Lecture Notes QFT Temp 3
Lecture Notes QFT Temp 3
Umut Gürsoy
1
2 CONTENTS
4 Interactions in QFT 39
4.3.2 Counterterms . . . . . . . . . . . . . . . . . . . . . . . . . 47
7 Spinors 61
7.1.2 Chirality . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
7.2.1 Chirality . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
7.2.2 Helicity . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
9 Quantum electrodynamics 76
Preliminaries
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 %
• Srednicki, M.
• Schwartz, M
and my own notes. There are other excellent material you can benefit from:
• Peskin, Schröder
• A. Zee
• Weinberg
• 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)
• Vector fields
• Path integrals
• Correlation functions
• Scattering
• Gauge invariance and QED
Chapter 1
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.
(i) State of the system at time t is represented by a vector in the Hilbert space
|ψ, ti ∈ H.
x0 = −x0 , xi = xi .
More generally, we define the metric tensor to raise and lower indices:
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.
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
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
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.
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)
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
The solution is
Now the equation in the position basis can be written succinctly as,
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.
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.
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
x y
σ
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
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
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.
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
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 + 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?
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).
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
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).
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|)
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)
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
to commutation relations3 :
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.
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 :
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 ).
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)
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
1. The need to pass from a single particles to fields to retain both Lorentz
invariance and unitarity.
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.
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
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.
Some comments:
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
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 .
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
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
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.
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 .
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
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 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)
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
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.
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
Figure 3.3: Definition of the contour C̃ in the complex frequency space that
defines the vacuum-to-vacuum amplitude.
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
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)
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:
where the sum is over all different pairs. The result (3.37) and (3.38) is called
the Wick’s theorem.
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 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
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
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.
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
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.
∞ 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.
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.
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.
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.
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
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
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}
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
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.
4.3.2 Counterterms
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.
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!?
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
Now we will answer the question how to fix the counterterms unambiguously.
Chapter 5
50
Chapter 6
Representations of the
Lorentz group and particles
51
52CHAPTER 6. REPRESENTATIONS OF THE LORENTZ GROUP AND PARTICLES
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
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
[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.
Ji = JiL + JiR
• 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
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.
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
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
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.
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 :
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
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
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.
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)
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
i σi
i0 0
S = (7.15)
2 0 −σ i
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:
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:
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
2 With a slight abuse of notation, I will denote the components of the γ matrices also by
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
where σ µ = (1̂, σ i ) and σ̄ µ = (1̂, −σ i ). We can find the equation of motion as:
(i∅ − m)ψ = 0.
(∂ 2 − m2 )ψ = 0.
Here we have
√
−p · σξ S
Us (p) = √
−p · σ̄ξS
Here we have:
√
√−p · σξS
Vs (p) =
− −p · σ̄ξS
−p · σ = −ηνµ pµ σ ν
= w(1, 0; 0, 1) ± w(1, 0; 0, −1)
2w 0 0 0
= ,
0 0 0 2w
7.5.1 Chirality
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
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
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
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 thatthere 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.
L = iψ̄∅ψ − mψ̄ψ
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
γ̄ then becomes
XZ
ψ̄ = dp(e−ip·x ŪS (p)â†S (p) + eip·x V̄S (p)b̂S (p))
S
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
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:
p
Here, it is also important to note that p0 = w = p2 | + m 2 .
|~
7.9. QUANTIZATION OF SPINOR FIELDS 69
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:
as when changing δ̃ → δ˜Λ , we set w → w0 and δ(~ p − ~q) → δ(Λ~ p − Λ~q). So, once
again:
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
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.
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.
Aµ = (1/2, 1/2). Spin states 0, 1. A constraint can remove the spin-0 state.
(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
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.
E i = ∂ 0 Ai − ∂i A0 (8.1a)
Bi = ijk ∂j Ak (8.1b)
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
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
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
∂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