(x)
R R
e =e e .

Now we apply this trick to an interacting QFT with a generic interaction given
by a polynomial in φ(x). Denoting this polynomial Lint and the free, quadratic
part of the Lagrangian L0 we have
Z R 4
R 4 δ
Z[J] = Dφ(x)ei d x(L0 +Lint +Jφ+) = ei d xLint [−i δJ(x) ] Z0 [J] . (4.2)

As we have already calculated Z0 [J] in (3.33) (4.2) is the final expression for the
interacting generating function, at least formally. Unfortunately this expression
cannot be evaluated analytically in most of the cases and we will apply a pertur-
bative expansion to calculate. In particular we will expand the first exponential
in a Taylor expansion when the coupling constants are small. This perturba-
tive evaluation of the generating function is best described in the language of
Feynman diagrams that we turn to in the next section.

4.2 Perturbative expansion and Feynman dia-


grams

4.2.1 Feynman rules

We will evaluate the generating function in (4.2) by expanding the expo-


nentials in (4.2) and (3.33) in a Taylor series. To be definite let us consider a
cubic interaction term Lint = g/3!φ3 . We obtain

∞ Z 3 !V X∞  Z P
X 1 ig 4 δ 1 i 4 4
Z[J] ∝ d x −i d yd zJ(y)∆(y − z)J(z) .
V! 3! δJ(x) P! 2
V =0 P =0
(4.3)
where the proportionality constant will be determined below. The letters P
and V stand for the number of propagators ∆ and vertices in this expression.
Now consider a specific term (P, V ) in the double sum. When the functional
derivatives hit the sources the leftover number of Js is given by

E = 2P − 3V . (4.4)

The letter E stands for the number of “external” sources. There are 3V func-
tional derivatives in this expression acting on 2P sources in 2P (2P −1) · · · (2P −
3V + 1) = 2P !/(2P − 3V )! combinations. Most of these combinations will be
4.2. PERTURBATIVE EXPANSION AND FEYNMAN DIAGRAMS 41

Figure 4.1: Feynman rules for the real scalar field with cubic interaction.

identical however. To organise these terms in an efficient way we introduce the


following “Feynman rules”: We denote a propagator −i∆(x − y) by a solid line,
an external source by a line with a blob at the end, and a cubic vertex by a
tri-star. We present the Feynman rules in figure 4.2.1 and give some examples in
figure 4.2.1. It is better to organise the expansion using the pair (E, V ), that is
the number of external sources and vertices. This is because we are eventually
interested in computing the n-point correlators that is obtained from the gen-
erating function by taking n functional derivatives with respect to the sources
J and setting J = 0 in the end. Therefore the only terms that contribute to
the n-point correlators have n = E. Furthermore, we define the perturbative
expansion for a small value of the coupling constant g, which makes possible to
expand in the number of vertices. In order to calculate an n-point correlator
then one needs to calculate all the Feynman graphs with E = n to the desired
order in the number of vertices.

4.2.2 Symmetry factor of diagrams

We will now determine the multiplicity factor in each Feynman diagram. In


a diagram with E external sources, V vertices and P = (E + 3V )/2 propagators
the naive multiplicity factor is V !×P !×(2!)P ×(3!)V . The first factor arise from
swapping the different vertices, the second arises from swapping different prop-
agators and the last two terms arise from swapping the end points in individual
propagator and vertex terms. This nicely cancels out the multiplicative factor
in the double Taylor expansion in (4.3). However this naive result overlooks the
fact that a some rearrangement of the derivatives give the same match-up to
sources as some rearrangement of sources. We have to divide the contribution of
each diagram by this factor which is called the symmetry factor. Therefore each
42 CHAPTER 4. INTERACTIONS IN QFT

Figure 4.2: Some examples of Feynman diagrams with different number of ex-
ternal sources E and vertices V constructed using the Feynman rules. Number
of propagators follow using the formula (4.4).

Feynman diagram has a symmetry factor associated to it. Clearly this mach-up
of derivatives with sources is represented by the symmetry of a diagram under
the simultaneous exchange of the legs in a vertex with the propagators that
connect to these legs, and a simultaneous exchange of two vertices and the end-
points of the propagators that connect these two vertices. Some examples are
presented in figure 4.2.1. As clear from the explanation above and the examples
in figure 4.2.2, the symmetry factor is literally the number of permutations of
vertices, propagators and external sources that leave the diagram invariant.

4.2.3 Example: One point function

As an example let us work out the O(g) contribution to the one-point func-
tion. The one-point function is obtained by taking a single derivative w.r.t the
source of the generating function(4.3) and setting J = 0 in the end. Therefore
only diagrams with E = 1 give a non-trivial contribution. We can organise all
these diagrams with E = 1 according to their number of vertices, that is, we can
expand the one-point function in powers of g V . As obvious from (4.4) there is
no contribution for E = 1 and V = 0. The first contribution arises from E = 1,
V = 1 which requires having P = 2. Let us calculate the corresponding term in
4.2. PERTURBATIVE EXPANSION AND FEYNMAN DIAGRAMS 43

Figure 4.3: Some examples of Feynman diagrams with their symmetry factors.

the generating function


Z  3  Z 2
ig δ 1 i
Z[J] ∝ d4 x −i d4 yd4 zJ(y)∆(y − z)J(z) .
3! δJ(x) 2! 2

Clearly all the contributions in this expression are of the same form, one δ/δJ
hitting one of the J∆J terms and the two other δ/δJ hitting the other J∆J.
The multiplicity is 2 (for choosing the J∆J that is hit by a single δ/δJ, 3 (for
choosing the δ/δJ that hits that J∆J) 2 (for choosing which of the Js in that
J∆J it hits) and finally 2 (for the order of the leftover Js hitting the other
J∆J, which gives 24. Note that in the final contribution we have 2 instead of
4 because exchanging two δ/δJs simultaneously with the two Js in J∆J gives
the same. This is the origin of the mismatch between the overall normalisation
term in the expansion of the generating function that is (2!)P (3!)V V !P ! = 42
and the multiplicity we found above, 24. This means that the symmetry factor
is 2. The result is,

−i
Z
δ
∆(0) d4 z ∆(x − z) .

h0| O(x) |0i = −i Z[J] = (4.5)
δJ(x) J=0 2

All this computation greatly simplifies by using the Feynman diagrams and
Feynman rules: The only Feynman diagram with a single external source and
two propagators is given by the topmost diagram in figure 4.2.1. The symmetry
factor of the diagram is clearly 2. Using the Feynman rules in figure 4.2.1 we
immediately arrive at the same result as above.

4.2.4 Connected vs. disconnected diagrams

All of the examples of Feynman diagrams we present above are called the
connected diagrams: we can trace a path through the diagram between any
44 CHAPTER 4. INTERACTIONS IN QFT

Figure 4.4: Example of a disconnected Feynman diagram with V = 4, E = 2


and P = 7.

two points on it. A generic diagram, on the other hand, consists of several
disconnected pieces. These arise in the sum (4.3) as a product of disconnected
integrals. An example is given in figure 4.2.4. Therefore the most general
diagram can be represented by the following expression:
1 Y n
D{n1 ,n2 ,··· } = (CI ) I . (4.6)
SD
I

Here we label all possible connected Feynman diagrams with the index I, CI
represent their values including the symmetry factors, nI is the number of the
same diagram I appearing in the product, and SD is the symmetry factor as-
sociated with the product itself. The latter gives for example r! if there are r
identical diagrams present in the sum. Generally it is given by
Y
SD = nI ! . (4.7)
I

Then the generating function (4.3) can be written in terms of the values corre-
sponding to the Feynman diagrams generically as follows:
!
X X Y nI
X
Z[J] ∝ D{n1 ,n2 ,··· } = nI ! (CI ) = exp CI , (4.8)
{n1 ,n2 ,··· } {n1 ,n2 ,··· } I I

where we exchanged the sum and the product in the last step and summed up
each of the sums over nI into exponentials. This is a remarkable result: the
generating function in general can be written as the exponential of sum over all
possible connected diagrams.

Now we can impose the normalisation condition that the vacuum should
remain vacuum in the absence of any source, that is Z[0] = 1. Note that this is
in general a different condition than the previous one Z0 [0] = 1 in (3.32) because
the vacuum state in the presence of interactions is generically different than the
free case. We denote both vacua by |0i with a slight abuse of notation. In the
absende of any sources, Feynman diagrams only include the so called vacuum
4.3. AN INTRODUCTION TO RENORMALISATION IN QFT 45

Figure 4.5: Examples of vacuum diagrams.

diagrams, diagrams with no sources, that only involve closed loops, see figure
4.2.4. Denoting the set of these vacuum diagrams by {0} the normalisation
condition above means
 
X 
Z[0] = exp CI = 1 ,
 
I∈{0}

thus we arrive at our final result


 
X
Z[J] = exp  CI  . (4.9)
I ∈{0}
/

4.3 An introduction to renormalisation in QFT

4.3.1 Loop divergences

The final expression we obtained for the generating function in (4.9) is not
completely well-defined. For one thing it involves infinitely many divergent
terms! To see this in a simple example let us consider the one-point function.
Using the (4.4) formula for E = 1 we see that the lowest order contribution
in g to the on-point function comes from the term with V = 1, P = 2. The
corresponding Feynman diagram is the top figure in figure 4.2. We can easily
evaluate this diagram using the Feynman rules in figure 4.1. The result of this
calculation is Z
∆(0)
g d4 xd4 y J(x) ∆(x − y) ,
2
where the factor of 1/2 comes from the symmetry factor of the diagram and the
propagator ∆(y − y) = ∆(0) comes from the propagator in the loop. The one-
point function is obtained from this by taking the functional derivative w.r.t.
the source:
Z
δ ∆(0)
h0| φ(x) |0i = −i Z[J] = −i g d4 y ∆(x − y) + O(g 2 ) , (4.10)
δJ(x) 2
46 CHAPTER 4. INTERACTIONS IN QFT

where the higher order contributions arise from Feynman diagrams with a single
source and higher number of vertices. This expression is divergent because the
factor in front ∆(0) is infinite. From (3.34) we have
d4 k
Z
1
∆(0) = .
(2π)4 k 2 + m2 − i
Recalling the pole structure of the propagator, shown in figure 3.4, we see that
we can rotate the integral of k 0 from −∞ to +∞ on the real axis to an integral
−i∞ to +i∞ on the imaginary axis. This is because, the contour C shown in
figure ?? does not contain any poles. Make then the change of variables k 0 = ikE
with k 2 = −(k 0 )2 + ~k · ~k = kE
2
+ kx2 + ky2 + kz2 ≡ |k|2 defined a 4-sphere with
radius |k|. This rotation is called the Wick rotation. After the Wick rotation,
the integral can be written as
Z ∞
d|k| |k|3
∆(0) = iΩ3 ,
0 (2π) |k| + m2 − i
4 2

where Ω3 = 4π/3 is the volume of a unit 3-sphere.

This integral is clearly divergent. These type of divergences are ubiquitous


in QFT: they appear whenever there is a closed loop in a Feynman diagram
which instructs us to integrate over the loop 4-momentum from −∞ to +∞.
One way to deal with this problem is to place a high 4-momentum cut-off, that
is to replace the upper limit of the integral with a large but finite number
Z ΛE
d|k| |k|3
∆(0) → iΩ3 ,
0 (2π) |k| + m2 − i
4 2

and then take ΛE → ∞ at the end of the calculation hoping that this diver-
gence does not show up in the observables. This is called the regularization
procedure. However, we need to be very careful with the way we regularize
the loop divergences: we should make sure that we preserve the symmetries of
the action in the regularization procedure. For instance the UV cut-off we just
describe destroys the Lorentz invariance of the theory, that maps to rotational
invariance in 4D in the Euclidean momentum space above. There are different
ways to regularize which respects the Lorentz symmetry: dimensional regular-
ization, Pauli-Villars regularization, differential regularization etc. Pauli-Villars
instructs us to add another hypothetical highly massive particle to the theory
and replace the propagator of the scalar field as
 2
1 1 Λ
→ 2 .
k 2 + m2 − i k + m2 − i k 2 + Λ2 − i
The k-integral with this replacement is now finite. The result of the integral is
i
∆(0) → Λ2 . (4.11)
16π 2
We still need to take Λ → ∞ at the end of the calculation. The only thing
we achieved by this regularization is that we made the one-point function, and
the generating function in general well-defined at the expense of introducing
another parameter in the theory. The final results should not depend on this
unphysical cut-off Λ and the way to make that such an unphysical dependence
cancels out is called the renormalization procedure.
4.3. AN INTRODUCTION TO RENORMALISATION IN QFT 47

Figure 4.6: The new Feynman rule associated to the linear counter-term.

Figure 4.7: Examples of Y-vertex contributions to Feynman graphs. Here we


show possible contributions to the one-point function h0| φ̂(x) |0i with the indi-
cated number of Y and g-vertices.

4.3.2 Counterterms

In these lectures I will not give a complete account of renormalization in


QFT which is a very important and rich subject. I will instead take a practical
approach and explain the minimum amount of modifications we need to make
in the original theory to make sure that the generating function is well-defined.
I will again do this using the example of the one-point function. Now, imagine
that the action we started with (4.1) is problematic for some reason – we will see
what the problem is shortly – and we need to modify it to make it well-defined.
The practical approach is to find the minimum number of modifications needed.
In the case of the one-point function, consider adding a linear term to the La-
gragian L → L + Y φ(x) where Y is a new coupling constant. Repeating the
calculation in section (4.2.1) with this new interaction term we find that we need
to add a new Feynman rule, that is a new vertex with a single line (because the
interaction is linear in φ), that we show in figure ??. This new rule should be in-
cluded in the Feynman graphs. Possible contributions to the one-point function
are shown in figure 13 with the various number of Y and g-vertices. Now, let’s
see that the divergences we encountered in the last section can completely be
cancelled by adjusting this new vertex. Anticipating that Y should be chosen
O(g) – to at least have a chance of cancelling the divergence (4.11) in (4.10)
– we see that there are two contributions to the one-point function now: the
one from the g-vertex that we calculated in (4.10) and the new, lowest order
contribution from the new vertex, that is the topmost diagram shown in fig. 13.
48 CHAPTER 4. INTERACTIONS IN QFT

Adding these we find,


 Z
1
h0| φ̂(x) |0i = −i iY + g∆(0) d4 y∆(x − y) + O(g 3 ). (4.12)
2

We learn that we can cancel the divergence completely by setting

1
Y =− g∆(0) .
32π 2
Therefore including a new vertex in the theory allows us to cancel the unwanted
divergences. The divergences that arise from loops in the Feynman graphs that
are higher order in g can similarly be cancelled by correcting Y to the desired
order.

4.3.3 Renormalisation procedure

All of this is called the renormalisation procedure. To recap the procedure


is: first regularise the theory by including a cut-off Λ, then add new terms to the
Lagrangian a.k.a the counterterms, and finally choose them wisely to cancel the
loop divergences. The result will be well-defined, finite n-point functions. As,
you should have noticed, there is a lot of ambiguity in this procedure: i) how do
we know which counterterms to add? ii) How do we fix these terms precisely?
1
e.g. why not to choose Y = − 32π 2 g∆(0) + constant which would still cancel

the divergence but would yield an arbitrary finite contribution to the one-point
function. iii) Do we need to introduce an infinite number of counterterms to
cancel all possible divergences in the Feynman graphs? If so, the theory would
have absolutely no predictive power!?

In this introductory course I will only provide answers to these questions


without proofs:

• The answer to i) is add all possible polynomials in φ up to the maximum


order of φ in the original Lagrangian, that is φ3 in the present case. Then
the full counterterm action for the real scalar theory is

1 1 1
Lct = − (Zφ −1)∂ µ φ∂µ φ− (Zm −1)m2 φ2 + (Zg −1)gφ3 +Y φ , (4.13)
2 2 3!
where the coefficients Zf , Zm , Zg and Y are to be fixed below. We
parametrised these coefficients so that when added to the original La-
grangian, the total Lagrangian reads

1 1 1
Ltot = − Zφ ∂ µ φ∂µ φ − Zm m2 φ2 + Zg gφ3 + Y φ . (4.14)
2 2 3!

• The answer to ii) is: renormalisation conditions. These are certain phys-
icality requirements that determine the counterterms solely by physical
requirements. We discuss them in the next section.
4.3. AN INTRODUCTION TO RENORMALISATION IN QFT 49

• The answer to iii) is: no we don’t need an infinite number of counterterms.


We only need the four terms in (4.13) in this theory. This is sufficient
to fix all the divergences! In general, when we need a finite number of
counterterms the theory is called renormalisable. A general rule of thumb
to determine whether a theory is renormalisable is to require that the
mass dimension of the coupling constants in the original Lagrangian are
non-negative. To calculate the mass dimension of the field [φ] just require
that the Lagrangian should be dimensionless. Then from the kinetic term
and the rest we learn that

[φ] = 1, [m] = 1, [g] = 1 . (4.15)

Hence the theory is renormalisable according to the rule. A quartic inter-


action g4 φ4 would still be renormalisable, but a quintic one g5 φ5 would
not, as [g5 ] = −1.

4.3.4 Renormalisation conditions

Now we will answer the question how to fix the counterterms unambiguously.
Chapter 5

From theory to experiment:


scattering in high energy
physics

50
Chapter 6

Representations of the
Lorentz group and particles

6.0.1 Group representations

We only discussed a specific type of quantum field up to here: the scalar


field φ. Under a Lorentz transformation
x̄µ = Λµ ν xν ,
this field transforms as a scalar:
φ̄(x̄) = φ(x) ⇒ φ(x) → φ̄(x) = φ(Λ−1 x) , (6.1)
hence we say that φ transforms trivially, that is the field in the transformed
frame is the same as in the old frame. Here is an example of field that transforms
non-trivially: ∂µ φ(x). By taking the derivative of the expression above and
making a change of variables we find,
Λνµ ∂¯ν φ̄(x̄) = ∂µ φ(x) ⇒ ∂µ φ(x) → ∂µ φ(x) = (Λ−1 )νµ (∂ν φ)(Λ−1 x) .
(6.2)
This is a non-trivial transformation because not only the index of the field but
also the field itself transforms ∂µ φ → (Λ−1 )νµ ∂µ φ by a factor of Λ−1 . This
is called the covariant vector representation. In general a covariant vector Aµ
transform as
Aµ → (Λ−1 )νµ Aν .
Since contraction of this with a vector with an upper index B µ Aµ is a scalar,
we learn that the vectors with upper indices transform as
B µ → Λµν B ν .
This is called a contravariant vector representation. By repeating this calcula-
tion we learn that a general object with n number of lower and m number of
upper indices transform as
···βm
Cαβ11 ···α n
→ (Λ−1 )µα11 · · · (Λ−1 )µαnn Λβν11 · · · Λβνm
m
Cµν11 ···ν
···µn .
m
(6.3)

51
52CHAPTER 6. REPRESENTATIONS OF THE LORENTZ GROUP AND PARTICLES

This general object is called a tensor representation. Classification of all possible


transformation rules under a group action is called the representation theory.
In order to find all possible representations under the Lorentz group we need to
work out the representation theory of the Lorentz group.

Let us first establish that the Lorentz transformations indeed form a group:

(i) If Λ and Λ0 are Lorentz transformations, then Λ00 = ΛΛ0 is also a Lorentz
transformation.
(ii) There exists an identity Λµν = δνµ
(iii) For ∀Λ, there exists an inverse Λ−1
(iv) Associativity: (Λ · Λ0 ) · Λ00 = Λ · (Λ0 · Λ00 ).

It is easy to show these properties using only the definition of a Lorentz trans-
formation:
ηµν Λµα Λνβ = ηαβ (6.4)
Let us denote a generic transformation rule as
φA → φ̄A = LB
A (Λ)φB ,
−1 µ1
where A, B denote a generic collection of indices. For example LB A = (Λ )α1 · · · (Λ−1 )µαnn
β1 βm
Λν1 · · · Λνm for a tensor. Now, a representation is defined by the following prop-
erty:
0 0
LB C C
A (Λ)LB (Λ ) = LA (Λ · Λ ) . (6.5)
The representation should become unity for a trivial Lorentz transformation:
LB B
A (1) = δA . From this it follows that
−1
LB
A (Λ ) = (L−1 )B
A (Λ) . (6.6)
Representations can be worked out by expanding the group action near unity:
Λµν = δνµ + δwνµ + O(δw2 ) . (6.7)
Inserting this in the definition (6.4) we find that w should be an antisymmetric
4 by 4 matrix. Therefore it has 6 independent components which correspond to
3 rotations and 3 boosts. Denoting a unit vector by n̂ we can parametrize them
as follows:
Rotations: δωij = −ijk n̂k δθ ≡ −ijk δθk generates a rotation around the k-
axis with angle δθ,
Boosts: δωi0 = n̂i δη ≡ δηi generates a boost in the i direction with rapidity
δη.
Now consider a generic representation L(Λ). Using LB B
A (1) = δA and demanding
continuity we find that, for a transformation Λ close to unity, L itself should
have an expansion near identity:
i ˆ µν )A ,
LB B
A (1 + δω) = δA + δωµν (M B (6.8)
2
where M µν are called the generators. They are a collection of N by N dimen-
sional matrices where N is called the dimension of the representation. They
should be anti-symmetric under the exchange of Lorentz indices.
53

6.0.2 Lorentz generators and their algebra

Starting from the generic properties of a representation (6.5) and (6.6) we


have
L−1 (Λ)L(Λ0 )L(Λ) = L(Λ−1 Λ0 Λ) .
Assuming that Λ0 is close to unity Λ0 = 1 + δw0 and expanding in δw0 we find
0
δwµν L−1 (Λ)M µν L(Λ) = δwµν
0
Λµρ Λνσ M ρσ
which needs to hold for all δw0 , thus we have:
L−1 (Λ)M µν L(Λ) = Λµρ Λνσ M ρσ . (6.9)
This is how the generators transform under a Lorentz transformation. Note
that this transformation rule is consistent with the general rule (6.3) that we
established above. Assuming that Λ is close to unity Λ = 1 + δw and expanding
(6.13) we discover that the Lorentz generators M should satisfy the Lorentz
algebra:
[M µν , M ρσ ] = iη µρ M νσ − iη νρ M µσ + iη νσ M µρ − iη µσ M νρ (6.10)
Lorentz generators generate the transformations associated with the parameters
δwµν that they couple to. In particular the boosts and rotations are generated
respectively by
M i0 ≡ Ki , M ij ≡ ijk Jk . (6.11)
Thus a general. transformation can be written as
1 1
δwµν M µν = δwi0 M 0i + δwij M ij = δηi K i − δθk J k . (6.12)
2 2
Decomposing the Lorentz algebra (6.10) into rotations and boosts we find
[Ji , Jk ] = iijk Jk (6.13)
[Ji , Kj ] = iijk Kk (6.14)
[Ki , Kj ] = −iijk Jk . (6.15)

6.0.3 Finite dimensional representations of the Lorentz


group

A very important result from the representation theory is that the repre-
sentations of a group algebra is in one-to-one correspondence with the repre-
sentations of the group elements that are continuously connected to the unit
transformation, that is, of the form (6.7). Generally the Lorentz group also
contain elements that are not connected to the identity transformation. They
are given by the parity and time-reversal transformations. For the moment we
ignore these elements–which we discuss parity and time-reversal in the appendix
– and focus on the connected subgroup. For this subgroup it suffices to work
out the representations of the Lorentz algebra (6.10) or (6.13-6.15). Define
1 1
JiL ≡ (Ji − iKi ), JiR ≡ (Ji + iKi ) . (6.16)
2 2
54CHAPTER 6. REPRESENTATIONS OF THE LORENTZ GROUP AND PARTICLES

From (6.13-6.15) we find that these combinations satisfy

[JiL , JjL ] = iijk JkL , [JiR , JjR ] = iijk JkR , [JiL , JjR ] = 0 , (6.17)

that is the algebra of two sets of SU(2)s. We already know the representations
of SU(2) algebra from a basic quantum mechanics course. A generic finite
dimensional representation of SU(2) is labelled by a non-negative (half)-integer
1 3
J = 0, , 1, , · · ·
2 2
and given by (2J + 1) × (2J + 1) dimensional matrices. These matrices have
2J + 1 eigenvalues that correspond to the value of J3 :

J3 : −J, −J + 1, · · · + J

In fact, these are the finite dimensional irreducible representations, usually ab-
breviated as an “irrep”. A more generic reducible representation is obtained by
tensor products of these. For example a 2(J + J 0 ) + 2 dimensional matrix that
can be put in a block diagonal form with 2J + 1 and 2J 0 + 1 dimensional blocks
is a reducible representation.

We conclude that the finite dimensional irreducible representations of the


Lorentz algebra (hence the connected Lorentz group) is labelled by two non-
negative (half)-integers (JL , JR ) that correspond to the subalgebras generated
by (6.16) and given by the tensor product of a 2JL + 1 and 2JR + 1 dimensional
matrices. Therefore the dimension of the representation is (2JL + 1) × (2JR + 1).
As the angular momentum generator is given by, see (6.16),

Ji = JiL + JiR

the states of different angular momentum contained in the representation (JL , JR )


are
|JL − JR |, |JL − JR | + 1, · · · JL + JR
. This means that, even though the representation (JL , JR ) is irreducible under
the full Lorentz algebra, it is generically reducible under the rotation subgroup.

Here are some examples of finite dimensional Lorentz irreps:

• Scalar, (0, 0): This is the trivial representation. According to the rule
above, it only contains a spin-zero state. The quantum field that is based
on this representation is nothing else but the scalar field φ(x) that we
studied above.
• Left and right handed Weyl spinors, ( 21 , 0) and (0, 12 ): These are spinor rep-
resentations that transforms as a 2 dimensional spinor under the left(right)
SU(2) and trivially under the right(left) SU(2) respectively. They only
contain a spin-1/2 state under rotations. The quantum field that is based
on these representation are called left(right)-handed Weyl spinor fields
and will be denoted by ψaL (x) ψaR (x). The index a = 1, 2 denotes the
components of the spinor with spins up and down respectively. These rep-
resentations play a key role in QFT and represent the the massless spin
55

1/2 fields. The Dirac spinor ψa is obtained as a tensor sum of the left and
right handed Weyl spinors. Therefore these representations form the basis
of the largest class of elementary particles, electrons, muons, quarks etc.
• Vector representation, ( 21 , 12 ): This is a 4 dimensional representation that
transforms as a (covariant or contravariant) vector. It contains two types
of spin states, a spin-zero and a spin-one. The quantum field that is based
on this representation is called a vector field and will be denoted by Aµ (x).
The index µ is the space-time index that runs from 0 to 3 and labels the
4 different components. The spin-zero and spin-one components can be
thought of the A0 and A ~ respectively1 . This representation plays a key role
in QFT and corresponds to the gauge field. When this field is massless, it
is nothing else but the photon!
• Self-dual and anti-self dual antisymmetric tensors, (1, 0) and (0, 1): These
are 3-dimensional representations that only contains a spin-one state under
rotations. The associated quantum field can be written as an antisymmet-
±
ric matrix and denoted by Bµν . Note that the property of antisymmetry
under exchange of space-time indices is preserved under Lorentz transfor-
mations. That is, the transformed matrix B̄µν = (Λ−1 )α µ (Λ
−1 β
)ν Bαβ is also
anti-symmetric as the Lorentz matrices are symmetric. Therefore the anti-
symmetry property is respected by the irreps, hence can be used to reduce
general reducible representations. An antisymmetric 4 by 4 matrix has 6
components whereas this representation is supposed to have 3. There is a
+ −
further decomposition of an antisymmetric matrix Bµν = Bµν +Bµν where
the individual components satisfy the duality and anti-duality conditions

Bµν+
= 21 µναβ (B + )αβ and Bµν = − 12 µναβ (B − )αβ . Note that these dual-
ity properties are also invariant under Lorentz transformations, hence are
respected by irreducible representations. The field strength of the gauge
field Fµν = ∂µ Aν − 6ν Aµ transforms as a sum of the B + and B − repre-
sentsations. They further play an important role in characterising certain
topological field configurations called instantons.
• Left and right-handed Rarita-Schwinger spinors, (1/2, 1) and (1, 1/2): These
are 6 dimensional representations that contain spin 1/2 and spin 3/2 states
each. The associated quantum fields possess one vector one spinor index
L R
and denoted as ψµ,a and ψµ,a . Here µ is a space-time vector index and
a = 1, 2 is a spinor index. This contains 4 x 2 = 8 components which are
further reduced to 6 by the gamma-tracelessness condition γ µ ψµ,a L,R
= 0
µ
that are two independent conditions, where γ is the Dirac-gamma ma-
trix. Note that this condition is preserved by the Lorentz transformations.
They play an important role in supergravity theories and represent the su-
persymmetric partner of the graviton, called the gravitino.
• Symmetric traceless tensor, (1, 1): This is a 9 dimensional representation
that can be represented by a symmetric matrix Gµν , which is 10 dimen-
sional that is further reduced by demanding tracelessness η µν Gµν = 0.
1 Thisdecomposition is not Lorentz invariant and holds only in a Lorentz center of mass
Lorentz frame. A Lorentz invariant way to decompose the spin states can p be obtained by
introducing the 4-velocity of the Lorentz frame uµ = γ(1, ~v ) where γ = 1/ 1 − |vecv|2 . Note
that u2 = −1 in our metric convention. Then the decomposition reads Aµ = −um u · A +
(Aµ + uµ u · A). The first component is spin-zero and the second one is spin-one.
56CHAPTER 6. REPRESENTATIONS OF THE LORENTZ GROUP AND PARTICLES

Notice that both symmetry and tracelessness is preserved by the Lorentz


transformations. This representation contains one spin-zero, one spin-one
and one spin-two states. The latter one is nothing else but the graviton!

The list goes on.

6.0.4 Representing the Poincare group in quantum me-


chanics

Reading the examples I listed above you should have noticed that the fi-
nite dimensional irreps of the Lorentz group are in one-to-one correspondence
with the different type of elementary (also non-elementary) particles. Indeed, in
general we label particle excitations by their spin, their mass and their charge
under the various global symmetry groups such as the electric charge, hyper-
charge, lepton number, baryon number etc. The most fundamental among these,
in the sense that it applies to all elementary particles is the mass and the spin.
This is indeed dictated by the Lorentz symmetry. There are other quantum
numbers that enter the labelling of a particle state, such as the 4-momentum
pµ , spin projected on some axis Ji etc. but these numbers are frame-dependent,
they change from one Lorentz frame to another. Only the Lorentz invariant
labels enter the classification, i.e. we do not say an electron is a particle with
momentum 2.5 MeV in some direction, we say it is a particle with mass 0.5
eV and spin 1/2. These invariant labels are indeed determined by the different
finite dimensional irreps of the Lorentz algebra as we discussed above.

The story is in fact more subtle. The reason for this subtlety is that in
a quantum theory we need to represent all symmetry operations by unitary
operators acting on the Hilbert space, because only unitary operators leave
the norm if a state invariant, hence preserve probabilities. That is, Lorentz
transformation acting on a quantum field in a specific representation is defined
by
φA (x) → φ̄A (x) = U (Λ)−1 φA U (Λ) = LBA φB (Λ
−1
x) , (6.18)
with U (Λ)−1 = U † (Λ). What we need to define an elementary particle is then
a unitary irrep. A general result from representation theory, on the other hand,
is that non-compact groups (groups with a non-compact transformation param-
eter, such as the boosts in the Lorentz group) do not possess any unitary finite
dimensional representations. This means that we need infinite dimensional rep-
resentations of the Lorentz group – as opposed to the examples discussed above
– in order to represent the Lorentz group in a quantum mechanical theory.

In fact we need infinite dimensional representations for a different reason


too: our full symmetry group is Poincare which consists of space-time transla-
tion in addition to the Lorentz transformations. Space-time translations act on
the Hilbert space by the unitary operator U (a) = exp(iaµ Pµ ) where Pµ is the
translation generator. Thus, one a one-particle state with definite 4-momentum
it would give
U −1 (a) |pi = e−ia·p |pi , (6.19)
57

that is to say 4-momentum of the particle is its “charge” under translations.


However, as we have seen above, Lorentz transformations change p and the irreps
of Lorentz contain states with all pµ but fixed mass and spin. But then how
can this Lorentz irrep, which is a collection of states with different p represent
translations? The way out is, as we have seen before, is to consider fields. A
field φ(x) indeed provide a unitary representation under translations:

U −1 (a)φA (x)U (a) = φ̄A (x) = φA (x − a) . (6.20)

This is a unitary operation because U (a) in a field representation is given by


U (a) = exp{ia · P } = exp{aµ ∂µ } which is a unitary operator. The reason that
representing translations as fields does not mess up the representation structure
under the Lorentz group is because a field φA (x) is comprised of one-particle
(and anti-particle) states with all different pµ . Let us consider the simplest
example of the free scalar field:
Z  
φ̂(x) = ˜ â(~k)eik·x + ↠(~k)e−ik·x .
dk (6.21)

First, let’s see how the translations (6.20) work out. When (6.20) applies to
(6.21) it acts on the creation and annihilation operators inside φ. From (6.19)
and the fact that |pi = a(p)† |0i we learn that

U −1 (a)a(p)† U (a) = e−ia·p a(p)† , U −1 (a)a(p)U (a) = e+ia·p a(p) , (6.22)

which immediately yields (6.20). At the same time it does not mess up the
transformation property under Lorentz because, similar to.(6.22) the Lorentz
generators act on the creation/annihilation operators as

U −1 (Λ)a(p)† U (Λ) = a(Λ−1 p)† , U −1 (Λ)a(p)U (Λ) = a(Λ−1 p) , (6.23)

therefore a change of integration variables k → Λk yields (6.18) for the trivial


representation, that is LB A = I. Note that, it is crucial to have an integral over
all k for this to work, so indeed representation of the Poincare symmetry in
terms of a field is sufficient and does the job. Representing the Poincare group
by fields turns out to also automatically solve the issue with unitary vs finite
dimensional representations of the Lorentz group. We will show this below. For
now, we can make the following precise definition of an elementary particle:

An elementary particle is a unitary irreducible representation of the Poincare


group.

As we discussed above this unitary representation is infinite dimensional and


given in terms of quantum fields. Now, it seems like our entire discussion of finite
dimensional representations of the Lorentz group in the previous section become
obsolete. Luckily, this is not the case. As Wigner showed in 1939, in fact, one
can construct these infinite dimensional unitary representations starting from
the finite irreps following a procedure called “induction from the Little group”.
58CHAPTER 6. REPRESENTATIONS OF THE LORENTZ GROUP AND PARTICLES

6.0.5 Wigner’s construction of unitary representations

First we identify the “Casimir operators” of the Poincare group. These are
operators that are quadratic in the group generators with the essential property
that they commute with all of the generators of the group. The Poincare group
is generated by Lorentz generators M µν and space-time translations P µ . It is
easy to show, by using the Poincare algebra, that there are only two Casimir
operators: the mass and the spin,
1
M 2 = P µ Pµ , J 2 = W µ Wµ , Wµ = µνρσ M νρ P σ , (6.24)
2
where W µ is called the Pauil-Lubansky vector. These are invariant under both
translations and Lorentz transformations thus can be used to label represen-
tations. Clearly the first operator corresponds to the invariant mass of the
representation.

To figure out the value of the second operator it is easiest to work in a


specific Lorentz frame. Clearly the value should be independent of this frame
choice. One should distinguish two cases M 2 > 0 and M 2 = 0. For M 2 > 0 one
can always choose the rest frame P µ = (m, 0, 0, 0). Then one finds J 2 = Ji Ji
that is the total spin operator whose eigenvalue is j(j + 1) in a representation
with spin j. Therefore we characterise a massive Poincare representation by its
mass M 2 and its total spin j. Both are conserved under time evolution because
both of the operators M 2 and J 2 commute with the Hamiltonian P 0 . They are
also clearly invariant under changes of Lorentz frame.

When M 2 = 0 there is no center-of-mass frame and the best one can do is


to fix the spatial momentum in a given direction, say z. Then P µ = (ω, 0, 0, ω).
In this frame one can show that J 2 = 0. Heuristically one can see this by
noticing that all the terms in J 2 are proportional to either (P~ · K)
~ 2 or (P~ × K)
~ 2
which should both vanish because we hold fix the momentum in this frame,
and also because a massless physical excitation cannot have non-trivial spin on
the plane of its motion. This means that W µ ∝ P µ with a single independent
component W 0 = P~ · J. ~ The representation in a generic frame is characterised
by the eigenvalue of W 0 /|P~ | which is indeed frame independent. This operator
is called the helicity operator:
~ ~
~h ≡ P · J . (6.25)
|P~ |
This is the value of spin in the direction of motion. It is quantised for the finite
dimensional representations discussed above. For a massless scalar h = 0, for
a massless Weyl spinor it is h = ±1/2. For a massless vector, h = ±1, for the
graviton h = ±2 etc. It is conserved under time evolution because both P~ and
J~ commute with P 0 . It is invariant under changes of Lorentz frame only when
M 2 = 0. Physically this is because one can only change the spin in the direction
of motion by a boost that reverses the direction of motion, which is impossible
for massless particles.

Now that we have identified the Casimir’s to characterise the representa-


tions, we move on to construct the infinite dimensional unitary representations
59

based on the finite dimensional representations of the Lorentz group above. The
problem of non-unitarity is caused by the boost generators because this is the
part of the transformations with a non-compact range parameters, rapidities.
We will see this non-unitarity explicitly when we discuss the spinors below.
Wigner’s idea is simply to get rid of these unwanted boosts by fixing a frame,
and then constructing the representations in that given frame.

Again, we have to distinguish the massive and the massless case. In the
massive case we go to the rest frame P µ = (m, 0, 0, 0). This means that all the
boosts are fixed, because a boost would change this fixed momentum. But the
rotation subgroup SO(3) leaves it invariant, and we need to identify the irrep
under rotations. This SO(3) is called the ”little group”. Finite dimensional
representations are just the usual spin states labelled by the (half)-integers
J = 0, 1/2, 1, · · · . In a given represenetation J there are 2J + 1 states −J, −J +
1 · · · + J. This representation is both finite dimensional and unitary.

For example, the field representation Aµ (x) for a massive vector particle
works as follows. Determine the wave-function of a given state with momen-
tum p~ for J = 1. This wave function µi (p) should be labelled by an index
i = 1, 2, 3 that corresponds to −1, 0, +1. In the special frame, they should sat-
isfy pµ µi (p) = 0 by Wigner’s construction. One can also normalise them by
requiring i,µ µi = 1. They are called the polarisation vectors. Then the generic
field representation can be obtained by coupling these wave-functions with the
creation and annihilation operators as usual and summing over all i and p~:
Z
µ
A (x) = dp ˜ e−ip·x ai (p)µ (p) + e−ip·x a† (p)µ∗ (p) . (6.26)
i i i

The trick works: even though the individual finite dimensional representations
µi (p), i.e. the wave-functions, transform non-unitarily under the Lorentz trans-
formations,
µ (p) → µ (p) = Λµν ν (p) (6.27)
— because the Λµν is a non-unitary matrix for the boosts — the infinite di-
mensional representation Aµ (x) transforms unitarily because the action of the
Poincare group on A is given in terms of its action on the basis of creation/annihilation
operators

U (Λ)−1 ai (p)U (Λ) = Λji ai (Λ−1 p), U (Λ)−1 a†i (p)U (Λ) = Λji a†i (Λ−1 p) ,
(6.28)
where Λji is a rotation generator and it is unitary. This means that the one-
particle states transform unitarily under the Lorentz transformations. Yet the
action (6.29) generates the desired transformation property of Aµ :

U (Λ)−1 Aµ (x)U (Λ) = Λµν Aν (Λ−1 x) , (6.29)

as you can easily show by using (6.29), (6.27) and a change of integration variable
p → Λ · p in (6.26).

In the massless case instead one fixes the boosts by choosing P µ = (ω, 0, 0, ω).
Then the little group is SO(2) that are rotations on the xy plane. Finite dimen-
sional representations are given by (half)-integers and they are labelled by the
60CHAPTER 6. REPRESENTATIONS OF THE LORENTZ GROUP AND PARTICLES

helicities h = 0, 1/2, 1, · · · . In a given representation there are only two states


s = +h, −h. This is a unitary representation. The infinite dimensional field
representation is determined by first obtaining the wave-functions associated to
these states for a given momentum, then coupling them to the corresponding
creation/annihilation operators and finally summing over the helicity states s
and momenta p~. For example the field representation for a massless left or
right Weyl spinor works as follows. The wave functions for each helicity state
and fixed momentum are the spin 1/2 particles denoted as usa (p) and the spin
1/2 anti-particles denoted as vas (p). Here s is the helicity index running over
s = −1/2, +1/2 and a = 1, 2 is the spin index. The field representation is given
by Z
ψa (x) = dpe ˜ −ip·x as (p)v s (p) + e−ip·x b† (p)v s (p) . (6.30)
a s a

Wigner’s trick works as before: even though the individual spinors usa (p) and
vas (p) transform non-unitarily, the creation/annihilation operators transform as
0 0
U (Λ)−1 as (p)U (Λ) = Λss as0 (Λ−1 p), U (Λ)−1 b†s (p)U (Λ) = Λss b†s0 (Λ−1 p) ,

where Λji is a rotation generator and it is unitary. Therefore, the spinor one-
particle states transform unitarily under the Lorentz group action.
Chapter 7

Spinors

7.1 Weyl and Dirac Lagrangians

In this chapter we analyse in detail the first non-trivial representation of


the Poincare group, namely the spinor representations, and construct the corre-
sponding quantum field theory. As we motivated in section 6.0.5 this representa-
tion, that is necessarily in terms of fields will be induced by the finite dimensional
spinor representation of the Lorentz algebra that we discussed in section 6.0.3.
In that section we have seen that there are two such spinors, namely the left
and right Weyl spinor representations (1/2, 0) and (0, 1/2). These are both 2
dimensional objects which we denote by ψaL and ψaR , a = 1, 2 respectively. Thus
the Lorentz transformation acting on them should be represented by a two di-
mensional matrix. Using (6.12) and (6.16) a generic Lorentz transformation
is

L(Λ) = exp [i(ηi Ki − θi Ji )] = exp −(iθi + ηi )JiL + (−iθi + ηi )JiR . (7.1)


 

The left and right Weyl representation correspond to the two dimensional rep-
resentations of the SU(2) x SU(2) algebra (6.17) in terms of Pauli’s σ matrices
on one of the SU(2)s and trivially in the other:

1
(JiL )ba ψbL = (σi )ba ψbL , (JiR )ba ψbL = 0 , (7.2)
2
1
(JiR )ba ψbR = (σi )ba ψbR , (JiL )ba ψbR = 0 . (7.3)
2
Expanding (7.1) for infinitesimal transformations we find the following infinites-
imal transformations on the fields
1 1
δψ L (x) = − (iθi +ηi )σi ψ L (Λ−1 x), δψ R (x) = (−iθi +ηi )σi ψ R (Λ−1 x) (7.4)
2 2
What is the simplest Lagrangian for these fields? We need to require that this
Lagrangian

61
62 CHAPTER 7. SPINORS

• is a Lorentz scalar ,
• is real ,
• contains a single derivative in the kinetic term.

The last requirement is based on our discussion in section 1.3.2. The second
requirement is non-trivial unlike in the case of the real scalar field, because —
as clear from the transformation properties (7.4) — components of ψ L,R are in
general complex numbers. The first requirement, Lorentz invariance, however
is the most powerful one in constructing Lagrangians.

First, consider the kinetic term for ψ L . To construct a Lorentz scalar that
contains a single derivative, one needs to contract ∂µ with another object that
contains a space-time derivative. One almost have such an object for the spinor
fields, that is the collection of the sigma matrices σ i . In fact, the following
combinations
σ µ = (1, ~σ ), σ µ = (1, −~σ ) , (7.5)
does the job, and the simplest kinetic terms for the Weyl spinors which satisfy
all the requirements above can be written as,
LR = iψ R† σ µ ∂µ ψ R , LL = iψ L† σ µ ∂µ ψ L . (7.6)
You should check explicitly, using (7.4) above, that they transform as a scalar
under Lorentz transformations.

Now let us consider a mass term. The simplest possibility mψ L† ψ L is not


Lorentz invariant. In fact the only Lorentz invariant quadratic combination of
the fields ψ L,R with no derivatives, is the one that couples the left and the right
spinors
Lm = m(ψ L† ψ R + ψ R† ψ L ) . (7.7)
This is called the Dirac mass term1 . We learn that a generic massive spinor
needs to transform in the combined left-right Weyl representation (1/2, 0) ⊕
(0, 1/2) which is of course reducible to left and right spinors. This representation
is called the Dirac spinor. It is conventional to represent it by a 4-component
spinor by writing  L 
ψ
ψ= . (7.8)
ψR
One can also combine the matrices (7.5) to form the 4x4 gamma matrices
0 σµ
 
µ
γ = (7.9)
σ̄ µ 0
The Lagrangian for the Dirac spinor is obtained by combining the left and
right components of the Weyl kinetic terms (7.6) and the Dirac mass (7.7) and
compactly written as
L = ψ̄(iγ µ ∂µ − m)ψ . (7.10)
where we also defined
ψ̄ ≡ ψ † γ 0 . (7.11)
1 It
is in fact possible to construct a mass term only for the left or the right spinors with help
of an additional “reality” condition. This is called the Majarona mass term and is discussed
in exercise XXX
7.2. DIRAC ALGEBRA AND LORENTZ TRANSFORMATIONS 63

7.2 Dirac algebra and Lorentz transformations

Before we discuss the solutions to the Dirac lagrangian and discuss the
quantization of spinors, it is important to note some properties of the Dirac
gamma-matrices and the spinors. As we already noted in section 1.3.2, the γ
matrices satisfy the Dirac algebra

{γ µ , γ ν } = −2η µν . (7.12)

The Lorentz generators in the Dirac representation are given by2



i
M µν ≡ S µν = [γ µ , γ ν ] .

(7.13)
Dirac 4

You can easily check that they reduce to (7.4) for the left and the right compo-
nents. It is also straightforward to verify that Sµν satisfy the Lorentz algebra
using (7.12). Generators of rotations are given by the spatial components

1 ijk σ k 0
 
ij
S =  (7.14)
2 0 σk

as expected. For example the upper(lower) components of both ψ L and ψ R


have spin +1/2 (-1/2) in the z-direction. The boost generators are given by

i σi
 
i0 0
S = (7.15)
2 0 −σ i

Using the hermiticity properties of the γ matrices (7.9)

γ i† = −γ i , γ 0† = +γ i , (7.16)

or directly from (7.14) and (7.15) that one finds that the spin generators are
hermitean whereas the boost generators are anti-hermitean:

S ij† = S ij , S i0† = −S i0 . (7.17)

This means that, as we alluded to in the previous sections, the boost trans-
formations on the Weyl and Dirac spinors are non-unitary. Because of this
non-unitarity, the simplest expression you would guess for a mass term i.e. ψ † ψ
is in fact not a Lorentz scalar:

ψ † ψ → ψ † exp iδωi0 S i0 − iδωij S ij exp iδωi0 S i0 + iδωij S ij ψ 6= ψ † ψ .


   

The boosts screw up the Lorentz invariance. However, noting that γ 0 satisfies

S i0 γ 0 = −γ 0 S i0 , S ii γ 0 = +γ 0 S ij ,

the desired minus sign can be generated in the combination ψ̄ψ where ψ̄ is
defined in (7.11). Indeed

ψ̄ψ = ψ † γ 0 ψ → ψ † γ 0 exp −iδωi0 S i0 − iδωij S ij exp iδωi0 S i0 + iδωij S ij ψ = ψ̄ψ .


   

2 With a slight abuse of notation, I will denote the components of the γ matrices also by

the index a, b, · · · which run from 1 to 4 now.


64 CHAPTER 7. SPINORS

Therefore the combination ψ̄ is essential in obtaining expressions that have well-


defined Lorentz transformation properties. For instance, the combination
ψ̄γ µ1 · · · γ µn ψ
transforms as a contravariant n-tensor. In particular ψ̄γ µ ∂µ ψ is a Lorentz scalar.
Thus the Dirac lagrangian in (7.10) is indeed the simplest Lorentz scalar that
can be constructed from ψ with both a kinetic and a mass term.

7.3 Chirality and helicity

The fact that the Dirac representation of the Lorentz algebra (7.13) is re-
ducible is directly seen by the presence of an operator that commutes with all
generators. This generator is called the chirality and given by
γ 5 ≡ iγ 0 γ 1 γ 2 γ 3 . (7.18)
Using the definitions (7.9) one can write it as,
 
5 −1 0
γ = (7.19)
0 1
Therefore it distinguishes the left and the right handed components in
 L 
ψ
ψ= ,
ψR

by assigning the eigenvalue -1(+1) to ψ L (ψ R ):


γ 5 ψ L = −ψ L , γ 5 ψ R = +ψ R . (7.20)
It is immediate to check that
{γ 5 , γ µ } = 0 , ⇒ [Sµν , γ 5 ] = 0 . (7.21)
Thus every state ψ in the Dirac representation should further be characterised by
its chirality eigenvalue, which means that the Dirac representation is reducible
to its left and right handed Weyl components. Even though chirality does not
change under Lorentz transformations, it is not necessarily conserved under time
evolution. For example a Dirac mass term in the Hamiltonian (7.7) transforms
a left-handed Weyl spinor into a right-handed one vice versa, thus chirality is
not conserved in the presence of a mass term. In the case m = 0it is conserved
and, as we demonstrate below, it coincides with the helicity defined in (6.25).

7.4 Solutions to the Dirac Lagrangian

The gamma matrix is written in the form:


 µ 
µ 0 σ2×2
γ4×4 = µ
σ̄2×2 0
7.4. SOLUTIONS TO THE DIRAC LAGRANGIAN 65

where σ µ = (1̂, σ i ) and σ̄ µ = (1̂, −σ i ). We can find the equation of motion as:

(i∅ − m)ψ = 0.

where ∅ = ∂µ γ µ . Then, we have:

(∂ 2 − m2 )ψ = 0.

The solutions to this equation


p are given as the momentum eigenvectors e±ipx
0
with the energy p = ± |~ p| + m2 . Here, the negative energy solutions corre-
2

spond to antiparticles. For particles


Z
ψs = dpUS (p)eip·x

Here we have
√ 
−p · σξ S
Us (p) = √
−p · σ̄ξS

where ξS = (1, 0) corresponds to spin up and ξS = (0, 1) corresponds to spin


down.

For antiparticles, we have:


Z
ψS = dpVS (p)e−ipx

Here we have:
√ 
√−p · σξS
Vs (p) =
− −p · σ̄ξS

p With energy w, momentum only in z-direction and m = 0, we have w =


w2 + p2z = ±pz . Choose pz = w. Then, we find the product

−p · σ = −ηνµ pµ σ ν
= w(1, 0; 0, 1) ± w(1, 0; 0, −1)
   
2w 0 0 0
= ,
0 0 0 2w

Then, we find the matrices


       
0 √0 0 √0
 0   2w  0   2w
U↑ = √  U↓ =  V↑ =  √  V↓ = 
 0 .
 
 2w  0  − 2w
0 0 0 0
66 CHAPTER 7. SPINORS

7.5 Chirality and Helicity

7.5.1 Chirality

Chirality gives information about whether we have right or left spinor. We


define the operator

γ 5 = iγ 0 γ 1 γ 2 γ 3

where we can show that {γ 5 , γ µ }=0. γ 5 commutes with all Lorentz generators
Sµν = 4i [γ µ , γ 0 ]: [Sµν , γ 5 ] = 0. Thus
 
5 −1̂ 0
γ =
0 1̂

Chirality does not change under Lorentz transformation. But, it is not


† †
necessarily conserved. For example, mγ̄γ = m(ψR ψL + ψL ψR ). If m = 0,
chirality is conserved. So, chirality is defined through the action of γ 5 .

7.5.2 Helicity

~
Define h = p~|~p·S| , spin projected onto momentum. For U↑ , h = 12 , γ 5 = 1 and
U↓ , h = − 12 , γ 5 = −1. Helicity coincides with chirality for m = 0.

~ are conserved.
Helicity is conserved as p~ and S

7.6 Spin and Statistics

For integer spin, the particles follow boson statistics. For half-integer spin,
the particles follow fermion statistics.

Consider a spinor and rotate the frame with an angle θ. We know that:
i µ ν
Sµν = [γ , γ ].
4
then

L1/2 (θ) = exp(iθS12 ).

where we have
1
S12 = diag(1, −1, 1, −1).
2
Thus, we have L1/2 (2π) = −1diag(1, 1, 1, 1). So, we can only obtain the identity
for L1/2 (4π) = 1diag(1, 1, 1, 1).
7.7. CANONICAL QUANTIZATION OF SPINORS 67

For example, assume that there are two spinors at x and −x. Their wave-
function can be given as ψ1/2 = |ψ↑ (x)ψ↑ (−x)i. Consider a rotation along the
z-axis by θ = π. L1/2 (π) = i diag(1, −1, 1, −1). Then we have |ψ↑ (x)ψ↑ (−x)i =
− |ψ↑ (−x)ψ↑ (x)i. This shows the intiutive reason behind fermi statistics.

7.7 Canonical Quantization of Spinors

Start with the Langrangian:

L = iψ̄∅ψ − mψ̄ψ

Take the canonical fields as ψ. Then, canonical momentum is πψ = δ∂δL 0ψ


=
0
iψ̄γ . Now, we can start the quantization with the following anti-commutation
relations:

{ψα (~x, t), ψβ (~y , t)} = 0 (7.22a)


3
{ψα (~x, t), πψ−β (~y , t)} = iδ (~x − ~y )δαβ (7.22b)
{πψ−α (~x, t), πψ−β (~y , t)} = 0 (7.22c)

Using these, we can determine the anticommutation relations between a and


b’s. As shown in the notes, we find the results
0
{aSp , aSp0 † } = (2π)3 δ 3 (~
p − p~0 )δSS 0 2w, (7.23a)
0
{bSp , bSp0 † } 3 3 0
p − p~ )δSS 0 2w.
= (2π) δ (~ (7.23b)
p
where w = p0 = p|2 + m2 and all the rest vanish.
|~

For spinors, we found the two solutions ψS and χS . Then, let us write the
classical solution
XZ
ψ= dp(eip·x US (p)aS + e−ip·x VS (p)b∗S )
S

then, through cannocial quantization, we have the quantized field as


XZ
ψ= dp(eip·x US (p)âS (p) + e−ip·x VS (p)b̂†S (p))
S

γ̄ then becomes
XZ
ψ̄ = dp(e−ip·x ŪS (p)â†S (p) + eip·x V̄S (p)b̂S (p))
S

Important difference between a real scalar is that instead of ap and a†p , we


have aSp , aS† S S† S
p , bp , bp . Here ap destroys a particle with momentum p and spin
S, bSp destroys an anti-particle.
68 CHAPTER 7. SPINORS

There are no finite dimensional unitary irreducable representation (irrep)


of a noncompact group such as the Lorentz. Wigner in 1939: Find infinite
dimensional unitary irreps starting from the finite non-unitary irreps. In Dirac’s
case, US is finite by non-unitary. Therefore, U † (p)U (p) is not conserved under
Lorentz transformation. Nonetheless, ψ being infinite dimensional fixes this
problem.

7.8 The One-particle State

p, si = as†
We define the individual components of Dirac solution as |~ p |0i,
where we
p call this one-particle state with momentum p
~ and spin ~
s . Its energy
is w = |~p|2 + m2 .

Consider hp, s|p, si. The probability should be Lorentz invariant. |p, si =
U (Λ) |p, si, with U U † = 1. Consider

XZ
0 0
XZ d3 p d3 p0 0
dp̃ dp̃0 hp , s |p, si = h0| aSp0 aS†
p |0i
(2π)3 2w (2π)3 2w
S,S 0 S,S 0
XZ d3 p d3 p0 0
= 3 3
h0| {aSp0 aS†
p } |0i
(2π) 2w (2π) 2w
S,S 0
XZ d3 p d3 p0
= p − p~0 )2wδSS 0
h0|0i (2π)3 δ 3 (~
(2π) 2w (2π)3 2w
3
S,S 0
XZ d3 p
= .
(2π)3 2p0
S

d3 p
where we note dp̃ = (2π)3 2w

7.9 Quantization of Spinor Fields

We can represent the spinors as left moving and right moving (jL , jR ) =
(1/2, 0) or (0, 1/2), where we define them as left/right Weyl spinors. In Dirac
representation, we write them together as (1/2, 0) ⊗ (0, 1/2). The Langrangian
for a Dirac spinor can be written as:

LDirac = iψ̄∅ψ − mψ̄ψ

where ψ̄ = ψ † γ 0 . Solution to the EOM is then:


XZ
ψ= ˜ ip·x US (p)aS + e−ip·x VS (p)b∗ )
dp(e S
S

p
Here, it is also important to note that p0 = w = p2 | + m 2 .
|~
7.9. QUANTIZATION OF SPINOR FIELDS 69

To quantize, we start by noting that πψ = δ∂δL 0ψ


. Then, with the commuta-
tion relations we defined, we end up getting:
XZ
ψ̂ = ˜ ip·x US (p)âS (p) + e−ip·x VS (p)b̂† (p))
dp(e S
S

Wigner: There are no finite dimensional unitary irreducable representations


for a non-compact group such as Lorentz group.

The representation of a Lorentz group can be written as L1/2 (Λ) = eηi Ki +iθi Ji ,
where boosts screw up unitarity. One particle states transforms unitarily under
boosts: |p, si = a†S (p) |0i. Let us start writing:

hp0 , s0 |p, si = h0| aS 0 d (p0 )a†S (p) |0i


p − p~0 )δSS 0 .
= h0|0i (2π)3 2wδ 3 (~

where by definition h0|0i = 1. Here, δ̃(~p − p~0 ) = (2π)3 2wδ 3 (~


p − p~0 ) is a Lorentz
invariant delta-function. Lorentz transformations leave the norm invariant as:
Z Z
˜ 0
p − p~ ) = d3 pδ(~
dpδ̃(~ p − p~0 ) = 1.

The lorentz transformation of a creation operator then should be:

a†S (p) → U (Λ)a†S (p)U −1 (Λ) = a†Λ1/2 S (Λp).

as when changing δ̃ → δ˜Λ , we set w → w0 and δ(~ p − ~q) → δ(Λ~ p − Λ~q). So, once
again:

|p, si = a†S (p) |0i → Λp, Λ1/2 s = a†Λ1/2 S (Λp) |0i


where |0i → |0i is assumed to be Lorentz invariant. Here, as a†S (p) → a†Λ1/2 S (Λp)
ensures that this representation is indeed a unitary representation. Then:
XZ
ψ̂ → U ψU −1 = ˜ ip·x US (p)âΛ S (Λp) + e−ip·x VS (p)b̂†
dp(e 1/2 Λ1/2 S (Λp))
S

where U a†S (p)U −1 = aΛ1/2 S (Λp). Now, we can change variables p = Λ−1 p0 .
Then, we have
XZ
ψ̂ → U ψU =−1 ˜ iΛ−1 p0 ·x US (Λ−1 p0 )âΛ S (p0 ) + ...)
dp(e 1/2
S

The spinor then transforms as:

US (Λ−1 p0 ) → L(Λ−1 )US (p)


US (p) → L(Λ)US (p)
ψ̂(x) → L(Λ)ψ(Λ−1 x)

which is indeed as expected.


Chapter 8

Vector fields and the gauge


symmetry

(jL , jR ) = (1/2, 1/2). Dimension (2jl + 1) × (2jR + 1) = 4. Then, we


define the vector field Aµ , where µ = 0, 1, 2, 3 and A†µ = Aµ . What is the
Langrangian for a vector field Aµ (x)? There will be a kinetic term, mass term
and an interaction term. The action should also be Lorentz invariant and real.

Let us try:
1 1
L = − ∂µ Aν ∂ µ Aν − m2 Aν Aν .
2 2
Here, the Langrangian is indeed Lorentz invariant due to contraction over all
indices. The EOM for Aµ becomes ( − m2 )Aµ = 0. Here, we realize that
there are no cross terms between Aν and Aµ . Thus, the EOM constitute four
decoupled Klein-Gordon fields. So we found 4 × (0, 0) representation not the
vector representation. We need another Langrangian, a more general one. Such
a Langrangian can be written as:
a b 1
L= ∂µ Aν ∂ µ Aν + ∂µ Aν ∂ ν Aµ − m2 Aν Aν
2 2 2
where Aν ∂ ν imposes the vector representation and the cross term.

A serious problem with 4 scalar Langrangian is that it contains negative


norm states upon quantization.
Z
Av = dp(e˜ ipx aν (p) + e−ipx a† (p))
ν

Commutation relations [aµ (~ p), a†ν (~


p0 )] = ηνµ δ 3 (~
p − p~0 )2w. For η00 = −1, let us
† 0 3 0
check a0 field. [a0 (~ p )] = −δ (~
p), a0 (~ p − p~ )2w. This then results in negative
norm states.

Gauge invariance will remove the negative norm states from the Hilbert

70
8.1. PROCA LAGRANGIAN: A = −1 71

space. Negative norm states could arise for any field with µ-index. Same prob-
lem for the graviton. Back to vector representation then.

Equations of motion for the vector representation can be given as:


−aAv − b∂ν ∂µ Aµ − m2 Aν = 0
Multiply both sides with ∂ ν .
− (a + b) + m2 ∂ν Aν = 0.


Assume m2 6= 0, if a = −b then m2 ∂νp


Aν = 0. Constraint equation. In the
µ
momentum space, Pµ A (p) = 0. Set w = |~p|2 + m2 , we have A0 = −pi Ai /w.

Aµ = (1/2, 1/2). Spin states 0, 1. A constraint can remove the spin-0 state.

8.1 Proca Lagrangian: a = −1

The Proca Lagrangian can be written as:


1 1 1
LP roca = − ∂µ Aν ∂ µ Aν + ∂µ Aν ∂ ν Aµ − m2 Aν Aν .
2 2 2

Define the ”field strength” Fµν . Then:


Fµν = ∂µ Aν − ∂ν Aµ
which allows us to write the Langrangian as:
1 m2
LP roca = − Fµν F µν − Av Av .
4 2

Remark: Defining Gauge transformation as Av (x) → Av (x) + ∂ν Λ(x), then


Fµν is gauge invariant. Lproca is only gauge invariant when m2 = 0.

8.2 Vector Fields: Recap

(0, 0) → KC − f ield, (1/2, 0), (0, 1/2) → W eyl and Dirac fields (1/2, 1/2)
vector field. For vector fields m2 = 0 or m2 ≥ 0. For m2 ≥ 0, we will deal with
the Proca Lagrangian.
1
L = − Fµν F µν − m2 Aµ Aµ ,
4
where Fµν = ∂µ Aν − ∂ν Aµ is called the field strength. Here, we demand a
constrant ∂µ Aµ = 0. The EOM becomes ∂µ F µν = m2 Aν =⇒ ∂ν Aν = 0.
For Aµ = (A⊥ , Ak ), by setting the constraint we get rid of Ak and get A0 =
√ ki Ai . The wave equation is given as
k k +m2
i i

( − m2 )Aν = 0,
where p2 = −m2 .
72 CHAPTER 8. VECTOR FIELDS AND THE GAUGE SYMMETRY

8.3 Quantum Vector Field

Z
(eipx âi (p)µi (p)∗ + e−ipx â†i (p)µi (p))
X
µ (x) = ˜
dp
i

where i is a label for polarization and µi are the solutions to the EOM. eipx µ∗
i
satisfies the K-G equation. We should also demand the constraint

pµ µ∗
i (p) = 0, ∀i.

We can find the solutions to this


p constraint. Choose an inertial frame such
that pµ = (w, 0, 0, pz ) with w = (pz )2 + m2 . Find all linearly independent
solutions to pµ µ∗ µ
i (p) = 0. Two such solutions are 1 = (0, a, 0, 0) and 2 =
µ

(0, 0, b, 0). We may also demand normalization. i (p)j (p) = δij , ∀i, j. These
are called transverse polarizations. For the third solution, we may try µ3 =
(pz /m, 0, 0, w/m). This is called the longitudinal polarization. Then, the most
general solution is
Z 3
(eipx âi (p)µi (p)∗ + e−ipx â†i (p)µi (p))
X
µ
 (x) = ˜
dp
i=1

Here eipx µ∗i is an infinite dimensional representation. One-particle states:


|p, ii = a†i (p) |0i. Commutation relations correspond to unitary representations.

For (1/2, 1/2) =⇒ m2 > 0: 4 = 1 + 3, where the 1 dimension is forbidden


by the constraint equation. If m2 = 0, 2 dimensional representation. If m2 = 0,
then we have
1
L = − Fµν F µν .
4
where EOM ∂µ F µν = 0 and ∂κ Fµν +∂ν Fκµ +∂µ Fνκ = 0. This leads to Maxwell’s
~ Then define A0 = V the scalar potential and A
equation. Set Aµ = (A0 , A). ~
as the vector potential. This gives us the physics behind the electrodynamics,
where

E i = ∂ 0 Ai − ∂i A0 (8.1a)
Bi = ijk ∂j Ak (8.1b)

where Fij = ijk Bk . We may then check the Lagrangian

1
L = − (2F0i F 0i + Fij F ij )
4
1
= − (2E i E i η00 + ijm ijm Bm B m )
4
1 2
= (E − B 2 )
2
8.4. GAUGE INVARIANCE 73

Before Aµ satisfied the constraint equation. Therefore ∂ν ∂µ F µν = 0 satisfied


trivially. If there was a source, then ∂ν ∂µ F µν = ∂ν J ν = 0 which is indeed
the continuity equation. Here, there is no constraint on Aν . There are a few
problems here

• No decomposition into spin 0 + spin 1 as in the Proca case


• A0 component will lead to negative norm states
• Longitidunal polarization µ3 = (pz /m, 0, 0, w/m) → ∞.

8.4 Gauge Invariance

Gauge invariance solves all the problem. Assume L = − 41 Fµν F µν is invari-


ant under Aµ → Aµ + ∂µ α for an arbitrary α(x). Note that the mass term
breaks the gauge invariance.

We can use the gauge invariance to solve the issues.

We can set A0 = 0 (temporal gauge). A0 → A0 + ∂ 0 α, so that there are no


negative norm states.

We can also set ∂µ Aµ = 0, which manifestly satisfies Lorentz invariant.


What about the negative norm states?

Finally, there is also a coulomb gauge ∂i Ai = 0. We want ∇ · A~ = 0.So,


~ 2
choose α such that ∇A = 0. Choose such that ∇ α = 0, then this still satisfies
the gauge condition. The resudial gauge freedom or ”leftover” gauge freedom:
All functions α(x) that satisfy ∇2 α = 0.

Consider ∂µ F µν = ∂µ (∂ µ Aν − ∂ ν Aµ ) = 0. For v = 0, grad2 A0 = 0. A0


satisfies the same equation as the leftover gauge freedom. Then, we can choose
α such that A0 → 0.
~ = 0. Then,
Overall, we choose α with the gauge freedom such that ∇ · A
we choose left-over gauge freedom to set A0 = 0.

For v = i: We realize that Ai = 0, which is indeed the Klein-Gordon


equation. We then found the conditions ∂i Ai = 0, Ai = 0 and A0 = 0. We
need to find their solutions.
Z 2
(eipx âi (p)µi (p)∗ + e−ipx â†i (p)µi (p))
X
µ
 (x) = dp ˜
i=1
2
where p = 0, i0 (p) = 0 and pµ iµ (p) = 0.

Choose a frame: pµ = (pz , 0, 0, pz ). Then, there are two possible solutions


µ1 = (0, 1, 0, 0) and µ2 = (0, 0, 1, 0). These are transverse polarizations. We
could also choose µR = √12 (0, 1, i, 0) and µL = √12 (0, 1, −i, 0). These are called
circular polarizations. These solutions are helicity eigenstates.
74 CHAPTER 8. VECTOR FIELDS AND THE GAUGE SYMMETRY

8.5 Quantum Gauge Field: Canonical Quanti-


zation

Commutation Relations:
1 1 1 1
L = − Fµν F µν = (∂t Ai )2 − ∂j Ai ∂j Ai + ∂i Aj ∂j Ai
4 2 2 2
where ∂i Aj ∂j Ai = 0 by gauge condition. Recall that for columb gauge ∂i Ai = 0
1 1 1
L = − Fµν F µν = (∂t Ai )2 − ∂j Ai ∂j Ai
4 2 2

Canonical fields: Ai . Conjugate canonical momenta:

i δL
πA = = ∂t Ai .
δ(∂t Ai )

Hamiltonian:
i
H = πA ∂t Ai − L
1 i 2 1
= (πA ) + (∇Ai )2 .
2 2

Commutation relation

[ΠiA (~x, t), Aj (~y , t)] = iδ ij δ 3 (~x − ~y )

does not satisfy the constraint ∂i Ai = 0. Define


 
∂i ∂j ∂i ∂j
Ai (x) = δij − 2 Aj + Ai
∇ ∇2
In momentum space
ki kj ki kj
Ai (k) = (δij − )Aj + Ai
~
|k| 2 |~k|2

To obtain comutation relations satisfied by the constrained field project


ki kj
Ai → Ai⊥ = (δij − |~k|2
)Aj . Then, the commutation relation is

∂i∂j 3
[ΠiA (~x, t), Aj (~y , t)] = i(δ ij − )δ (~x − ~y )
∇2
d3 k i~k·(~x−~y) ij k i k j
Z
=i e (δ − )
(2π)3 |~k|2
where the vector field is
Z 2
(eikx ai (k)ji (k)∗ + e−ikx a†i (k)ji (k))
X
Aj (x) = dk ˜
i=1
8.5. QUANTUM GAUGE FIELD: CANONICAL QUANTIZATION 75

Here, the commutation relations for the creation and annihilation operators
become

[aik , ajk0 ] = 0
[ai† j†
k , ak0 ] = 0

[aik , aj† 3 3 ~ ~ 0 ij
k0 ] = (2π) δ (k − k )δ .

E
For a photon ai†
~
k |0i = k, i . Gauge invariance of observables such as
D E
~k 0 , j ~k, i in general guaranteed by the ”Ward identity”.
Chapter 9

Quantum electrodynamics

76

You might also like