QFT Lecture Note
QFT Lecture Note
Lecture Notes
Joachim Kopp
3
Contents
5 Quantum Electrodynamics 81
5.1 The QED Lagrangian from Symmetry Arguments . . . . . . . . . . . . . . 81
5.2 The Feynman Rules for QED . . . . . . . . . . . . . . . . . . . . . . . . . 83
5.3 e+ e− → µ+ µ− . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
5.3.1 Feynman Diagram and Squared Matrix Element . . . . . . . . . . 86
5.3.2 Trace Technology . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
5.3.3 The Squared Matrix Element for e+ e− → µ+ µ− (Part II) . . . . . 89
5.3.4 The Cross Section — General Results . . . . . . . . . . . . . . . . 90
5.3.5 The Cross Section for e+ e− → µ+ µ− . . . . . . . . . . . . . . . . . 91
5.3.6 e+ e− → µ+ µ− : Summary . . . . . . . . . . . . . . . . . . . . . . . 92
5.4 More Technology for Evaluating QED Feynman Diagrams . . . . . . . . . 93
5.4.1 Scattering of Polarized Particles . . . . . . . . . . . . . . . . . . . 93
5.4.2 External Photons . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
6 Path Integrals 95
6.1 Path Integrals in Quantum Mechanics . . . . . . . . . . . . . . . . . . . . 95
6.2 The Path Integral for a Free Scalar Field . . . . . . . . . . . . . . . . . . . 97
6.3 The Feynman Propagator from the Path Integral . . . . . . . . . . . . . . 100
6.4 Wick’s Theorem from the Path Integral . . . . . . . . . . . . . . . . . . . 101
4
Contents
Bibliography 139
5
Contents
6
Introduction and Motivation
1
1.1 Do you recognize the following equations?
∂t2 ψ − ∇2 ψ − m2 ψ = 0 (1.3)
i∂/µ ψ − mψ = 0 (1.4)
δL(φ, ∂µ φ) δL(φ, ∂µ φ)
∂µ − =0 (1.5)
δ(∂µ φ) δφ
dσ
= |f (θ, φ)|2 (1.6)
dΩ
δL(φ, ∂µ φ)
jµ = ∆φ − J µ (1.7)
δ(∂µ φ)
~ = c = 1. (1.8)
7
Chapter 1 Introduction and Motivation
which is also employed by Peskin and Schroeder [1], and by most particle physicists
today. Note that the book by Srednicki, for instance, uses the “East Coast Metric”, with
−1 in the timelike component and +1 in the spacelike components [2].
8
The Klein–Gordon Field
2
2.1 Necessity of the Field Viewpoint
From relativistic quantum mechanics, we know how to deal with the dynamics of a single
relativistic particle: the Klein-Gordon equation
describes the time evolution of the wave function of a relativistic scalar particle in vac-
uum. (Here, x = (t, x) is the coordinate 4-vector and m is the mass of the particle.)
The Klein–Gordon equation can be easily generalized to motions in an external field by
making the replacements pµ → pµ + ieAµ . The Dirac equation
achieves the same for a fermionic particle. However, these equations only describe the
motion of a single particle. The relativistic equivalence of mass and energy tells us
that kinetic energy can be converted into the production of new particle (e.g. e+ e−
pair production when an ultrarelativistic electron travels through matter), therefore it
is clear that a complete description of a particles’ dynamics must provide for creation
and annihilation processes as well. Note that even in processes where too little energy
is available to actually create an e+ e− pair, the intermittent existence of such pairs for
very short time intervals is allowed by the Heisenberg uncertainty relation ∆E · ∆t ≥ ~
and will therefore happen.
Therefore a relativistic theory that can describe multiparticle processes, including
in particular particle creation and annihilation, is highly desirable. Actually, we have
encountered creation and annihilation processes before: when studying the quantum
mechanical harmonic oscillator, the operators a and a† that transformed its wave function
to a lower or higher energy state were called creation and annihilation operators, though
9
Chapter 2 The Klein–Gordon Field
at the time this seemed like a bit of a misnomer. Quantum field theory (QFT) generalizes
the concept of the harmonic oscillator to an extent that makes the terms “creation”
and “annihilation” operator appropriate. It describes particles of a given species (e.g.
electrons) as a field, i.e. a function φ(x) that maps each spacetime point x to a scalar value
(scalar fields, e.g. the Higgs boson), a Dirac spinor (fermion fields, e.g. the electron) or a
Lorentz 4-vector (vector fields, e.g. the photon). We will show below that the equation
of motion for each Fourier mode of this field has exactly the same form as the equation
of motion of the harmonic oscillator. It can therefore be quantized in exactly the same
way. The creation and annihilation operators then become ap and a†p , e.g. they carry a
momentum index p corresponding to the Fourier mode they describe. The interpretation
of a field is then the following: its ground state, where each momentum mode carries
only the zero point energy, corresponds to the vacuum |0i. If a momentum mode is in
its first excited state a†p |0i, this means that one particle with momentum p exists. The
operators a†p and ap then create and annihilate particles. The equations of motion of
the theory describe the rules according to which such creation and annihilation processes
occur.
In the following, we will first review a few concepts of classical (non-quantized) field
theory (section 2.2) and then make the above statements more precise and more mathe-
matical (section 2.3).
is stationary, i.e.
δS = 0 , (2.4)
where L(x, ẋ) is the Lagrange functional and δS means the variation of S with respect
to x and ẋ. In field theory, the Lagrange functional becomes the Lagrange density (or
Lagrangian for short) L(φ, ∂µ φ), a functional of the field value and its spactime deriva-
tives. (It is called a density because we will see that it has units of [length]−3 [time]−1 , as
opposed to the Lagrange functional in classical mechanics, which has units of [time]−1 .)
The action is defined as
Z
S ≡ d4 x L(φ, ∂µ φ) . (2.5)
10
2.2 Elements of Classical Field Theory
where in the last step we have integrated by parts. Since eq. (2.7) is required to be
satisfied for any variation δφ, the term in square brackets must vanish. This leads to the
Euler-Lagrange equations
δL δL
∂µ − = 0. (2.8)
δ(∂µ φ) δφ
As an example, consider a scalar field φ (for instance the Higgs field), for which the
Lagrangian reads
1 1
L = (∂µ φ)(∂ µ φ) − m2 φ2 . (2.9)
2 2
The Euler–Lagrange equations then lead to the equation of motion
∂µ ∂ µ φ + m2 φ = 0 , (2.10)
H = π ẋ − L , (2.11)
H ≡ π(x)φ̇(x) − L , (2.13)
δL
π(x) ≡ . (2.14)
δ φ̇(x)
11
Chapter 2 The Klein–Gordon Field
L → L + α∂µ J µ . (2.16)
Thus,
δL δL
α∆L ≡ α∂µ J µ = (α∆φ) + ∂µ (α∆φ) (2.17)
δφ δ(∂µ φ)
δL δL δL
= α∂µ ∆φ + α − ∂µ ∆φ . (2.18)
δ(∂µ φ) δφ δ(∂µ φ)
In the second step, we have used the product rule of differentiation backwards. The term
in square brackets vanishes due to the Euler–Lagrange equations. Therefore, we are left
with the conclusion that the current
µ δL
j ≡ ∆φ − J µ (2.19)
δ(∂µ φ)
is conserved:
∂µ j µ = 0 . (2.20)
This implies in particular for the associated charge (which, except in the case of electro-
magnetic gauge transformations has nothing to do with electric charge)
Z
Q ≡ d3 x j 0 (2.21)
is constant:
Z Z Z
Q̇ = d x∂0 j = d x∂µ j − d3 x∂k j k = 0 .
3 0 3 µ
(2.22)
12
2.3 Quantization of the Klein–Gordon Field
Here, as usual, the greek index µ runs from 0 to 3, while the latin index k runs only from
1 to 3. The term containing ∂µ j µ vanishes by virtue of Noether’s theorem (eq. (2.20)),
the one containing ∂k j k vanishes thanks to Gauss’ theorem.
As an example, consider infinitesimal spacetime shifts
xµ → xµ − αaµ , (2.23)
for some constant 4-vector aµ , applied to the free scalar field from eq. (2.9). The corre-
sponding field transformation is
Plugging this into the Lagrangian eq. (2.9), we obtain (remembering that α is infinitesi-
mal)
This implies
J ν = aν L (2.26)
and
h i
j µ = (∂ µ φ) ∂ν φ(x) − L δ µν aν .
(2.27)
Consider in particular time-like shifts, i.e. aν = (1, 0, 0, 0). Then the term in square
brackets becomes
j 0µ ≡ (∂ µ φ) ∂0 φ(x) − L δ µ0 .
(2.28)
The 0-component of this expression is just the Legendre transform of the Lagrangian, i.e.
the Hamiltonian density. The d3 x integral over the Hamiltonian density (or Hamiltonian
for short) is just the energy. Thus, invariance under time-like shifts implies energy
conservation. Had we chosen aν to correspond to space-like shifts, we would in analogy
have found momentum conservation.
13
Chapter 2 The Klein–Gordon Field
When going from classical mechanics to quantum mechanics, the procedure is to promote
the position q and the canonical momentum π ≡ (δL)/(δ q̇) to operators that satisfy
canonical commutation relations
[q j , pk ] = iδ jk (2.30)
j k j k
[q , q ] = [p , p ] = 0 . (2.31)
We now do this analogously for the field φ(x) by promoting φ(x) and
δL
π(x) ≡ (2.32)
δ φ̇(x)
Note that the time coordinate of the spacetime points x and y is the same here.1 Since
we normally deal with particles of known energy (for instance the particles in the LHC
beams), without being too much interested in their position, it is convenient to go to
Fourier space and write
d3 p ipx
Z
φ(x, t) = e φ(p, t) , (2.34)
(2π)3
d3 p ipx
Z
π(x, t) = e π(p, t) . (2.35)
(2π)3
Note that φ(−p, t) = φ(p, t) because φ(x, t) was assumed to be real. The Klein–Gordon
equation
∂µ ∂ µ φ + m2 φ = 0 , (2.36)
then becomes
2
∂ 2 2
+ |p| + m φ(p, t) = 0 . (2.37)
∂t2
This is precisely the equation of motion for a harmonic oscillator with frequency
p
ωp = |p|2 + m2 . (2.38)
1
Allowing for different time coordinates would not make sense. For points with a spacelike separation,
causality dictates that all commutators must be zero because measurements at points that are outside
each other’s light cone cannot influence each other. Measurements at points within the light cone,
however, can affect each other. Thus, the commutator of fields at different t is more complicated.
14
2.3 Quantization of the Klein–Gordon Field
We know how to quantize the harmonic oscillator: we introduce creation and annihilation
operators a†p (t) and ap (t), defined such that
1
ap (t) + a†−p (t)
φ(p, t) = p (2.39)
2ωp
r
ωp
ap (t) − a†−p (t) .
π(p, t) = −i (2.40)
2
Note that, by convention, we assign the creation operator an index −p. It will become
clear later why this choice makes sense. Note also that we work in the Heisenberg picture
here where the field operators are time dependent. We will see below in sec. 2.4 what
the explicit form of this time dependence is. From the commutation relations eq. (2.33),
we can then derive
† 1 √ i √ 0 i 0
[ap (t), ap0 (t)] = ωp φ(p, t) + √ π(p, t), ωp φ(−p , t) − √
0 π(−p , t)
2 ωp ωp0
(2.41)
r
i ωp ωp0
r
φ(p, t), π(−p0 , t) + φ(−p0 , t), π(p, t)
=− (2.42)
2 ωp0 ωp
Z Z r
i ωp ωp0
r
3 −ipx 3 0 ip0 x 0 0
=− d xe d x e φ(x, t), π(x , t) + φ(x , t), π(x, t)
2 ωp0 ωp
(2.43)
Z r
1 ωp ωp0
r
0
= d3 x e−i(p−p )x + (2.44)
2 ωp0 ωp
= (2π)3 δ (3) (p − p0 ) . (2.45)
reads
Z
3 1 2 1 2 1 2 2
H= d x [π(x, t)] + [∇φ(x, t)] + m [φ(x, t)] (2.46)
2 2 2
√
d3 p d3 p0 i(p+p0 )x ωp ωp0
Z Z
= d3 x ap (t) − a†−p (t) ap0 (t) − a†−p0 (t)
3 3
e −
(2π) (2π) 4
0 2
−p · p + m † †
+ √ ap (t) + a−p (t) ap0 (t) + a−p0 (t) (2.47)
4 ωp ωp0
d3 p
Z
ωp a†p (t)ap (t) + 12 [ap , a†p ] .
= 3
(2.48)
(2π)
In the last line, we have used that ω−p = ωp . As for the harmonic oscillator in quantum
mechanics, the operator a†p (t)ap (t) is the particle number operator. When applied to a
15
Chapter 2 The Klein–Gordon Field
quantum state |ψi, it gives the number of particles with momentum p in that state:
a†p (t)ap (t)|ψi = np |ψi (2.49)
The term [ap , a†p ] is an infinite constant according to the commutation relation eq. (2.45)
and corresponds to the zero point energies of all the individual p-modes. However, exper-
iments measure only energy differences, therefore a constant (albeit infinite) contribution
to H can be dropped without changing the physics. This is what we will do from now
on.2
We now discuss the eigenstates of the Hamiltonian eq. (2.48). As before, we call the
vacuum state |0i. It isnormalized according to
h0|0i = 1 . (2.50)
A state containing exactly one particle of momentum p will be written as |pi ≡ ca†p |0i.
Here, ap (without the argument t) denotes ap (0). Remember that we are working in the
Heisenberg picture, so the states are time-independent and all the time-dependence is
included in the operators. The normalization constant c is chosen such that
hp|qi = 2Ep (2π)3 δ (3) (p − q) , (2.51)
p
where Ep = p2 + m2 . This normalization condition has the advantage of being Lorentz
invariant, as can be easily shown by applying a Lorentz transformation and using the
properties of the δ-function (Exercise!). It implies
|pi ≡ 2Ep a†p |0i .
p
(2.52)
This is easily seen by directly computing
hp|qi = 2 Ep Eq h0|ap a†q |0i
p
(2.53)
= 2 Ep Eq h0|a†q ap |0i + h0|[ap , a†q ]|0i
p
(2.54)
= 2Ep (2π)3 δ (3) (p − q) . (2.55)
16
2.4 The Feynman Propagator for the Klein–Gordon Field
which implies
We again use the notation ap ≡ ap (0). In a similar way we can show that
Consequently, making the t-dependence explicit, the Klein–Gordon field becomes, ac-
cording to eqs. (2.34) and (2.39),
d3 p
Z
ap e−ipx + a†p eipx .
φ(x, t) = 3
p (2.61)
(2π) 2Ep
As usual, the symbols p and x denote 4-vectors, i.e. px = Ep t−px. In the term containing
a†p (t), we have substituted p → −p.
Note that the field described by eq. (2.61) contains a term with positive frequency
(proportional to e−ipx ) and a term with negative frequency (proportional toe e+ipx ).
The positive frequency term comes with an operator that destroys a positive energy
state, and the negative frequency term comes with an operator that creates a positive
energy state. So the Hilbert space contains only positive energy states, and thus has
a straightforward physical interpretation—in contrast to the wave function solutions to
the Klein–Gordon equation in relativistic quantum mechanics.
17
Chapter 2 The Klein–Gordon Field
Figure 2.1: The integration contour for the p0 integral in the Feynman propagator,
shown in the complex p0 plane. Figure taken from [1].
(In general, we should write D(x, y), but translational invariance implies that the depen-
dence can only be on x − y.) A solution of the equation
d4 p −ip(x−y)
Z
D(x − y) = e D̃(p) (2.65)
(2π)4
d4 p −ip(x−y) d4 p −ip(x−y)
Z Z
2 2
(∂x + m ) e D̃(p) = −i e (2.66)
(2π)4 (2π)4
d4 p −ip(x−y) d4 p −ip(x−y)
Z Z
2 2
⇔ e (−p + m ) D̃(p) = −i e . (2.67)
(2π)4 (2π)4
We arrive at
d4 p ie−ip(x−y)
Z
D(x − y) = . (2.68)
(2π)4 p2 − m2
Note that in going from eq. (2.67) to eq. (2.68), we have divided by p2 − m2 , which
is only well-defined if p2 6= m2 . This, of course, need not be the case—in particular,
p2 = Ep2 − p2 = m2 is just the relativistic mass-shell condition, satisfied for “real”
particles. We therefore have to regularize eq. (2.68). We do so by shifting the poles away
from the real axis by an inifinitesimal amount . There are four ways of doing this, and
all of them lead to valid Green’s functions.
18
2.4 The Feynman Propagator for the Klein–Gordon Field
19
Chapter 2 The Klein–Gordon Field
d3 p
Z
= e−ip(x−y) . (2.75)
(2π)3 2Ep
In other words, the Feynman propagator always describes the propagation of a particle
in the positive time direction. This motivates the particular choice for the shift of the
poles in eq. (2.69). We also see now why DF (x − y) deserves the name “propagator”.
A more compact way of writing DF (x−y) is obtained by introducing the time ordering
symbol T , which tells us to order any following field operators according to the zero
component of their argument in descending order:
20
The Dirac Field
3
3.1 The Dirac Equation and its Solutions
3.1.1 The Equation and the Corresponding Lagrangian
Now that we have some understanding of how to quantize a scalar field, let us repeat the
same for fermions. Our starting point is the Dirac equation (written in covariant form)
where we use the notation ∂/ ≡ γ µ ∂µ , and γ µ are the Dirac matrices, which satisfy the
algebra
{γ µ , γ ν } = 2g µν . (3.2)
Here, g µν = (1, −1, −1, −1) is the Minkowski metric and {·, ·} is the anticommutator, i.e.
{γ µ , γ ν } ≡ γ µ γ ν + γ ν γ µ . (3.3)
One can get quite far without specifying a specific representation for the Dirac matrices
γ µ , but it is much easier to have one in mind. In the following, we will always use the
chiral representation, which reads
σi
0 0 12×2 i 0
γ = γ = . (3.4)
12×2 0 −σ i 0
Here, σ i are the Pauli matrices. Often, we will also encounter the 5-th gamma matrix
γ 5 ≡ iγ 0 γ 1 γ 2 γ 3 (3.5)
i
= − µνρσ γµ γν γρ γσ . (3.6)
4!
21
Chapter 3 The Dirac Field
The matrix γ 5 has the property that it anticommutes with the other γ matrices, as can
be easily seen by using eq. (3.6) and the anticommutator (3.2).
Note that, by taken the Hermitian transpose of the Dirac equation and multiplying
by γ 0 from the right, we can immediately derive an equation for ψ̄ ≡ ψ † γ 0 . (We will see
shortly why it is useful to consider ψ̄ instead of simply ψ † .)
γ 0 γ µ† γ 0 = γ µ , (3.9)
which holds for the chiral representation (but not all representations) of the Dirac ma-
trices, as can be easily checked by direct computation. The complex conjugate Dirac
equation then becomes
The Lagrangian from which the Dirac equation and its complex conjugate are obtained
is
To check this, simply apply the Euler–Lagrange equations eq. (2.8), taking into account
that ψ(x) and ψ̄(x) are independent fields. (We could equivalently treat the real and
imaginary parts of ψ(x) as the independent degrees of freedom, but this would lead to
much more cumbersome equations.)
where p is an on-shell 4-momentum (i.e. p0 = Ep ) and the index s denotes the spin
orientation. The physical meaning of the negative energy solutions is a highly nontrivial
topic in relativistic quantum mechanics. Here, we will neglect it for the moment since
field theory will provide an elegant solution once we have proceeded to the quantized
22
3.1 The Dirac Equation and its Solutions
Dirac field. The spinors us (p) and v s (p) must satisfy the relations (momentum space
Dirac equation)
s
/ − m)u (p) = 0 ,
(p
s
(p
/ + m)v (p) = 0 ,
(3.14)
ūs (p)(p
/ − m) = 0 ,
v̄ s (p)(p
/ + m) = 0
for ψp,s,+ (x) and ψp,s,− (x) to satisfy the Dirac equation eq. (3.1).
Explicitly, we have
√ √
p · σ ξs p · σ ξs
s s
u (p) = √ and v (p) = √ . (3.15)
p · σ̄ ξ s − p · σ̄ ξ s
We introduced here even more notation, namely we define the 4-vectors σ ≡ (12×2 , σ)
and σ̄ ≡ (12×2 , −σ). The 2-component vectors ξ 1 and ξ 2 distinguishing the two spin
orientations are simply two orthonormal basis vectors of R2 , for instance ξ 1 = (1, 0) and
ξ 2 = (1, 0). We can see that us (p) and v s (p) are indeed solutions of eq. (3.14) in the
following way:
√
p · σ ξs
s 0 p·σ
p
/u (p) = p · σ̄ √
0 p · σ̄ ξ s
p 2 p · σ̄ ξ s
(p · σ)
= p
(p · σ̄)2 p · σ ξ s
p 2 (E + p · σ) ξ s
(E − p · σ)
= p
(E + p · σ)2 (E − p · σ) ξ s
p
(E 2 − 2E(p · σ) + p2 ) (E + p · σ) ξ s
= p 2
(E + 2E(p · σ) + p2 ) (E − p · σ) ξ s
p 3
E − 2E 2 (p · σ) + Ep2 + E 2 (p · σ) − 2Ep2 + p2 (p · σ) ξ s
= p
E 3 + 2E 2 (p · σ) + Ep2 − E 2 (p · σ) − 2Ep2 − p2 (p · σ) ξ s
√
m √E − p · σ ξ s
=
m E + p · σ ξs
√
m p · σ ξs
= √
m p · σ̄ ξ s
= mus (p) . (3.16)
23
Chapter 3 The Dirac Field
where ūs ≡ (us )† γ 0 . Equation (3.17) can be easily checked by explicit calculation. It
is also easy to check that normalizing u† u rather than ūu would not yield a Lorentz-
invariant normalization condition. In fact,
us† (p) ur (p) = v s† (p) v r (p) = 2Ep δ rs . (3.18)
Therefore, in QFT, we almost always work with ψ̄ instead of ψ † . Note that the u and v
spinors are orthogonal:
ūs (p) v r (p) = −v̄ s (p) ur (p) = 0 . (3.19)
A similar relation does not hold for us† (p) v r (p) and v s† (p) ur (p) — these products are
in general nonzero. A relation that is, however, useful sometimes is
us† (p) v r (−p) = v s† (p) ur (−p) = 0 . (3.20)
(We write the argument of the u and v spinors as 3-vectors here to emphasize the sign
of the 3-momentum. It is of course implied that the 0-component of the momentum
4-vector is set by the relativistic energy–momentum relation.)
Plugging in eq. (3.15) for us (p) and using s=1,2 ξ s ξ s† = 12×2 , we obtain
P
X X √p · σξ s √ √
s s
u (p)ū (p) = √ s (ξ s† p · σ̄, ξ s† p · σ) (3.22)
p · σ̄ξ
s=1,2 s=1,2
√ √ √ √
p · σ p · σ̄ p·σ p·σ
= √ √ √ √ (3.23)
p · σ̄ p · σ̄ p · σ̄ p · σ
m p·σ
= (3.24)
p · σ̄ m
=p
/ + m. (3.25)
In the third equality, we have used
(p · σ)(p · σ̄) = (p0 − pi σ i )(p0 + pi σ i ) (3.26)
= (p0 )2 − (pi )2 (3.27)
2
=m . (3.28)
In analogy to eq. (3.25), we can also show that
X
v s (p)v̄ s (p) = p
/ − m. (3.29)
s=1,2
24
3.2 Quantization of the Dirac Field
where a, b = 1 · · · 4 are spinor indices. Note that, again, we take the time coordinate
of the two fields to be the same. Let us try to quantize the Dirac field based on these
commutation relations and see what goes wrong.
We write the field ψ(x) as a superposition of all solutions of the free Dirac equation,
with operator-valued coefficients:
d3 p
Z X
s s −ipx s s ipx
ψ(x) = p a p u (p)e + bp v (p)e . (3.31)
(2π)3 2Ep s=1,2
Note that, unlike for the real Klein–Gordon field, the coefficients of eipx and e−ipx are
not related to each other for the Dirac field, which is complex. If the creation and
annihilation operators asp and bsp satisfy the commutation relations
(and all other commutators being zero) we can show that ψ and ψ † satisfy the postulated
relations eq. (3.30):
d3 p d3 q
Z X
†
[ψ(x, t), ψ (y, t)] = 6
p p eipx−iqy
(2π) 2Ep 2Eq r,s
s†
× [arp , as†
q ]u r
(p)ūs
(q) + [br
,
−p −qb ]v r
(−p)v̄ s
(−q) γ0 (3.33)
d3 p
Z
eip(x−y) γ 0 Ep − γ i p + m + γ 0 Ep + γ i p − m γ 0
= 3
(2π) 2Ep
(3.34)
= δ (3) (x − y) 14×4 . (3.35)
Let us look at the quantized Hamilton operator, which is as usual obtained from a
Legendre transform of the Lagrangian:
Z
3 δL
H= d x ∂0 ψ − L (3.36)
δ(∂0 ψ)
Z
= d3 x iψ̄γ 0 ∂0 ψ − ψ̄(i∂/ − m)ψ
(3.37)
25
Chapter 3 The Dirac Field
Z
d3 x ψ̄ − i(γ · ∇) + m ψ
= (3.38)
d3 p d3 q
Z Z Z Xh
−ipx
= d3 x as† s ipx
+ bs† s
3
p
3
p p ū (p)e p v̄ (p)e
(2π) 2Ep (2π) 2Eq s,r
i
· − i(γ · ∇) + m arq ur (q)e−iqx + brq v r (q)eiqx
(3.39)
d3 p
Z h
as† s
r r
= 3 p ū (p) γ · p + m ap u (p)
(2π) 2Ep
+ as† s
r r
p ū (p) γ · p + m b−p v (−p)
+ bs† s
r r
p v̄ (p) − γ · p + m a−p u (−p)
r r i
+ bs†
p v̄ s
(p) − γ · p + m bp u (p) (3.40)
d3 p X
Z
Ep as† s s† s
= 3 p ap − Ep bp bp . (3.41)
(2π) s
In the fourth equality, we have plugged in eq. (3.31) and its complex conjugate. In the
sixth equality, we have used γ · p = −p / + Ep γ 0 and we have then invoked the momen-
tum space Dirac equation, eq. (3.14), as well as the orthogonality relations eqs. (3.18)
and (3.20).
We encounter a serious problem here: if we interpret the operators as† s s† s
p ap and bp bp as
particle number operators again, we are led to the conclusion that, the more particles we
create using the operator bs†
p , the lower the energy becomes. This seems like a terribly
unstable system.
A discussion of additional problems with the above procedure is given in [1].
d3 p X
Z
Ep as† s s s†
H= 3 p ap − Ep bp bp (3.42)
(2π) s
d3 p X
Z
Ep as† s s† s 3 (0)
= 3 p ap − Ep bp bp − Ep (2π) δ (0) . (3.43)
(2π) s
26
3.2 Quantization of the Dirac Field
and postulate
(We have already dropped the infinite constant.) The field operators ψ(x, t) and ψ † (y, t)
now also satisfy anticommutation relations:
d3 p d3 q
Z X
† ipx−iqy
{ψ(x, t), ψ (y, t)} = p p e
(2π)6 2Ep 2Eq r,s
r†
× {arp , as†
q }u r
(p)ūs
(q) + {b ,
−p −q bs
}v r
(−p)v̄ s
(−q) γ 0 (3.49)
d3 p
Z
eip(x−y) γ 0 Ep − γ i p + m + γ 0 Ep + γ i p − m γ 0
= 3
(2π) 2Ep
(3.50)
= δ (3) (x − y) 14×4 . (3.51)
27
Chapter 3 The Dirac Field
The interpretation of the operators asp and bsp is that both of them create (distinct)
particles with momentum p and positive energy Ep .
28
3.3 The Feynman Propagator for the Dirac Field
where the 4 × 4 matrix (in spinor space) in the denominator means the inverse of that
matrix. If one wishes to avoid matrix inverses, one can equivalently write
i(p
/ + m)
S̃(p) = , (3.58)
p2 − m2
using that p
/p/ = p2 . With this S̃(p), eq. (3.55) is infinite, so we again need to introduce
a regularization scheme. As for the Klein–Gordon field, and by the same arguments, we
use the Feynman prescription for shifting the poles away from the real axis. This defines
the Feynman propagator for the Dirac field :
d4 p i(p
/ + m) −ip(x−y)
Z
SF (x − y) = e . (3.59)
(2π) p − m2 + i
4 2
Like the Klein–Gordon propagator, also the Feynman propagator can be related to
correlation functions. Indeed note that, by integrating over p0 , we can write
d3 p
Z h i
−ip(x−y) 0 0 ip(x−y) 0 0
SF (x − y) = e ( p
/ + m)θ(x − y ) + e (−p/ + m)θ(y − x ) .
(2π)3 2Ep
(3.60)
On the other hand, we also have
d3 p
Z X
h0|ψ(x)ψ̄(y)|0i = 3
us (p)ūs (p)e−ip(x−y) (3.61)
(2π) 2Ep s
d3 p
Z
−ip(x−y)
= (p
/ + m)e . (3.62)
(2π)3 2Ep
and
d3 p
Z X
h0|ψ̄(y)ψ(x)|0i = v s (p)v̄ s (p)eip(x−y) (3.63)
(2π)3 2Ep s
d3 p
Z
ip(x−y)
= / − m)e
(p . (3.64)
(2π)3 2Ep
Thus, we find again, similar to what we found for the Klein–Gordon field,
SF (x − y) = h0|ψ(x)ψ̄(y)|0i θ(x0 − y 0 ) − h0|ψ̄(y)ψ(x)|0i θ(y 0 − x0 ) (3.65)
= h0|T ψ(x)ψ̄(y)|0i . (3.66)
For the Klein–Gordon field, there was a plus sign between the two terms. The fact that
we now find a minus sign is a reflection of the anticommutating nature of fermion fields.
In the second line of eq. (3.66), we have extended the definition of the time ordering
symbol T to fermion fields. It is implied that any interchange of adjacent fermion field
operators that is necessary to bring the field order into time order contributes a minus
sign.
The physical interpretation of the Feynman propagator for Dirac fields is the following:
it describes the propagation of a particle from y to x (if x0 > y 0 ), or the propagation of
an antiparticle from x to y (if y 0 > x0 ).
29
Chapter 3 The Dirac Field
Figure 3.1: An active transformation (in this case a rotation) of a field configuration.
Figure taken from [1].
We write the transformation of the spinor field ψ(x) under this transformation as
30
3.4 Symmetries of the Dirac Theory
This consistency condition basically says that the index µ on γ µ can indeed be treated
like any other Lorentz index.
Let us now construct S(Λ) explicitly. To do so, consider an infinitesimal Lorentz
transformation and write Λµν as
Λµν = g µν + ω µν ,
µ (3.73)
(Λ−1 ) ν = g µν − ω µν ,
where ω µν is infinitesimal. We know from the explicit form of Λµν , familiar from special
relativity, that ω µν is antisymmetric.
We can write S(Λ) and its inverse as
i
S(Λ) = 1 − σµν ω µν ,
4 (3.74)
−1 i
S (Λ) = 1 + σµν ω µν ,
4
with a yet-to-be-determined tensor σµν . Since ω µν is antisymmetric in µν, so must σµν
be as well. Plugging the infitesimal forms of Λµν and S(Λ), eqs. (3.73) and (3.74), into
the transformation law for γ µ , eq. (3.72), we find
i i
σαβ ω αβ γ µ − γ µ σαβ ω αβ = ω µν γ ν , (3.75)
4 4
or, equivalently,
In the last step, we have used the antisymmetry of ω αβ . Since eq. (3.76) must be satisfied
for any ω αβ , it reduces to
31
Chapter 3 The Dirac Field
In an analogous way, also the action of a Lorentz transformation on a spinor field ψ(x),
given by eq. (3.68), can be expressed in terms of a generating operator Jµν . Consider
again an infinitesimal Lorentz transformation as in eq. (3.73) and write
32
3.4 Symmetries of the Dirac Theory
(The factor i/2 is mere convention.) Plugging in our expressions for S(Λ) and Λ in the
infinitesimal case, eqs. (3.73) and (3.74), we obtain
i i
1 − σµν ω µν ψ(xµ − ω µν xν ) = 1 − Jµν ω µν ψ(x) . (3.88)
4 2
This leads to
i i
− σµν ω µν − ω µν xν ∂µ = − Jµν ω µν , (3.89)
4 2
or
1
Jµν = σµν + i(xµ ∂ν − xν ∂µ ) . (3.90)
2
In the last step, we have used the antisymmetry of ω µν . Note that the Jµν are block
diagonal in the chiral basis for the Dirac matrices. This follows from the definition of
σµν , eq. (3.78), together with the explicit form of the γ µ in the chiral basis, eq. (3.2). It
means that the spinor representation of the Lorentz group is reducible: if the upper or
lower two components in a 4-spinor are zero, they will remain zero under any Lorentz
transformation.
Spin
We know that angular momentum is the conserved quantity associated with symmetry
under spatial rotations. Therefore, we can now use the transformation properties of the
field operator under such rotations to derive the form of the angular momentum operator
for a fermion field. This will in particular help us better understand the internal angular
momentum (spin) of fermions.
Under an infinitesimal Lorentz transformation (of which spatial rotations are of course
a special case), a fermion field transforms as
with
(see eq. (3.68)). Using the explicit form eq. (3.74) for S(Λ) and eq. (3.73) for Λ, this
becomes
i ∂
δψ(x) = − σµν ω µν ψ(x) − ω µν xν µ ψ(x) . (3.93)
4 ∂x
We now specialize to a rotation about the z axis (ω 12 = −ω21 = θ, all other ω µν = 0),
and use that
i
σ 12 = [γ 1 , γ 2 ] (3.94)
2
33
Chapter 3 The Dirac Field
i −σ 1 σ 2 + σ 2 σ 1
= (3.95)
2 −σ 1 σ 2 + σ 2 σ 1
3
σ
= (3.96)
σ3
≡ Σ3 . (3.97)
The right hand side is thus the third component of the angular momentum operator for a
fermion field. In exact analogy, also the other two components of the angular momentum
operator can be found, leading to
Z
J = d3 x ψ † x × (−i∇) + 21 Σ ψ .
(3.100)
The first term in square brackets is just the angular momentum operator in quantum
mechanics (dressed here with two field operators). For non-relativistic fermions, it can be
interpreted as giving the orbital angular momentum. The second term in square brackets
gives the internal angular momentum (spin). For relativistic fermions, this division into
spin and orbital angular momentum is not so trivial due to spin–orbit coupling.
Nevertheless, eq. (3.100) allows us to check that a Dirac fermion had indeed spin 1/2.
For this, it is sufficient to consider particles at rest (or very nearly so), so that the orbital
angular momentum term is negligible. In this case, we can plug the Fourier expansion
of ψ(x), eq. (3.44) into eq. (3.100) (omitting the orbital angular momentum term):
d3 p d3 p0
Z Z
0
j
J = d x 3
6
p p e−ip x eipx (3.101)
(2π) 2Ep 2Ep0
X r0 † 0 0 0
Σj
· ap0 ur † (p0 ) + br−p0 v r † (−p0 ) arp ur (p) + br†
−p v r
(−p) . (3.102)
0
2
r,r
To apply this operator (or specifically its third component) to a one-particle state of zero
momentum, as† z
0 |0i, we use the fact that the vacuum has zero spin, i.e. J |0i = 0 and
therefore
J z as† z s†
0 |0i = [J , a0 ]|0i . (3.103)
34
3.4 Symmetries of the Dirac Theory
(3.108)
35
Chapter 3 The Dirac Field
We say that a particle whose spin is pointing in the direction of motion has a right-
handed (RH) helicity and a particle with its spin is pointing in the opposite direction
has a left-handed (LH) helicity.
We observe that, in the ultrarelativistic limit, for spinors for LH particles and RH
antiparticles, the lower components are zero, while for RH particles and LH antiparticles
the upper components vanish. This shows again that the upper and lower components of
a spinor are quite independent: they correspond to different spin orientations. We have
argued previously, below eq. (3.90), on more formal grounds that the upper and lower
components of a Dirac spinor are independent because the 4-dimensional spinor represen-
tation of the Lorentz group is reducible. It factorizes into two irreducible representations,
one describing the transformation properties of the upper two components of the Dirac
spinor, one describing the transformation properties of the lower two components.
It also makes sense that LH particles and RH antiparticles are in the same irreducible
representation of the Lorentz group. In the Dirac hole theory, an antiparticle corresponds
to the absence of a particle from the Dirac sea. If a left-handed particle is absent, the hole
it leaves back effectively has opposite spin and corresponds to a right-handed antiparticle.
To highlight the factorization of the Lorentz representation, one sometimes splits up
a Dirac spinor field into
ψL (x)
ψ(x) = (3.109)
ψR (x)
and works with the left-chiral Weyl fermion ψL (x) and the right-chiral Weyl fermion
ψR (x) separately. Note that, according to eqs. (3.107) and (3.108), in the ultrarelativistic
limit the Weyl fermions are helicity eigenstates, justifying the attribute “left” and “right”.
(Even though a left-chiral field corresponds to a LH particle, but a RH antiparticle.) In
terms of the Weyl fields the Dirac Lagrangian eq. (3.11) then becomes
†
L = ψL† (iσ̄ · ∂)ψL + ψR (iσ · ∂)ψL − mψ̄L ψR − mψ̄R ψL . (3.110)
We see that it is only the mass term that mixes the left- and right-chiral fields, thereby
destroying the one-to-one correspondence between chirality and helicity. This is a reflec-
tion of the fact that for massive fermions, helicity is a frame-dependent quantity. One
can always boost into a frame where the momentum of the particle is reversed.
When working with Weyl spinors, it is convenient to define the projection operators
1 − γ5 1 + γ5
PL ≡ PR ≡ , (3.111)
2 2
which project out the left-chiral and right-chiral components, respectively, of a Dirac
spinor. They satisfy PL/R2 = PL/R and PL PR = 0.
Weyl fermions play a crucial role in the theory of weak interactions, where the couplings
of left- and right-chiral fields are completely different.
36
3.4 Symmetries of the Dirac Theory
3.4.2 Parity (P )
In addition to the continuous Lorentz transformations discussed above, there are several
discrete symmetries under which many quantum field theories are invariant. The first of
these is parity or space inversion which transforms a Lorentz 4-vector (t, x) into (t, −x).
A parity transformation thus reverses the 3-momentum of a particle, but leaves its spin
orientation unchanged. To see this, remember that orbital angular momentum L is a cross
product of x and p, and since both of these change sign under parity transformations,
L remains invariant. Spin and orbital angular momentum must transform the same
way if spin is to be interpreted as a form of angular momentum, therefore also spin
must remain invariant under parity. Consequently, the parity operator must transform
a quantum state asp |0i into as−p |0i. This implies that the parity operation P in Hilbert
space must act as
P asp P = ηa as−p and P bsp P = ηb bs−p , (3.112)
where ηa and ηb are phase factors. We are in principle free to choose these phase factors
arbitrarily—they are not restricted by the requirement that parity reverses space. They
are, however, constrained if we demand that applying the parity operator twice should
leave any physical observable unchanged. Since observables (such as the Hamiltonian)
are constructed from even numbers of fermion field operators, this implies ηa , ηb = ±1.
When we discussed continuous Lorentz transformations, we found that their action of
the field operator can be written as a multiplication by a matrix S(Λ) in spinor space, see
eq. (3.68). We will now find a similar transformation law also for parity transformations.
We write
d3 p
Z
ηa as−p us (p)e−ipx + ηb∗ bs†
X
s ipx
P ψ(x)P = 3
p −p v (p)e . (3.113)
(2π) 2Ep
We now substitute p = (p0 , p) → p0 ≡ (p0 , −p) in the integral. In applying this
substitution, we must express us (p) and v s (p) in terms of p0 . To do so, we use the explicit
form of these spinors (see eq. (3.15)) and observe that p · σ = p0 · σ̄ and p · σ̄ = p0 · σ.
This leads to
√
p · σξ s p · σ̄ξ s
√ 0
us (p) = √ = √ = γ 0 us (p0 ) , (3.114)
p · σ̄ξ s p0 · σξ s
√ √ 0
p · σξ s s
p · σ̄ξ
s
v (p) = √ = √ = −γ 0 v s (p0 ) . (3.115)
− p · σ̄ξ s − p0 · σξ s
With these relations, we arrive at
d3 p0
Z
s 0 s 0 −ip0 x0 ∗ s† 0 s 0 ip0 x0
X
P ψ(x)P = p η a ap0 γ u (p )e − ηb bp0 γ v (p )e . (3.116)
(2π)3 2Ep0
where we have defined x0 ≡ (x0 , −x). We now choose ηb∗ = −ηa . Then, we have the
transformation law
P ψ(x)P = ηa γ 0 ψ(x0 ) . (3.117)
37
Chapter 3 The Dirac Field
Table 3.1: Transformation properties of Dirac field bilinearies under parity (P ), time
reversal (T ) and charge conjugation (C). The shorthand notation (−1)µ means +1 for
µ = 0 and −1 for µ = 1, 2, 3.
Let us discuss the transformation propertiers of these expressions under parity. Since
ψ̄(x) appears in all of them, we need the result that
In other words, ψ̄(x)γ µ ψ(x) transforms like a Lorentz vector. The transformation prop-
erties of the other bilinearies can be derived in a similar way. They are summarized in
table 3.1.
T asp T = a−s
−p and T bsp T = b−s
−p , (3.122)
In principle, we should again allow for arbitrary phase factors, but since they would not
affect the following derivations, we omit them here. This is, however, not the full story
38
3.4 Symmetries of the Dirac Theory
yet: given some quantum mechanical transition amplitude hψ1 |ψ2 i, time reversal should
interchange the initial and final states, i.e. hT ψ1 |T ψ2 i = hψ2 |ψ1 i. (Such an operator is
called antiunitary.)
This implies that T not only acts on Hilbert space states, but also on complex numbers
by sending them to their complex conjugate. We thus have
d3 p
Z
−s† s∗
X
−s s∗ +ipx −ipx
T ψ(x)T = p a−p u (p)e + b−p v (p)e . (3.123)
(2π)3 2Ep
To write the right hand side as a linear transformation of ψ(x), we need a way of writing
us (p) in terms of u−s (p0 ), where p0 = (p0 , −p).
To do so, we first note again that p0 · σ = p · σ̄ and p0 · σ̄ = p · σ. Moreover, note that
a spin flip (sending s to −s) is achieved by the transformation
ξ s → ξ −s = −iσ 2 (ξ s )∗ . (3.124)
(We are not worrying about the inconsequential phase factors here.) To see this, assume
ξ s describes a spin along a unit vector n, i.e. ξ s is an eigenvector of the helicity operator
n · σ:
n · σξ s = +ξ s . (3.125)
Then, using σ i σ 2 = −σ 2 σ i∗ ,
Therefore, we have
√
p · σ ξs
s √
u (p) = (3.128)
p · σ̄ ξ s
2 (ξ −s )∗ ]
√ 0
p · σ̄ [−iσ
= √ 0 . (3.129)
p · σ [−iσ 2 (ξ −s )∗ ]
We would like to move the σ 2 matrices to the left of the square roots. To √ this end,
go to a reference frame where p is aligned along the z-axis. In this frame p0 · σ =
1−σ 3 1+σ 3
i 2 2 i∗
p p
Ep0 + |p0 |( 2 ) + Ep0 − |p0 |( 2 ). Therefore, using again σ σ = −σ σ , we find
p p
p0 · σ σ 2 = σ 2 p0 · σ̄ ∗ . (3.130)
Similarly,
p p
p0 · σ̄ σ 2 = σ 2 p0 · σ ∗ . (3.131)
39
Chapter 3 The Dirac Field
σ2
= −i [u−s (p0 )]∗ (3.133)
σ2
= −γ 1 γ 3 [u−s (p0 )]∗ . (3.134)
d3 p0
Z
−ip0 x0 s† s ip0 x0
X
T ψ(x)T = −γ 1 γ 3 s s
p ap0 u (p)e + bp0 v (p)e (3.136)
(2π)3 2Ep0
= −γ 1 γ 3 ψ(x0 ) . (3.137)
To write the action of C on the field ψ(x) as a linear operation as we did for the P and T
transforms, we will need a transformation that converts a u spinor into a v spinor. Here
it is:
√
p · σ ξs
s √
v (p) = (3.139)
− p · σ̄ ξ s
√
p · σ [−iσ 2 ξ −s ]∗
= √ (3.140)
− p · σ̄ [−iσ 2 ξ −s ]∗
√
−iσ 2 p · σ̄ ∗ (ξ −s )∗
= √ (3.141)
iσ 2 p · σ ∗ (ξ −s )∗
√ ∗
0 −iσ 2 p · σ ξ −s
= √ (3.142)
iσ 2 0 p · σ̄ ξ −s
0 −iσ 2
= [us (p)]∗ (3.143)
iσ 2 0
= −iγ 2 [us (p)]∗ . (3.144)
In the second equality, we have used eq. (3.124). Note that in the last step we have us (p)
and not u−s (p) because we have seen in section 3.4.1 that the association between the
40
3.4 Symmetries of the Dirac Theory
ξ s 2-spinors and the physical spin orientation is opposite for u and v spinors: ξ s = (1, 0)
corresponds to spin up along the z axis for a u spinor, but to spin down along the z axis
for a v spinor. Similarly, we also have
(Note that to obtain eq. (3.145) directly from eq. (3.144), one has to take into account
that, in our conventions from eq. (3.124), ξ −(−s) = −ξ s .) Thus,
d3 p
Z X
bsp us (p)e−ipx + as† s ipx
Cψ(x)C = 3
p p v (p)e (3.146)
(2π) 2Ep
d3 p
Z X
2
bsp [v s (p)]∗ e−ipx + as† s ∗ ipx
= −iγ 3
p p [u (p)] e (3.147)
(2π) 2Ep
= −iγ 2 [ψ(x)]∗ (3.148)
0 2 T
= −i(ψ̄γ γ ) . (3.149)
The transformation properties of the various Dirac field bilinearies under this transfor-
mation are again summarized in table 3.1.
41
Chapter 3 The Dirac Field
42
Interacting Fields and Feynman Diagrams
4
In the previous chapters, we have considered rather boring systems: non-interating scalar
and fermion fields. Of course, our ultimate goal is to study the interactions among fields,
and in particular to compute the transition amplitude from some initial state to some
final state as well as the associated cross section. We will now develop the tools to do
that. We will first develop a formalism for evaluating so-called correlation functions in
sections 4.1 to 4.3. Then, we will learn how to relate correlation functions to scattering
matrix elements in section 4.4. In section 4.5, we will finally employ our formalism to
evaluate scattering amplitudes. Along the way, we will get to know Feynman diagrams
as a pictorial way of representing the mathematical expressions for correlation functions
and scattering amplitudes. Up to this point, we will work in a theory containing only
a real scalar field φ(x), but no fermions. We will generalize the formalism to include
fermions in section 4.6.
43
Chapter 4 Interacting Fields and Feynman Diagrams
grangian:
1 1 λ
L = (∂µ φ)2 − m2 φ2 − φ4 . (4.1)
2 2 4!
With the Fourier expansion of φ(x) in terms of creation and annihilation operators in
mind, we realize that the φ4 term describes for instance the scattering of two φ particles,
which can be viewed as the annihilation of two incoming particles, followed by the cre-
ation of two outgoing particles with possibly different momentum vectors. The quantity
λ in eq. (4.1) is a coupling constant that determines the strength of the interaction be-
tween φ particles, and the factor 4! in the denominator is included by convention. The
motivation for writing the interaction term in this way will become clear later in this
chapter.
We have already encountered such a correlation function in the free Klein–Gordon theory:
the Feynman propagator was given by the two-point correlation function
d4 p ie−ip(x−y)
Z
DF (x − y) = h0|T φ(x)φ(y)|0i = , (free theory) (4.3)
(2π)4 p2 − m2 + i
44
4.1 Time-Dependent Perturbation Theory for Correlation Functions
see eqs. (2.69) and (2.77). In the interacting theory, the expression will be more compli-
cated of course. In fact, it is impossible to write the propagator of the interacting theory
in closed form, but if λ is not too large, the interaction can be treated as a small per-
turbation, and we can use time-dependent perturbation theory to approximate it quite
well. This is what we are going to do now. After we have studied the two point corre-
lation function, we can easily generalize the formalism to more complicated correlation
functions and then to transition matrix elements.
We will make the connection between correlation functions and transition matrix ele-
ments in section 4.4 when we discuss the LSZ reduction formula.
H = H0 + Hint , (4.4)
where
d3 p
Z
Ep a†p ap
H0 ≡ 3
(4.5)
(2π)
(Remember that the Legendre transform that relates the Lagrangian and the Hamilto-
nian, eq. (2.13) acts nontrivially only on terms containing derivatives of the field, while
those without derivatives just get a minus sign.)
d3 p
Z
ipx † −ipx
φ(x, t0 ) = p ap (t0 )e + a p (t0 )e . (4.7)
(2π)3 2Ep
45
Chapter 4 Interacting Fields and Feynman Diagrams
In quantizing the field, we impose as before that the creation and annihilation operators
satisfy the canonical commutation relations eq. (2.45).1 The full time-dependent field at
arbitrary time t is then formally given by
We will define the interaction picture field operator φI (x, t), which contains the time-
dependence due to H0 , but not the one due to Hint . In other words, we define
This expression is of course (and not by chance) identical in form to the expression
for the time-dependent field operator in the free Klein–Gordon theory, eq. (2.61). In
the following, we will write the creation and annihilation operators at t = t0 simply
as a†p and ap , omitting the explicit mentioning of t0 . φI is an operator that we can
work with: it contains only c-numbers and creation and annihilation operators that obey
the canonical commutation relations. All the technology developed in chapter 2 can be
reused. Therefore, we would like to rewrite the two point correlations hΩ|T φ(x)φ(y)|Ωi
entirely in terms of φI . We have
which describes all the non-trivial time-dependence of the theory. It satisfies the Schrödinger-
like equation
∂
U (t, t0 ) = eiH0 (t−t0 ) − H0 + H e−iH(t−t0 )
i (4.14)
∂t
= eiH0 (t−t0 ) − H0 + H e−iH0 (t−t0 ) U (t, t0 )
(4.15)
≡ Hint,I (t)U (t, t0 ) . (4.16)
46
4.1 Time-Dependent Perturbation Theory for Correlation Functions
Z
λ 4
= d3 x φ (x, t) . (4.18)
4! I
The solution of the above Schrödinger equation eq. (4.16), with the initial condition
U (t0 , t0 ) = 1, can be written as
Z t
U (t, t0 ) = 1 − i dt1 Hint,I (t1 )U (t1 , t0 ) (4.19)
t0
by simply integrating both sides. Plugging this expression into itself repeatedly, we
obtain the infinite series
Z t Z t Z t1
U (t, t0 ) = 1 + (−i) dt1 Hint,I (t1 ) + (−i)2 dt1 dt2 Hint,I (t1 )Hint,I (t2 )
t0 t0 t0
Z t Z tn−1
+ . . . + (−i)n dt1 · · · dtn Hint,I (t1 ) · · · Hint,I (tn ) + . . . . (4.20)
t0 t0
If λ, and thus Hint,I , is small, we can truncate the series after the first order or second
order term. Note that the integration boundaries are such that the integration vari-
ables t1 · · · tn satisfy t1 > t2 > · · · > tn . Therefore, using the time ordering symbol T
introduced in section 2.4.3, we can simplify the second order term in eq. (4.20) into
Z t Z t1 Z t Z t
2 1
(−i) dt1 dt2 Hint,I (t1 )Hint,I (t2 ) = (−i)2 dt1 dt2 T [Hint,I (t1 )Hint,I (t2 )] .
t0 t0 2 t0 t0
(4.21)
Here, we have used the fact that, with the time ordering symbol included, the integrans
is symmetric under t1 ↔ t2 , i.e. extending the domain of the t2 integral from [t0 , t1 ] to
[t0 , t] just double counts every point (see fig. 4.1. This is compensated by a prefactor
1/2.
The above arguments can be extended for the higher order terms in the perturbation
series:
Z t Z tn−1
n
(−i) dt1 · · · dtn Hint,I (t1 ) · · · Hint,I (tn )
t0 t0
Z t Z t
1
= (−i)n dt1 · · · dtn T [Hint,I (t1 ) · · · Hint,I (tn )] . (4.22)
n! t0 t0
Then, the perturbation series eq. (4.20) becomes an exponential series and we can write
Z t
0 0
U (t, t0 ) = T exp − i dt Hint,I (t ) , (4.23)
t0
where it is understood that the time ordering applies to each term in the exponential
series.
47
Chapter 4 Interacting Fields and Feynman Diagrams
Figure 4.1: Illustration of the arguments given below eq. (4.21), justifying why the
integration domain of the t2 integral in that equation can be extended from [t0 , t1 ] to
[t0 , t]. Figure taken from [1].
It is easy to check that also U (t, t0 ) satisfies the Schrödinger-like eq. (4.16),
∂
i U (t, t0 ) = Hint,I (t)U (t, t0 ) . (4.25)
∂t
but now with the initial condition U (t, t0 ) = 1 at t = t0 . Another way of writing U (t, t0 )
is
0 0
U (t, t0 ) = eiH0 (t−t0 ) e−iH(t−t ) e−iH0 (t −t0 ) . (4.26)
This can be seen by checking that both eq. (4.24) and eq. (4.26) satisfy eq. (4.25) with
the same initial condition. Equation (4.26) shows in particular that U (t, t0 ) is unitary,
i.e. [U (t, t0 )]† U (t, t0 ) = 1 and that
48
4.1 Time-Dependent Perturbation Theory for Correlation Functions
where n runs over a complete set of energy eigenstates of the interacting theory. The
interacting vacuum |Ωi is of course an energy eigenstate, with eigenvalue E0 , and can
therefore be separated from the sum:
X
e−iHT |0i = e−iE0 T |Ωi hΩ|0i + e−iEn T |ni hn|0i , (4.31)
n6=0
We would like to take the limit T → ∞, but for the oscillating exponentials, this limit
does not exist. We therefore have to regularize by taking the limit to a slightly imaginary
direction: T → ∞(1−i). Then, the term containing E0 is the slowest to die and therefore
dominates. We thus get rid of all states |ni except the desired vacuum state |Ωi. The
latter is thus given by
e−iHT |0i
|Ωi = lim . (4.32)
T →∞(1−i) e−iE0 T hΩ|0i
Since H0 |0i = 0 (see the explicit form eq. (4.5)), we can insert a factor eiH0 T before |0i.
Moreover, since T is very large, we can impose a small shift T → T + t0 without changing
the result. This leads to
e−iH(T +t0 ) eiH0 (T +t0 ) |0i
|Ωi = lim (4.33)
T →∞(1−i) e−iE0 (T +t0 ) hΩ|0i
U (t0 , −T )|0i
= lim −iE
. (4.34)
T →∞(1−i) e 0 (T +t0 ) hΩ|0i
49
Chapter 4 Interacting Fields and Feynman Diagrams
50
4.2 Wick’s Theorem
To generalize this two n > 2, it is useful to separate the field operator into a piece
containing only creation operators and a piece containing only annihilation operators:
−
φI (x) = φ+
I (x) + φI (x) , (4.43)
with
d3 p
Z
φ+
I (x) = p e−ipx ap , (4.44)
(2π)3 2Ep
d3 p
Z
φ−
I (x) = p e+ipx a†p . (4.45)
(2π)3 2Ep
(The superscripts + and − stand for positive and negative frequency.) This decomposi-
tion is useful because we know that
φ+
I (x)|0i = 0 and h0|φ−
I (x) = 0 . (4.46)
In the second line, we have rewritten the expression in such a way that all terms except
the commutator are in normal order, i.e. the creation and annihilation operators are
ordered such that all annihilation operators are on the right and all creation operator
are on the left. Normally ordered expression have the advantage that they vanish when
sandwiched between the vacuum |0i. There is even a special notation for normal ordering:
when operators are sandwiched between colons (: · :), it is implied that they should be
put into normal order. For example,
d3 p d3 p0
Z
0
+ −
[φI (x), φI (y)] = 6
p p e−ipx+ip y [ap , a†p0 ] (4.50)
(2π) 2Ep 2Ep0
d3 p 1 −ip(x−y)
Z
= e . (4.51)
(2π)3 2Ep
51
Chapter 4 Interacting Fields and Feynman Diagrams
If we consider, instead of the simple product φI (x)φI (y) the time-ordered product
T φI (x)φI (y), we obtain for x0 > y 0 exactly the above expression eq. (4.48). For x0 < y 0 ,
−
the normal ordered terms are the same, but the commutator is [φ+ I (y), φI (x)]. There is
again a special symbol to simplify the notation in this situation. The contraction of two
fields is defined as
(
−
[φ+ 0
I (x), φI (y)] for x ≥ y
0
φI (x)φI (y) ≡ −
. (4.52)
[φ+ 0
I (y), φI (x)] for x < y
0
d3 p 1 h −ip(x−y)
Z i
0 0 +ip(x−y) 0 0
φI (x)φI (y) = e θ(x − y ) + e θ(y − x ) . (4.53)
(2π)3 2Ep
Comparing to eq. (2.72), we see that the right hand side is just the Feynman propagator:
This was to be expected because we know that h0|T φI (x)φI (y)|0i = DF (x − y) and we
have argued that, in decomposition like eq. (4.48), only the commutator but not the
normally ordered terms contribute.
To summarize the discussion so far, we have shown that
From here on, we will omit the subscript I on the field operators. Any field operator
appearing in our expressions in this section will always be meant to be written in the
interaction picture. In subsequent sections, it should be clear from the context where
φ(x) denotes the Heisenberg picture operator and where it denotes the interaction picture
operator.
Let us now generalize eq. (4.55) to time-ordered products of more than two fields. The
result we are going to prove is called Wick’s theorem and reads
where “all possible contractions” means one term for each way of contracting pairs of
fields. For instance, for n = 4, Wick’s theorem reads (with the shorthand notation
φa ≡ φ(xa ))
T φ1 φ2 φ3 φ4 = : φ1 φ2 φ3 φ4 + φ1 φ2 φ3 φ4 + φ1 φ2 φ3 φ4 + φ1 φ2 φ3 φ4
+ φ1 φ2 φ3 φ4 + φ1 φ2 φ3 φ4 + φ1 φ2 φ3 φ4
+ φ1 φ2 φ3 φ4 + φ1 φ2 φ3 φ4 + φ1 φ2 φ3 φ4 : . (4.57)
52
4.3 Feynman Diagrams
This expression means that all operators that are not contracted are in normal order,
while each pair of contracted operators gives a factor DF , irrespective of whether they
are adjacent or not. For instance:
: T φ1 φ2 φ3 φ4 : ≡ DF (x1 − x3 ) · : φ2 φ4 : . (4.58)
Note that, in vacuum expectation values of the form h0|T φ(x1 ) · · · φ(xn )|0i, only those
terms where all field are contracted contribute.
Let us now prove Wick’s theorem by induction. We have already proven it for n = 2.
Assume now that the theorem has been proven for n − 1 fields. Let us moreover assume
without loss of generality that x01 ≥ x02 ≥ · · · ≥ x0n . (If this condition is not satisfied, we
can simply relabel the xj to fulfill it.) Then,
−
T φ(x1 ) · · · φ(xn ) = (φ+
1 + φ1 ) · : φ(x2 ) · · · φ(xn )
+ [all contractions not involving φ1 ] : . (4.59)
− −
We would like to pull φ+ 1 and φ1 into the normally ordered expression. For φ1 , this is
easy because, according to the normal ordering prescription, it is anyway supposed to
stay on the left. The φ+ 1 operator, on the other hand, has to be commuted past all other
uncontracted fields. (Remember that fields that are contracted give just c-numbers that
commute with everything.) In a term with m uncontracted fields (which we take to be
φ2 , . . . , φm for definiteness), this leads to
φ+ + +
1 : φ2 · · · φm : = : φ2 · · · φm : φ1 + [φ1 , : φ2 · · · φm :]
+ − + −
= : φ+
1 φ2 · · · φm : + : [φ1 , φ2 ]φ3 · · · φm + φ2 [φ1 , φ3 ]φ4 · · · φm + · · · :
= : φ+
1 φ2 · · · φm : +φ1 φ2 φ3 · · · φm + φ1 φ2 φ3 · · · φm + · · · . (4.60)
For m = n, the first term on the right hand side gives the first term on the right hand
side of Wick’s theorem, eq. (4.56). The other terms for n = m give all possible terms
involving a contractions of φ1 with one of the other fields, and no other contractions. For
m < n, we obtain all terms involving an uncontracted φ1 together with contractions of
other pairs of fields, and all terms involving a contracted φ1 , together with contractions
of other pairs of fields. These are all the terms appearing on the right hand side of Wick’s
theorem, eq. (4.56), and the theorem is therefore proven.
53
Chapter 4 Interacting Fields and Feynman Diagrams
3 4 3 4 3 4
This diagrammatic notation also admits a straightforward physical interpretation: par-
ticles are created at two of the vertices, then each particle propagates to another vertex
and is annihilated there. There are three different ways of picking pairs of vertices, cor-
responding to the three Feynman diagrams. Each diagram corresponds to a transition
amplitude, and the amplitude for the overall process is the sum of the three Feynman
diagrams.
Z T
0T φ(x)φ(y) exp − i dt Hint (t) 0
−T
D Z T E
= 0T φ(x)φ(y) + T φ(x)φ(y) − i dt Hint (t) + · · · 0 , (4.62)
−T
where on the right hand side, we have expanded the perturbation series up to first order
in λ. The zeroth order term is simple: h0|T φ(x)φ(y)|0i = DF (x − y). The first order
term reads, after plugging in the expression for Hint
Z
λ
d4 z 0T φ(x)φ(y)φ(z)φ(z)φ(z)φ(z)0 .
−i (4.63)
4!
2
It is implicit here that the field operators φ(x), φ(y) appearing hΩ|T φ(x)φ(y)|Ωi are the full Heisenberg
fields, while those appearing in eq. (4.62) are the interaction picture operators which we denoted by
φI when we first introduced them.
54
4.3 Feynman Diagrams
x y z + x y (4.65)
z
Note that neither the prefactors nor the integral over d4 z is explicitly written out in
the diagrammatic notation. It is implied that each
R 4 internal vertex (i.e. each vertex that
comes from Hint ) comes with a factor −iλ/4! d z. There is, however, no direct way
of reading off the prefactors 3 and 12 from the diagrams. When translating Feynman
diagrams into algebraic expressions, we therefore have to be careful not to forget these
factors. This can sometimes be a headache.
1 −iλ 3
Z Z Z
d z φφφφ d w φφφφ d4 u φφφφ 0
4 4
0 T φ(x)φ(y)
3! 4!
3 Z
1 −iλ
= d4 z d4 w d4 u DF (x − z) DF (y − w) DF (z − z)
3! 4!
· DF (z − w) [DF (w − u)]2 DF (u − u) . (4.66)
(4.67)
x z w y
3
It has become a sort of sport to name Feynman diagrams according to plants, animals and other
objects that they resemble. For instance: penguin diagram, lobster diagram, tadpole diagram, setting
sun diagram—just google for them.
55
Chapter 4 Interacting Fields and Feynman Diagrams
Of course, the contraction shown on the left hand side of eq. (4.66) is not the only
one leading to the expression on the right hand side. For instance, interchanging the
vertices z, w, u does not change the algebraic expression because the vertex coordinates
are all integrated over. There are 3! ways of interchanging the vertices. Moreover, the
contractions that involve φ(z) can connect to any of the four φ(z) factors. This leads to
a 4! = 24-fold ambiguity. Since, however, one contraction has its beginning and its end
on a φ(z) factor, the ambiguity is only 12-fold in our case. Similarly, the ambiguity in
how the contractions connect to the φ(w) fields is 24-fold and the one for the φ(u) fields
is 12-fold. Finally, note that ther are two contractions of the form DF (w − u):
φ(w)φ(w)φ(u)φ(u) . (4.68)
Interchanging the two φ(w) factors and the two φ(u) factors in this expression simul-
taneously does not change the contraction structure. Therefore, the multiplicity of the
diagram is reduced by a factor of 2. In total, the multiplicity factor is thus
4! 4! 1
3! × × 4! × × . (4.69)
|{z} 2 |{z} 2 2
interchange of |{z} interchange of |{z} |{z}
vertices interchange of φ(w) factors interchange of interchange of
φ(z) factors φ(u) factors the two
φ(w)–φ(u)
contractions
We see that the factor coming from the interchange of vertices cancels against the pref-
actor 1/3! from the Taylor series. This is true at any order in perturbation theory: the
n-th order term comes with a factor 1/n! from the series expansion, and interchanging
vertices leads to a multiplicity of n!. Also, each vertex has a prefactor 1/4! that cancels
against the factor 4! from the different ways in which contractions connect to the four
field operators in each vertex. (It now becomes clear why we have pulled the factor 1/4!
out of λ in defining the theory!) What remains is a factor 1/S = 1/8, called a symme-
try factor that accounts for permutations of vertices or fields that leave the contraction
structure invariant.
56
4.3 Feynman Diagrams
w u
(symmetry factor of the “setting
S = 3! · 2 sun ” diagram times factor 2 from (4.73)
x z y interchanging u and w vertices.)
Z T
hΩ|T φ(x1 ) · · · φ(xn )|Ωi ∝ 0T φI (x1 ) · · · φI (xn ) exp − i dt Hint (t) 0 (4.74)
−T
up to a given order p in perturbation theory, draw all possible Feynman diagrams with
n external vertices (each of which connects to exactly one propagator) and p internal
vertices (each of which has four connection points). The algebraic expression for each
diagram is then composed of factors determined by the following Feynman rules
Z
2. For each internal vertex z = −iλ d4 z (4.76)
Let us briefly reiterate the physical interpretation of these rules: each propagator gives
57
Chapter 4 Interacting Fields and Feynman Diagrams
the amplitude for a particle to propagate between two spacetime points. At each external
vertex, one particle is created or annihilated. At each internal vertex, four propagators
meet. This describes processes where 4 particles are created from the vacuum, or 1
particle is annihilated and 3 are created, or 2 are annihilated and 2 are created, or 3 are
annihilated and 1 is cretaed, or 4 are annihilated. The amplitude for these interaction
processes is given by the factor −iλ. Finally, the integral over d4 z instructs us to integrate
over all spacetime points where an interaction can happen. This is a manifestation of
the superposition principle in quantum mechanics. If a process can happen in different
ways (in this case: if the interaction can happen at different spacetime points z), the
amplitudes for all these different ways should be added up.
d4 p ie−ip(x−y)
Z
DF (x − y) = , (4.78)
(2π)4 p2 − m2 + i
58
4.3 Feynman Diagrams
p2 p4
Z
z = −iλ d4 ze−i(p1 +p2 +p3 +p4 )z = −iλ (2π)4 δ (4) (p1 + p2 − p3 − p4 ) .
p1 p3 (4.80)
If any of the arrows was reversed, the sign of the corresponding momentum would be
flipped on the right and side. The δ-function simply implies energy-momentum conser-
vation at the vertex: the 4-momentum carried into the vertex by the incoming particles
must be equal to the 4-momentum carried away by the outgoing particles.
The only pieces that are left are the 4-momentum integrals from the propagators.
Many of these can be carried out directly using the δ-functions from the vertices. The
remaining ones have to be included in the amplitude explicitly.
To summarize, the momentum space Feynman rules read
i
1. For each propagator = (4.81)
p2 − m2 + i
d4 p
Z
5. Integrate over momenta not determined by 4.: (4.84)
(2π)4
59
Chapter 4 Interacting Fields and Feynman Diagrams
x y
(4.85)
z
x1 x2
x3 x4
Each disconnected diagram is a product of several connected pieces. Some of these con-
tain external vertices, others contain only internal vertices. The latter type of diagrams
is called vacuum bubbles because they correspond to processes where particles are created
from the vacuum and annihilate back into the vacuum without coupling to an external
vertex.
Let us define the set {Vj } of all vacuum bubbles and the set {Ej } of diagrams in which
each internal vertex is connected to at least one external vertex. Note that the Ej do
not need to be connected—the only requirement is that each disconnected piece has to
contain at least one external vertex. In other words, there should be no vacuum bubbles.
Since every Feynman diagram is a product of one Ej and an arbitrary number of Vj ,
combinatorics tells us that the perturbation series can be reordered in the following way:
Z T i X X Y 1
h
ni
0 T φ(x1 ) · · · φ(xn ) exp − i
dt Hint,I (t) 0 =
Ej × (Vi ) .
−T ni !
j {ni } i
(4.86)
Here, the sum over {ni } runs over all ordered sets {n1 , n2 , · · · } of non-negative integers,
where each ni denotes how often the i-th vacuum bubble appears in the diagram. In any
fixed order Feynman diagram, only a finite number of the ni will be nonzero of course.
The factor 1/ni ! is a symmetry factor arising from the fact that interchanging identical
vacuum bubbles leaves the diagram and the underlying contraction structure unchanged.
We can simplify the contribution from the vacuum bubbles to eq. (4.86) further:
XY 1 YX ∞
ni 1
(Vi ) = (Vi )ni (4.87)
ni ! ni !
{ni } i i ni =0
Y
= exp Vi (4.88)
i
60
4.4 The LSZ Reduction Formula
X
= exp Vi . (4.89)
i
d3 p
Z
ap (t)eipx + a†p (t)e−ipx ,
φ(x, t) = 3
p (4.93)
(2π) 2Ep
the creation and annihilation operators a†p (t) and ap (t) acquire a time dependence. In
section 4.1.4, we have imposed the canonical commutation relations at t = t0 to obtain
an object whose action on the free vacuum |0i we know. However, this does not help us
61
Chapter 4 Interacting Fields and Feynman Diagrams
define one-particle states in the interacting theory because we do not know how a†p (t0 )
acts on |Ωi.
We will assume that, in the interacting field theory, asymptotic one-particle states
with definite momentum are given by
ap (±∞)|Ωi = 0 . (4.95)
We will assume that the one-particle states defined by eq. (4.94) are normalized in the
same way as in the free field theory:
A further complication arises because the initial state of a scattering process involves
more than one particle. As soon as the wave functions of these particles overlap, they
will start to interact. If particles were described by plane wave states (momentum eigen-
states), their wave functions would overlap at all times. This would make it difficult to
define suitable initial and final states as combinations of one-particle states. In partic-
ular, even tough one particle states a†p1 (±∞)|Ωi and a†p2 (t = ±∞)|Ωi are momentum
eigenstates, the two particle state a†p2 a†p1 (±∞)|Ωi is not a momentum eigenstate because
of the interaction. The way out is to consider localized wave packets instead of plane
waves. Then, at t = ±∞, the wave packets will have negligible overlap. The different
particles in the initial and final states will know of each other only when they get close
to the origin, so their interaction does not hinder the definition of the asymptotic states.
We define the creation operator for a wave packet state centered around a momentum
pj as
d3 p
Z
ã†j (t) ≡ f (p; pj ) a†p (t) , (4.97)
(2π)3
where f (p; pj ) is the shape factor of the wave packet. Its exact form is unimportant,
but we could for instance choose a Gaussian:
(2π)3/4 (p − pj )2
f (p; pj ) = exp − . (4.98)
σ 3/2 4σ 2
(The normalization is chosen such that hΩ|ãj ã†j |Ωi = 1.) To be specific, let us consider
a 2 → 2 scattering process in which two initial state wave packets with central momenta
62
4.4 The LSZ Reduction Formula
p1 and p2 scatter into two final state wave packets with central momenta p3 and p4 .
The initial and final states are thus
and
←→
where f (x) ∂0 g(x) ≡ f (x)(∂g(x)/∂x0 ) − (∂f (x)/∂x0 )g(x). The construction with the
derivative is necessary to extract only a†p (t) instead of a combination of a†p (t) and ap (t),
as can be easily shown by direct computation:
←→
Z
i
−p d3 x e−ipx ∂0 φ(x)
2Ep t→±∞
3
−ipx ←
→
Z Z
i d q −iqx
† iqx
= −p d3 x p e ∂0 aq e + aq e
(2π)3 2Eq
2Ep t→±∞
i −iEp t
=− e − iEp ap e−iEp t + iEp a†p eiEp t + iEp ap e−iEp t + iEp a†p eiEp t
2Ep t→±∞
Note that here we have expanded the full Heisenberg field φ(x) in the same way as in
the free theory. This is not possible in general—the general time dependence of the
field operator is more complicated than that. However, by our assumption, at t → ±∞,
the operators a†p (t → ±∞) and ap (t → ±∞) create and annihilate energy-momentum
eigenstates of the full theory. This means that they commute with the Hamiltonian at
t → ±∞, and consequently, according to eq. (4.8) (or according to the derivation from
section 2.3.3) the time evolution of the field operator at t → ±∞ is indeed as in the free
theory.
To rewrite a scattering amplitude, it is moreover useful to be able to rewrite ã†j (−∞)
in terms of ã†j (+∞) and vice versa. Here is the relevant relation:
Z ∞
ã†j (+∞) − ã†j (−∞) = dt ∂0 ã†j (t)
−∞
63
Chapter 4 Interacting Fields and Feynman Diagrams
d3 p −i ←
→
Z Z
= f (p; pj ) d4 x p ∂0 e−ipx ∂0 φ(x)
(2π)3 2Ep | {z }
←
→ ←
→
= −ip0 [e−ipx ∂0 φ(x)] + e−ipx ∂0 (∂0 φ(x))
= e−ipx [(p0 )2 + ∂02 ]φ(x)
d3 p −i −ipx 2
Z Z
d4 x p ∂0 + p2 + m2 φ(x)
= f (p; pj ) e
(2π)3 2Ep
d3 p −i −ipx 2 ← −
Z Z
f (p; pj ) d4 x p ∂0 − ∇2 + m2 φ(x)
= e
(2π)3 2Ep
d3 p −i −ipx 2
Z Z
f (p; pj ) d4 x p ∂ + m2 φ(x) .
= e (4.103)
(2π)3 2Ep
In the last step, we have integrated by parts twice to turn the left-acting derivative
←
−
operator ∇2 into a right-acting operator ∇2 . Note that (∂ 2 + m2 φ(x) would vanish in
the free theory due to the Klein–Gordon equation. In the interacting theory, however,
this expression is in general nonzero.
On the right hand side of eq. (4.103), the shape factor f (p, p0 ) of the wave packet no
longer plays an important (we only needed it to give meaning to the asymptotic states).
Therefore, we will from now on assume the momentum distribution q of the wave packets
3
to be very narrow, so that we can approximate f (p, pj ) ' (2π) 2Epj δ (3) (p−pj ). (The
normalization here is such that the states ã†j |Ωi are normalized in the same way as the
states |pj i, see eq. (4.96).
We can now go back to the scattering amplitude and write
hf |ii = hΩ|T ã4 (+∞) ã3 (+∞) ã†1 (−∞) ã†2 (−∞)|Ωi . (4.104)
On the right hand side, we have added a time ordering symbol T . We can do this because
the operators that follow it are in time order already. We now use eq. (4.103) and its
hermitian conjugate to turn all ãj (+∞) operators into ãj (−∞) operators and all ã†j (−∞)
into ã†j (+∞). The time ordering symbol then moves the annihilation operators to the
right, where they act on the vacuum to yield zero. What is left is only a term containing
field operators, but no ãj or ã†j operators any more. This term reads
Z
hf |ii = d4 x1 d4 x2 d4 x3 d4 x4 e−ip1 x1 (∂x21 + m2 ) · e−ip2 x2 (∂x22 + m2 )
· eip3 x3 (∂x23 + m2 ) eip4 x4 (∂x24 + m2 ) hΩ|T φ(x1 )φ(x2 )φ(x3 )φ(x4 )|Ωi . (4.105)
This is the desired relation between a scattering amplitude and a correlation function.
It is called the Lehmann-Symanzik-Zimmermann (LSZ) reduction formula. Generalizing
it from the 2 → 2 case to an arbitrary process with n particles in the initial state
(momenta p1 , . . . , pn ) and m particles in the final state (momenta pn+1 , . . . , pn+m ) is
64
4.5 Computing S-Matrix Elements from Feynman Diagrams
straightforward:
Z
n+m
hf |ii = i d4 x1 · · · d4 xn+m
Let us finally reiterate the main assumptions that went into the LSZ reduction formula.
The formula hinges on the fact that a†p (±∞)|Ωi is indeed a one-particle state. This can
only be the case if
hΩ|φ(x)|Ωi = 0 , (4.107)
i.e. if the vacuum expectation value of φ(x) vanishes. If eq. (4.107) was not satisfied, this
would mean that one or several of the a†p create a superposition of a one-particle state
and the vacuum.
The LSZ formula also relied on the normalization of the one-particle states, see eq. (4.96).
If the normalization was different, the numerical factors in the LSZ formula would change.
The normalization condition eq. (4.96) implies
65
Chapter 4 Interacting Fields and Feynman Diagrams
4-momentum p and a factor e−ipxj for an outgoing particle (arrow pointing out of the
diagram towards the external vertex) with 4-momentum p. This factor is the only depen-
dence of the correlation function on xj . Therefore, applying the Klein–Gordon operator
∂x2j +m2 from the LSZ formula to these exponential factors yields simply (−p2 +m2 )e±ipxj .
The term in parentheses (together with one of the factors of i in the LSZ formula) cancels
exactly the i/(p2 − m2 + i) from the propagator connected to that external vertex.
The remaining exponential e±i(pj −p)xj , together with the integral over d4 xj , yields a
delta function (2π)4 δ (4) (pjR − p). These delta function can be used to evaluate some of
the momentum integrals d4 p/(2π)4 . Only one of them remains, enforcing overall 4-
momentum conservation among the external particles: (2π)4 δ (4) (±p1 ± p2 ± p3 ± · · · ),
where the plus signs stand for incoming particles and the minus signs for outgoing par-
ticles (or vice-versa). In practice, this delta function is usually not associated with the
matrix element, but is reintroduced in the formula for the cross section. We will follow
this convention here.
Beyond the overall delta function, neither external vertices nor propagators connected
to them contribute to the Feynman diagrams for the scattering amplitude. Propagator
factors need to be included only for internal propagators, i.e. propagators connecting
two internal vertices. To emphasize this point, incoming and outgoing particles in the
Feynman diagram for a scattering amplitude are depicted as lines with only one end
attached to a vertex.
These rules can be summarized as follows
i
1. For each (internal) propagator = (4.109)
p2 − m2 + i
d4 p
Z
5. Integrate over momenta not determined by 4.: (4.112)
(2π)4
66
4.5 Computing S-Matrix Elements from Feynman Diagrams
Here are two examples: first, the amplitude for the scattering of a φ particle off another
φ particle a lowest order in λ is simply
p2 p4
= −iλ (4.113)
p1 p3
Here pj are the momenta of the external particles. It is implied that p1 + p2 = p3 + p4 .
Going one order higher in λ, there are two diagrams:
p2 p4 p2 p4
p1 p3 p1 p3
d4 q
Z
2 i i
= (−iλ) · 2
(2π) (p1 + p2 + q) − m + i q − m2 + i
4 2 2
d4 q
Z
2 i i
+ (−iλ) · 2 . (4.114)
(2π) (p1 − p3 + q) − m + i q − m2 + i
4 2 2
We call diagrams like eq. (4.113), where no path taken through the diagrams hits a vertex
twice, a tree level diagram. Diagrams with closed loops, i.e. paths that return to the
vertex at which they originated are called loop diagrams. We will develop the technology
to evaluate loop diagrams, i.e. to simplify algebraic expressions like the right hand side
of eq. (4.114), in the second part of this course.
In principle, when considering the second order (in λ) corrections to the scattering rate
of two φ particles, one could also draw diagrams with loops connected to the external
legs, for instance
(4.115)
We will see later, when discussing renormalization, that such diagrams in which loops
are attached to one of the external legs, need not be considered in many cases. For
the moment, we simply accept this result and will therefore drop such contributions
henceforth.
A few comments are in order:
• Note that we have omitted the arrows on the lines. Since the only place where the
4-momenta appear is in the propagators, where they appear as p2 , the direction of
momentum flow is irrelevant.
67
Chapter 4 Interacting Fields and Feynman Diagrams
+ + , (4.116)
• In a typical Feynman diagram, most momenta are determined by the fixed momenta
of the external particle, together with 4-momentum conservation at the vertices.
The undetermined momenta that we have to integrate over according to rule 5.
correspond to momentum flows inside the diagram along closed loops, i.e. beginning
and ending on the same vertex. For instance, in the diagram
(4.117)
there are two such loop momenta. To see this, note that there are three internal
propagators with per se undetermined momenta that should be integrated over.
Momentum conservation at the vertices imposes two constraints (in the form of
4d δ-functions). One of them becomes the overall energy-momentum-conserving
δ-factor that we have pulled out of the matrix element, the other eliminates the
integral over one of the internal momenta. We are left with two momentum inte-
grations, and we say that the diagram is a 2-loop diagram.
68
4.6 Feynman Rules for Fermions
Rt Rt
to bring them into time order (see eq. (4.21), where we rewrote t0 dt1 t01 dt2 in terms
Rt Rt
of t0 dt1 t0 dt2 and the last step leading to the master formula eq. (4.40), where we
pulled the various U -factors together into one exponential). For fermions, reordering
field operators leads to extra minus signs due to their anticommuting nature. However,
these minus signs can be fully accounted for if we define the time ordering symbol T such
that any interchange of two fermion fields required to restore time ordering yields one
extra minus sign. For instance:
T ψ1 ψ̄2 ψ3 ψ4 = (−1)3 ψ3 ψ1 ψ4 ψ̄2 if x03 > x01 > x04 > x02 . (4.118)
– – ( ) ( )
Here, we have again used the shorthand notation ψ j ≡ ψ (xj ). For the special case of
only two fermion fields, we have already used this definition in section 3.3.
With this definition, we can easily generalize Wick’s theorem. Consider first again the
case of just two fermion fields (cf. eq. (4.48) et seqq.):
Remember that the contraction term arises from those contributions that require an
interchange of fields in order to bring them into normal order. For scalar fields, this was
achieved by a commutator, for fermion fields we should use the anticommutator instead.
Since the time ordering symbol and the normal ordering symbols imply an extra minus
sign if y 0 > x0 , also the definition of the contraction has to include such an extra minus
sign:
(
{ψ + (x), ψ̄ − (y)} for x0 ≥ y 0
ψ(x)ψ̄(y) ≡ (4.121)
−{ψ̄ + (y), ψ − (x)} for x0 < y 0
= SF (x − y) . (4.122)
Here, ψ + (x) and ψ − (x) are again defined as the positive and negative frequency com-
ponents of ψ(x), respectively. The fact that the contraction of ψ(x) and ψ̄(y) equals
SF (x − y) is completely analogous to the proof of the corresponding identity for scalar
fields, eq. (4.54). Note that, unlike for the scalar propagator, for fermions SF (x − y) 6=
SF (y − x), so care must be taken to put ψ(x) and ψ̄(y) into the appropriate order before
evaluating the contraction.
69
Chapter 4 Interacting Fields and Feynman Diagrams
The time ordered product of two ψ fields or two ψ̄ fields is equal to the normally
ordered product of these fields because their anticommutator vanishes. In order for our
equations to have the same form in this case as in eq. (4.120), we define
ψ(x)ψ(y) = 0 (4.123)
ψ̄(x)ψ̄(y) = 0 . (4.124)
To carry out the induction step in the proof of Wick’s theorem for fermions, we need
to worry about one more source of minus signs. The analogue of eq. (4.60) for fermions
reads
= : ψ1+ ψ2 · · · ψm : + : ψ1 ψ2 ψ3 · · · ψm : + : ψ1 ψ2 ψ3 · · · ψm : + · · · .
(4.125)
(Replacing any of the ψ fields by ψ̄ fields would not change this result.) This implies
that a contraction of two non-adjacent field operators comes with a minus sign if an
odd number of anticommutations is necessary to move these field operators next to each
other. For example:
To summarize, Wick’s theorem for fermions takes exactly the same form as for scalars:
For fermions, however, it is understood that an extra minus sign appears on the left hand
side when x1 , . . . , xn are such that bringing the field operators into time order requires
an odd permutation. Similarly, in the normally ordered terms on the right hand side,
any summand that requires an odd permutation to restore normal order gets a minus
sign. Finally in replacing contracted field operators by Feynman propagators, we should
first move the contracted fields next to each other, possible paying the price of another
minus sign if an odd number of anticommutations is required.
70
4.6 Feynman Rules for Fermions
· eipn+1 xn+1 ūsn+1 (pn+1 )(i∂/xn+1 − m) αn+1 · · · eipn+m xn+m ūsn+m (pn+m )(i∂/xn+m − m) αn+m
· hΩ|T ψαn+m (xn+m ) · · · ψαn+1 (xn+1 )ψ̄(x1 )α1 · · · ψ̄(xn )αn |Ωi
←− ←−
· (−i ∂/ x1 − m)us1 (p1 ) α1 e−ip1 x1 · · · (−i ∂/ xn − m)usn (pn ) αn e−ipn xn .
(4.128)
Here, α1 , . . . αn+m are spinor indices. If any of the incoming fermions is replaced by an
antifermion (created by ψ instead of ψ̄), the corresponding factor is replaced according
to
←−
−i(−i ∂/ xj − m)usj (pj )e−ipj xj → ie−ipj xj v̄ sj (pj )(i∂/xj − m) (4.129)
and is moved from the right of the correlation function to the left. Similarly, if an outgoing
fermion is replaced by an antifermion (annihilated by ψ̄ instead of ψ), the corresponding
factor is replaced according to
←−
−ieipj xj ūsj (pj )(i∂/xj − m) → i(−i ∂/ xj − m)v sj (pj )eipj xj (4.130)
and is moved from the left of the correlation function to the right.
There are two main differences between the LSZ formula for fermions compared to
the one for scalars: first, the replacement of the Klein–Gordon operator by the Dirac
operator comes about because the analogue of eq. (4.101) which allows us to write a
creation operator in terms of a field operator does not involve a derivative in the case of
fermions. Instead, one exploits the orthogonality relations of the u and v spinors, which
also explains the appearance of such spinors in the LSZ formula for fermions.
71
Chapter 4 Interacting Fields and Feynman Diagrams
Z Z
(−ig)2
4
d4 y ψ̄ρ (y)ψρ (y)φ(y) 0 .
2! 0 T ψα (x1 )ψβ (x2 )ψ̄γ (x3 )ψ̄δ (x4 ) d x ψ̄τ (x)ψτ (x)φ(x)
(4.134)
Contracting ψα (x1 ) with ψ̄τ (x) and ψβ (x2 ) with ψ̄ρ (y) instead yields another nonzero
contribution. As usual, interchanging the two vertices x and y is always possible, and
the resulting factor 2! cancels the 1/2! prefactor. Therefore we can drop this prefactor
from now on. Note that, in Yukawa theory, the vertex couples three different fields, so
no extra factors arise from interchanging those. In other words, Yukawa theory does not
have symmetry factors!. Up to a possible minus sign, eq. (4.134) yields
Z Z
2
(−ig) d x d4 y [SF (x1 − y)]αρ [SF (x2 − x)]βτ [SF (y − x3 )]ργ [SF (x − x4 )]τ δ DF (x − y) .
4
(4.135)
We now apply the LSZ reduction formula to this expression. The terms for the incoming
fermions have the structure
←
−
Z
− i d4 x3 SF (y − x3 ) (−i ∂/ x3 − mψ )us3 (p)e−ipx3
d4 q i(/q + mψ ) −iq(y−x3 ) ←−
Z Z
= −i d x34
4 2 2 e (−i ∂/ x3 − mψ )us3 (p)e−ipx3
(2π) q − mψ + i
d4 q (/q + mψ )(/q − mψ ) −iq(y−x3 ) s3
Z Z
= d4 x3 e u (p)e−ipx3
(2π)4 q 2 − m2ψ + i
| {z }
=1
−ipy s3
=e u (p) . (4.136)
0
Similarly, for each outgoing fermion we obtain a factor of the form e+ip y ūs1 (p0 ) or
0
e+ik x ūs2 (k 0 ). The overall expression for the matrix element is thus, after plugging in
the Fourier expansion of DF (x − y),
iM · (2π)4 δ (4) (p + k − p0 + k 0 )
72
4.6 Feynman Rules for Fermions
d4 q ie−iq(x−y)
Z
0 0
= (−ig)2 [ūs1 (p0 )us3 (p)][ūs2 (k 0 )us4 (k)]e−i(p−p )y e−i(k−k )x
(2π)4 q 2 − m2φ + i
i
= (−ig)2 [ūs1 (p0 )us3 (p)][ūs2 (k 0 )us4 (k)] · (2π)4 δ (4) (p + k − p0 + k 0 )
(p0 − p)2 − m2φ
(4.137)
As mentioned earlier, we pull the overall 4-momentum conserving delta function out of
the matrix element by convention. The procedure we have followed can obvious also be
applied to more general processes, and to higher order terms in the perturbation series.
What we have found can be summarized in the Feynman rules for Yukawa theory:
73
Chapter 4 Interacting Fields and Feynman Diagrams
p
−−→
2. Incoming antifermion = v̄ s (p) (4.139)
p
−−→
4. Outgoing antifermion = v s (p) (4.141)
i
7. Scalar propagator = (4.144)
p p2 − m2φ + i
i(p
/ + mψ )
8. Fermion propagator = (4.145)
p p2 − m2ψ + i
d4 p
Z
11. Integrate over momenta not determined by 10.: (4.147)
(2π)4
74
4.6 Feynman Rules for Fermions
As mentioned above, there are no symmetry factors in Yukawa theory. With these rules,
the scattering amplitude we just computed corresponds to the diagram
p p0
q
k k0
(4.148)
and the second diagram relevant for the process of fermion–fermion scattering is
p0
p
q
(4.149)
k
k0
• By convention, the arrows on the fermion lines do not represent momentum flow,
but particle number flow: particle number flows into the diagram along an in-
coming fermion line and out of the diagram along an outgoing fermion line. For
antiparticles, the flow is reversed: for each incoming antifermion, one unit of par-
ticle number flows out of the diagram (or, equivalently, a unit of negative particle
number flows into the diagram) and for outgoing antifermions, particle number
flows into the diagram. On the external lines, the arrows remind us whether the
line corresponds to a barred spinor or an unbarred spinor. At the vertices, they
remind us of which propagator is contracted with the ψ̄ factor from the vertex and
which one is contracted with the ψ factor. The arrows thus also help us avoid
mistakes: if we end up with a vertex that has two arrows pointing towards it or
two arrows pointing away from it, that diagram is forbidden.
• Spinor indices are always contracted along fermion lines. This can be understood
by noting that fermion lines are joint together at vertices, where the indices of
the ψ field corresponding to the line pointing towards the vertex and the ψ̄ field
corresponding to the line pointing away from the vertex are contracted. In practice,
when translating a Feynman diagram to an algebraic expression, we start at the end
75
Chapter 4 Interacting Fields and Feynman Diagrams
of a fermion lines (which gives ū or v̄ spinor) and then work our way backwards
along the lines, in the direction opposite to the arrows, applying the rules for
vertices and propagators as we encounter them.
• For fermion propagators, the direction of momentum flow matters. The above rule
holds if the momentum p flows along the arrow. (In case the momentum flows in
the opposite direction p should be replaced by −p.) This rule arise from the way the
oscillating exponential combine into 4-momentum conserving delta functions. To
see this explicitly, let us work out another example from first principles. Consider
the process of fermion–antifermion annihilation into two scalars:
p k
x→ −
q
p0 k0
−
→ →
−
y (4.150)
Z Z
0 T ψ̄(x1 )ψ(x2 )φ(x3 )φ(x4 )(−ig) d4 x ψ̄ψφ (−ig) d4 y ψ̄ψφ 0
Z Z
= (−ig)2 d4 x d4 y (−1)5 SF (x−x1 )SF (x2 −y)SF (y−x)DF (x−x4 )DF (y−x3 ) ,
(4.151)
iM · (2π)4 δ (4) (p + p0 − k − k 0 )
d4 q 0 i(/
q + mψ )
Z Z Z
0 0
= (−ig) 2 4
d x d y4
4
v̄(p ) 2 2 u(p) e−i(p−k−q)x e−i(p −k +q)y .
(2π) q − mψ
(4.152)
We see that the d4 x and d4 y integrals give delta functions that require the momen-
tum in the internal fermion propagator to flow from x to y, i.e. along the arrow.
• Perhaps the least useful Feynman rule is #12 (figure out the overall sign) because
it doesn’t really tell us what to do. In the above example of fermion–antifermion
annihilation, we have already seen what is needed: we have to count the number of
76
4.6 Feynman Rules for Fermions
p0
p p0 p
+ (4.153)
k k0
k
k0
We repeat here for convenience the contractions from eq. (4.134), corresponding to
the first diagram:
Z Z
4
0 T ψ(x1 )ψ(x2 )ψ̄(x3 )ψ̄(x4 ) 2!1 (−ig) d4 y ψ̄ψφ 0 . (4.154)
d x ψ̄ψφ (−ig)
As it is written here, this contraction structure does not come with an extra minus
sign. The contractions corresponding to the second diagram, however, are
Z Z
4
0 T ψ(x1 )ψ(x2 )ψ̄(x3 )ψ̄(x4 ) 2!1 (−ig) d4 y ψ̄ψφ 0 . (4.155)
d x ψ̄ψφ (−ig)
Therefore, when applying the Feynman rules to write down the expression for the
amplitude of fermion–fermion scattering, the two diagrams should come with a
relative minus sign:
i
iM = (−ig) [ūs1 (p0 )us3 (p)] 0
2
[ūs2 (k 0 )us4 (k)]
(p − p)2 − m2φ
77
Chapter 4 Interacting Fields and Feynman Diagrams
s1 0 s4 i s2 0 s3
− [ū (p )u (k)] 0 [ū (k )u (p)] . (4.156)
(k − p)2 − m2φ
One special case is that of closed fermion loops. Consider for instance the leading
order diagram contributing to the scattering of two scalars:
We see that a closed fermion loop in a diagram always leads to an extra minus
sign.
78
4.6 Feynman Rules for Fermions
where, by assumption, the masses of the two fermions satisfy mψ2 mψ1 . The scattering
amplitude is as usual
The subscripts ±T on the bra and ket vectors indicate that the states are taken a
t = T → ±∞ and then evolved in time with the Hamiltonian H = H0 + V (t, x). In the
second scalar product in eq. (4.162), which comes from the very heavy fermion, we have
taken into account that this fermion is too heavy to be affected by the scattering process,
so the potential can be neglected in its Hamiltonian. Note also that, for very heavy mψ2 ,
the particle can absorb any momentum without its state changing. Therefore, we can
without loss of generality replace hk0 | by hk0 − k − p|. (We do this in order to bring the
final form of hf |ii to the same form as eq. (4.161), see below.)
The first factor on the right hand side of eq. (4.162) can be evaluated perturbatively
by expanding the exponential in V (t, x). The zeroth order term in this expansion is
uninteresting because no momentum is exchanged. We thus focus on the first order
term. Assuming that the potential is time-independent, it yields
Z RT 0 Rt 0
lim +T hp | − i dt e−i t dt H0 V (x)e−i −T dt H0 |pi−T
0
(4.163)
T →∞
Z Z
0
= lim −i dt d3 x e−i(Ep −Ep0 )t+i(p−p )x V (x) (4.164)
T →∞
In the second line, we have used that |pi and |p0 i are eigenstates of the unperturbed
Hamiltonian H0 , and we have replaced the bra and ket vectors by the (time-dependent)
wave functions of the initial and final states. In the third line, we have introduced the
3-dimensional Fourier transform Ṽ (p0 − p) of the potential V (x). In the fourth line, we
have rewritten the δ-function to bring it to the form appearing in the QFT expression
eq. (4.161). In doing so, we have used that mψ2 is very large, so that Ek − Ek0 ' 0.
After a similar calculation for the second term in eq. (4.162), we obtain
Before directly comparing eq. (4.167) to the QFT expressions eqs. (4.160) and (4.161),
we need to remember that in QM a one-particle state |pi is usually normalized according
79
Chapter 4 Interacting Fields and Feynman Diagrams
−g 2
Ṽ (p0 − p) = . (4.168)
|p0 − p|2 + m2φ
All that’s left to do to obtain V (r) is to carry out an inverse Fourier transform:
d3 q −g 2
Z
V (x) = eiqx
(2π)3 p2 + m2φ
−g 2 ∞ iqr − e−iqr
Z
2 e 1
= dq q
2
4π 0 iqr q + m2φ
2
∞
−g 2 qeiqr
Z
= 2 dq 2 (4.169)
4π ir −∞ q + m2φ
This integral can be carried out using the residual theorem. The integration contour
can be closed in the complex q-plane by an infinite half-circle in the upper half-plane.
Thanks to the exponential, the integrand vanishes along that half-circle. We then pick
up the pole at q = +imφ and get
−g 2 imφ e−rmφ
V (x) = · 2πi
4π 2 ir 2imφ
g 2 1 −rmφ
=− e , (4.170)
4π r
where r ≡ |x|. The minus sign indicates that the Yukawa interaction leads to an attractive
force. The exponential limits the range of the force to distances . m−1
φ . For instance, for
nucleon–nucleon interactions mediated by pions, the range of the interaction is ' m−1π ∼
−1
(100 MeV) ∼ O(fm).
80
Quantum Electrodynamics
5
After lots of formal developments, this section will be somewhat more practical. We will
abandon the toy models considered so far and move on to a theory of great practical
relevance: quantum electrodynamics (QED). In the context of QED, we will develop the
technology to efficiently calculate tree level Feynman diagrams and the associated cross
sections and decay rates.
We demand that the theory of electrons be invariant under U(1) gauge transformations
of the form
where α(x) is an arbitrary function of x. The crucial point here is the x-dependence of α.
This is what makes the symmetry a gauge symmetry or local symmetry, as opposed to a
global symmetry, for which the transformation law is the same at each spacetime point.
The symmetry is called a U (1) gauge symmetry because, at fixed x, the set of all possible
transformations forms a mathematical group called U (1) (for “unitary transformations
in one dimension”).
81
Chapter 5 Quantum Electrodynamics
Obviously, the free Dirac Lagrangian eq. (5.1) is not invariant under U (1) gauge trans-
formations because of the derivative. Rather, Lψ transforms as
Gauge invariance could be achieved if the theory contained another field Aµ (judiciously
called the photon) that couples to ψ through a term
F µν ≡ ∂ µ Aν − ∂ ν Aµ . (5.7)
It is easy to check that, with the Lagrangian (5.6), the Euler–Lagrange equations lead
to Maxwell’s equations:
δLγ
0 = ∂µ
δ(∂ µ Aν )
1
There are massive gauge bosons ins nature—the W and Z bosons mediating the weak interaction.
Understanding how they obtain their mass through the Higgs mechanism will be one of the climaxes
of the second part of this course.
82
5.2 The Feynman Rules for QED
δ
= ∂µ
ρ σ σ ρ
∂ A − ∂ A ∂ ρ A σ − ∂σ A ρ
δ(∂ µ Aν )
δ
= ∂µ
ρ σ
µ ν
2(∂ A )(∂ρ Aσ ) − 2(∂ ρ Aσ )(∂σ Aρ ) (5.8)
δ(∂ A )
= ∂ µ 4∂µ Aν − 4∂ν Aµ
= 4 ∂ µ Fµν . (5.9)
This is just Maxwell’s first equation in covariant form: ∂ µ Fµν = 0.2 In summary, the
QED Lagrangian reads in total
1
LQED = ψ̄ i∂/ − m ψ − Fµν F µν − eψ̄γ µ ψAµ .
(5.10)
4
As a final remark related to considerations of gauge symmetry, let us apply Noether’s
theorem. The consdeved current is
δLQED
jµ = · ∆ψ ∝ ψ̄γ µ ψ . (5.11)
δ(∂µ ψ) |{z}
=−iαψ
d3 p
Z X
s s −ipx s† s∗ ipx
Aµ (x) = p ap µ (p)e + ap µ (p)e . (5.12)
(2π)3 2Ep s
We have here exploited the fact that Aµ is a real field. Otherwise, the coefficients of
the positive and negative frequency components would not be the complex conjugates of
each other. Since Aµ is bosonic, the creation and annihilation operators as† s
p and ap obey
canonical commutation relations of the form of eq. (2.45).
The numerical coefficients sµ account for the Lorentz structure of the field and will
be interpreted as photon polarization vectors. This is in full analogy to the u and v
spinors in the expansion of the Dirac field. The u and v spinors were solutions to the free
Dirac equation, and the field was expanded in these solutions. Here, the sµ are solutions
2
Note that the second of Maxwell’s equations, ∂ µ (µνρσ F ρσ ) = 0, is automatically satisfied because
∂ µ F ρσ + ∂ ρ F σµ + ∂ σ F µρ = 0 .
83
Chapter 5 Quantum Electrodynamics
to the free Maxwell exquations. The sum runs over a complete basis of such solutions.
This is actually where the complications with quantizing the photon field come into play:
we know that physical photons are always transversely polarized. Therefore, we must
restrict the polarization vectors of external photons to have the form µ
µ
√ = (0, ), with
· p = 0. A suitable choice for p along the z axis is = (0, 1, ±i, 0)/ 2, where the plus
sign corresponds to right-handed polarization and the minus sign corresponds to left-
handed polarizition. With this restriction, we can guess the Feynman rules for incoming
(outgoing) photons: they will contribute just a factor µ (µ∗ ), just as the appearance
of u and v spinors in the expansion of the Dirac field lead to the appearance of these
spinors in the Feynman rules.
The restriction of transversality does not necessarily need to apply to off-shell photons
(photons with p2 6= 0) which can propagate along an internal photon line in a Feynman
diagram. Nevertheless, even off-shell photons do not have 4 degrees of freedom because
of gauge invariance. The freedom to choose a gauge removes one degree of freedom.
This has to be carefully taken into account when deriving the expression for the photon
propagator. Here, we will settle for a naı̈ve guesstimate: the propagator should somehow
resemble that of the scalar field. However, since the propagator should again be just a
two point correlation function of the form h0|Aµ (x)Aν (y)|0i, it must have some Lorentz
structure. The simplest choice is a factor g µν . We will settle for that, but we will add
an extra minus sign and define the photon propagator as
d4 p −ig µν e−ip(x−y)
Z
µν
∆ (x − y) ≡ . (5.13)
(2π)4 p2 + i
The rationale for the extra minus sign in the numerator is as follows: For x → y and µ =
ν, the propagator is just the norm of Aµ (x), and that norm should be positive. Evaluating
the integral over p0 in eq. (5.13) using complex contour integration and assuming x0 > y 0 ,
we have
d3 p 1 −ip(x−y)
Z
µν x0 &y 0
∆ (x − y) = e · (−g µν )
(2π)3 2|p|
−g µν |p|2 ei|p||x−y| − e−i|p||x−y|
Z
= d|p|
4π 2 2|p| i|p||x − y|
−g µν
= . (5.14)
8π|x − y|
Thus, the spatial components of Aµ will have positive norm, while the 0-component has
negative form. This means that at least physical (transversely polarized) photons are
properly normalized. Regarding the unphysical time-like photons, we do not have the
tools yet to deal with them. We will see later that their contributions exactly cancel
against contributions from longitudinal photons (polarized along p), so that only the
physical polarization states remain.
The Feynman rule for the interaction vertex between photons and fermions can be
directly read off from the Lagrangian. As in the previously considered theories, all
84
5.2 The Feynman Rules for QED
fields appearing in the interaction term end up contracted with other field operators,
yielding propagators, and the coupling constant (which here includes a γ µ matrix is
copied verbatim, but receives an extra factor
R of i from the fact that
R the master formula
eq. (4.40) involves an exponential of −i Hint (or equivalently +i Lint ), not just Lint .
We can thus summarize the QED Feynman rules as follows:
−ig µν
7. Photon propagator = (5.21)
p p2 + i
i(p
/ + m)
8. Fermion propagator = (5.22)
p p2 − m2 + i
85
Chapter 5 Quantum Electrodynamics
5.3 e+ e− → µ+ µ−
We now apply the QED Feynman rules to a simple process: the annihilation of an electron
and a positron into a muon and an antimuon, e+ e− → µ+ µ− . This was for instance
one of the most important processes at the Large Electron Positron (LEP) collider at
CERN. Similar processes (with the electron and the positron replaced by a quark and an
antiquark) are of utmost importance to the LHC. Our first goal is to compute the cross
section for e+ e− → µ+ µ− , developing along the way several important computational
techniques that are crucial to the efficient evaluation of Feynman diagrams.
p k
0 −igµν 0
= v̄ s (p0 )(−ieγ µ )us (p) ūr (k)(−ieγ ν )v r (k 0 ) . (5.25)
q q2
→
→
−
−
e+ p0 k0 µ+
Since our goal is to compute a cross section, we ultimately need not the matrix element
iM, but its modulus square, |M|2 . We thus have to multiply eq. (5.25) by its complex
conjugate. Along the way, we encounter the complex conjugates of Dirac field bilinearies,
e.g.
0 0
(v̄ s (p0 )γ µ us (p))∗ = [us (p)]† (γ µ )† (γ 0 )† v s (p0 )
0
= ūs (p)γ µ v s (p0 ) . (5.26)
γ µ† = γ 0 γ µ γ 0 , (5.27)
which is most easily shown by direct calculation using the explicit form eq. (3.4) of the
gamma matrices. The squared matrix element is therefore
e 4 s0 0 µ s ν s0 0
r0 0 r0 0
|M|2 = v̄ (p )γ u (p)ūs
(p)γ v (p ) ūr
(k)γ µ v (k )v̄ (k )γ ν ur
(k)
q4
e4 0 0
0 0
= 4 tr v s (p0 )v̄ s (p0 )γ µ us (p)ūs (p)γ ν tr ur (k)ūr (k)γµ v r (k 0 )v̄ r (k 0 )γν .
q
(5.28)
86
5.3 e+ e− → µ+ µ−
The rearrangements in the second line will come in handy below. In practice, we often
have no control over the spin orientation of the initial state particles, and we do not
measure the spin orientation of the final state particles. For instance, in a particle collider,
it is very difficult to produce polarized beams—it requires that the electrons have some
polarizing interaction and to retain their polarization until they collide. One possibility
is the so-called Sokolov–Ternov effect: when electrons emit synchrotron radiation in a
magnetic field, their spin flips. Since the two spin states have different energies in the
magnetic field, transitions from the higher energy state to the lower energy one are
slightly more probable than interactions going the opposite way. After the electrons
have been in a storage ring for a long time, they can achieve a polarization up to 70%.
Measuring the spin of a final state particle can be done by observing the kinematics
(in particular angular correlations) between its decay products. In the case of muons,
however, this is difficult because they are too long-lived to decay inside a typical detector.
The upshot is that we are often interested in unpolarized cross sections, i.e. cross sec-
tions averaged over initial state spins and summed over final state spins.
P We have already
seen how toP evaluate spin sums in section 3.1.3. Using the relations s=1,2 us (p)ūs (p) =
/ + m and s=1,2 v s (p)v̄ s (p) = p
p / − m, we obtain
1 X 1 XXX
|M|2 ≡ |M|2
2 s 2 0 r 0 s r
e4 0 µ ν
0
= / − me )γ (p
tr (p / + me )γ tr (k/ + mµ )γµ (k/ − mµ )γν (5.29)
4q 4
Don’t let yourself get confused by the fact that µ appears both as a Lorentz index here
and as an index distinsguishing the electron mass me from the muon mass mµ .
tr γ µ = 0 . (5.30)
In fact the trace of any product of an odd number of gamma matrices vanishes because,
using (γ 5 )2 = 1 and the cyclic property of the trace, we have
tr γ µ γ ν = tr 2g µν · 14×4 − γ ν γ µ
87
Chapter 5 Quantum Electrodynamics
= 8g µν − tr γ µ γ ν . (5.34)
Therefore,
tr γ µ γ ν = 4g µν . (5.35)
For any even number 2n of gamma matrices, we can use the anticommutator {γ µ , γ ν } =
2g µν to move the leftmost gamma matrix all the way to the right and then use the cyclic
property of the trace to bring it back to the left. This leads to the relation
which can be used to reduce the trace of 2n gamma matrices to a sum of traces of 2n − 2
gamma matrices. (In the last term on the right hand side, the cyclic property of the trace
needs to be exploited to bring it to the same form as the left hand side.) In particular,
for four gamma matrices, this leads to
tr γ µ γ ν γ ρ γ σ = tr g µν γ ρ γ σ − γ ν g µρ γ σ + γ ν γ ρ g µσ
= 4 g µν g ρσ − g µρ g νσ + g µσ g νρ .
(5.37)
Even though these are all the relations we need to evaluate the squared matrix element
for the process e+ e− → µ+ µ− , let us also mention some identities we will encounter later.
This concerns in particular traces involving γ 5 . Since γ 5 = iγ 0 γ 1 γ 2 γ 3 , it counts as an
even number of gamma matrices. This means in particular that any trace involving one or
several γ 5 and an odd number of γ 0 , . . . , γ 3 , vanishes. Also, the trace of γ 5 itself vanishes.
This is immediately obvious from the explicit form γ 5 = diag(−12×2 , 12×2 ) in the chiral
representation, but can also be shown more abstractly using only the representation
invariant properties of the gamma matrices:
tr γ 5 γ µ γ ν γ ρ γ σ = −4iµνρσ . (5.39)
88
5.3 e+ e− → µ+ µ−
From here on, we will set me = 0 to keep our expressions shorter. At a typical high
energy collider, where kinetic energies are GeV, the electron mass 0f 511 keV is indeed
negligible. The squared matrix element is then, after contracting Lorentz indices,
8e4 h 0 0 0 0 2 0
i
|M|2 = (p · k )(p · k) + (p · k)(p · k ) + mµ (p · p) . (5.42)
q4
At this point, it is useful to be a bit more explicit about the kinematics. Let us therefore
consider the collision of an e+ e− pair in the center of mass frame, where the two particles
are moving in opposite directions along the z axis. (This is the conventional coordinate
system chosen for processes at a high energy collider.) This kinematic setup is illustrated
in fig. 5.1. With the definitions given in the figure, we can compute the following dot
products:
q 2 = (p + p0 )2 = 4E 2 , p · p0 = 2E 2 ,
(5.43)
p · k = p0 · k 0 = E 2 − E|k| cos θ , p · k 0 = p0 · k = E 2 + E|k| cos θ .
e4 h 2 2 2 2 2 2
i
|M|2 = E (E − |k| cos θ) + E (E + |k| cos θ) + 2m µ E
2E 4
m2µ m2µ
= e4 1 + 2 + 1 − 2 cos2 θ . (5.44)
E E
89
Chapter 5 Quantum Electrodynamics
We have written the last line such that the high energy limit E mµ , relevant for
instance for the LEP collider, is easy to take. We have eliminated |k|2 in favor of E 2 and
m2µ .
Similar expressions hold of course for the delta functions in the other components of
the 4-momentum vector. Here, the integration interval [−L/2, L/2] is interpreted as the
boundaries of a large box of volume V = L3 , to which the process is confined.
We also have to worry about the normalization of the external states: remember that
one-particle plane wave states are normalized according to hp|qi = (2π)3 2Ep δ (3) (p − q).
To derive a physical observable, we would rather like to have them normalized to one.
This can be achieved by adding a factor [2Ep ]−1 for each particle. For the initial state
particles, we should also divide by V (the height of the δ-peak in a finite volume). For
the final states, we instead integrate over the 3-momentum (the phase space integral).
Finally, in a collider environment we are not colliding individual electron with individ-
ual positrons, but rather beams containing many particles. We should therefore multiply
by the total number of electrons, N− , and positrons, N+ .
In total, we thus have for the transition probability:
d3 k d3 k 0
Z
1 1
P = lim T V N+ N−
T,L→∞ 2Ep V 2Ep0 V (2π)3 2Ek (2π)3 2Ek0
90
5.3 e+ e− → µ+ µ−
where we have arbitrarily called the e− the “target particles” and the e+ the “incident
particles”. The number of target e− in the denominator cancels the factor N− from
eq. (5.47). The e+ flux is given by the e+ number density, N+ /V times the relative
velocity vrel of the e+ and the e− . Note that also the factors T and V cancel, so that
our final expression for the cross section carries no trace of the fact that we temporarily
worked with a finite spacetime volume. In other words, the limit for T, L → ∞ is trivial
to take and we obtain a master formula for 2 → 2 cross sections:
d3 k d3 k 0
Z
+ − + − 1
σ(e e → µ µ ) =
2Ep 2Ep0 vrel (2π)3 2Ek (2π)3 2Ek0
For processes with more than two R particles in the final state, one simply has to add more
final state phase space factors d3 kj /[(2π)3 Ekj ].
For the 2 → 2 case, the 4-dimensional delta function can be used to remove for instance
the integral over d3 k 0 and the integral over d|k|, leaving only the integral over the solid
angle of the first final state particle. If the masses of the two final state particles are m
and m0 , we find
d3 k d3 k 0
Z
(2π)4 δ (4) (p + p0 − k − k 0 )
(2π)3 2Ek (2π)3 2Ek0
|k|2
Z Z p p
= dΩ d|k| δ E + E 0 − |k| 2 + m2 − |k|2 + m02
p p
(2π)2 4Ek Ek0
|k|2 |k| −1
|k|
Z
= dΩ +
(2π)2 4Ek Ek0 Ek Ek0
|k|
Z
= dΩ 2
(5.50)
(2π) 4(Ek + Ek0 )
If we are not interested in the total cross section, but in a differential cross section (for
instance dσ/dΩ), we simply omit the integral over the corresponding kinematic variable
(for instance Ω) in the final state. For instance, for e+ e− → µ+ µ− , we have in the center
of mass frame,
s
dσ |M| 2 m2µ
= 1 − , (5.51)
dΩ 128π 2 E 2 vrel E2
where E is the common energy of the electrons and muons.
91
Chapter 5 Quantum Electrodynamics
Figure 5.2: The total cross section for e+ e− → µ+ µ− as a function of the center of
2 to better expose
mass energy Ecm ≡ 2E. The vertical axis has been multiplied by Ecm
the asymptotic behavior for Ecm → ∞. Figure taken from [1].
Here we have introduced the electromagnetic fine structure constant α ≡ e2 /(4π). The
total cross section is
s
2 m2µ 1 m2µ
πα
σtotal = 1− 2 1+ . (5.53)
3E 2 E 2 E2
5.3.6 e+ e− → µ+ µ− : Summary
Let us briefly summarize how we obtained the cross section for the annihilation of an
electron–positron pair to muons:
1. Draw all relevant Feynman diagrams (here, there was only one)
3. Square it and average over initial state spins/sum over P final state spins. Use
eqs. (3.25) and (3.29) to rewrite expressions of the form s ūs (p)us (p) as traces
over momenta, masses and Dirac matrices.
5. Pick a suitable reference frame and write the 4-momenta in terms of the measurable
kinematic variables in that frame.
92
5.4 More Technology for Evaluating QED Feynman Diagrams
6. Integrate |M|2 over the final state phase space according to the master formula
eq. (5.49) to obtain the cross section.
Here, ↔ means that the left hand side and the right hand side are not identical, but
when computing squared matrix elements they can be used interchangeably. Here is the
93
Chapter 5 Quantum Electrodynamics
√
proof: consider the two physical transverse polarization states ±
µ (p) ≡ (0, 1, ±i, 0)/ 2,
assuming that p is aligned with the z axis. Then
0
X 1
∗µ (p)ν (p) =
1
s=+,−
0
where êµt = (1, 0, 0, 0) and êµz = (0, 0, 0, 1). Next, note that
pµ − (êt · p)êµt
êµz = . (5.57)
êt · p
We now argue that the term proportional to pµ on the right hand side does not contribute
to a Feynman amplitude. The argument is gauge invariance: the matrix element of any
physical process involving an external photon field Aµ (x) is invariant under Aµ (x) →
Aµ (x) + 1e ∂ µ α(x). In momentum space, this gauge transformation rule reads
i
õ (p) → õ (p) − pµ α̃(p) . (5.58)
e
For the external photon, õ (p) ∝ µ (p). Thus, any contribution ∝ pµ to µ (p) must
vanish. Therefore,
X
∗µ (p)ν (p) ↔ −gµν + êtµ êtν − êtµ ẑtν = −gµν . (5.59)
s=+,−
94
Path Integrals
6
Now that we have developed and applied the techniques required to compute physical
observables in quantum field theory, it is useful to pause a moment and think about
the physics hiding behind the algebra. In this chapter, we will therefore discuss an
approach to QFT different from the canonical quantization procedure introduced in
chapters 2 and 3: the path integral formalism. In that formalism, a scattering amplitude
is computed by considering all possible phase space trajectories that particle could take
to go from the initial state to the final state. The amplitudes for all these paths are then
integrated up, true to the superposition principle of quantum mechanics.
p̂2
Ĥ(p, q) = + V (q̂) . (6.1)
2m
Here, V (q) is the potential. We wish to compute the amplitude hq 0 |e−iĤt |qi for the
particle to propagate from q at time 0 to q 0 at time t. We split the time interval [0, t]
into n + 1 small subintervals of length δt ≡ t/(n + 1). Between intervals, we insert a
complete set of coordinate eigenstates:
n
Z Y
0 −iĤt
hq |e |qi = dqj hq 0 |e−iĤ δt |qn ihqn |e−iĤ δt |qn−1 i · · · hq1 |e−iĤ δt |qi . (6.2)
j=1
1
For the moment, we write operators with a hat, e.g. Ĥ and c-numbers without.
95
Chapter 6 Path Integrals
We would like to split each exponential factor into the kinetic part and the potential
part using the Campbell-Baker-Hausdorff formula
1
eT̂ +V̂ = eT̂ eV̂ e− 2 [T̂ ,V̂ ]+··· . (6.3)
In the limit n → ∞, i.e. δt → 0, the commutator term, which is of order (δt)2 and all
higher order terms can be neglected. Let us moreover insert a complete set of momentum
eigenstates. Then, each factor in eq. (6.2) takes the form
p̂2 ED
Z
−iĤδt dpj D E
hqj+1 |e |qj i ' qj+1 exp − i δt pj pj exp − iV (q̂) δt qj
2π 2m
(6.4)
Z
pj2
dpj h i
' exp − i δt − iV (qj ) δt eipj (qj+1 −qj ) (6.5)
2π 2m
p2j
Z
dpj h i
' exp − i δt − iV (qj ) δt + ipj q̇j δt . (6.6)
2π 2m
In the second line, we have used that hq|pi = eipq , and in the last line, we have used that
(q2 − q1 )/δt ' q̇. To write the factors corresponding to first and the last time step in the
form of eq. (6.6), we define q0 ≡ q, qn+1 ≡ q 0 . Equation (6.2) now becomes
Z Y n Y n n 2
0 −iĤt dpk X pk
hq |e |qi = dqj exp − i + V (qk ) − pk q̇k δ t . (6.7)
2π 2m
j=1 k=0 k=0
The physical interpretation of eq. (6.9) is the following: each set {q1 , . . . , qn , p0 , . . . pn }
corresponds to one possible phase space Rtrajectory leading from q to q 0 . The amplitude
for each path is given by the action exp(i dt L). The path integral (integral over Dq Dp)
sums up the amplitudes for all the paths. In that sense, any quantum mechanical tran-
sition process can be viewed as a continuum generalization of the two-slit experiment.
Already at this stage, we can glimpse why the path integral formalism may be useful
in QFT. Imagine we want to compute the expectation value of the particle’s position at
a time t1 , somewhere in the interval [0, t]. It is given by
hq 0 |q̂(t1 )|qi ≡ hq 0 |e−iH(t−t1 ) q̂e−iH(t1 −0) |qi
96
6.2 The Path Integral for a Free Scalar Field
Z Z
= Dq Dp q(t1 ) exp i dt L(q, p) . (6.10)
So far, so good. Now, let’s insert one more q factor. If t1 > t2 , we have
Thus, overall,
(
hq 0 |q̂(t1 )q̂(t2 )|qi
Z Z if t1 > t2
Dq Dp q(t1 )q(t2 ) exp i dt L(q, p) = (6.13)
hq 0 |q̂(t2 )q̂(t1 )|qi if t2 > t1
= hq 0 |T q̂(t1 )q̂(t2 )|qi . (6.14)
In other words, if we had an efficient way of evaluating path integrals, we’d have a new
way of computing time-ordered correlation functions. And we have seen in our discussion
of the LSZ reduction formula (section 4.4) that these are of paramount importance for
computing scattering amplitudes.
We will now introduce the path integral formalism in QFT, we will argue that it leads
to the same Feynman rules as our previous approach involving creation and annihilation
operators (though in an arguably more elegant way), and we will use path integrals to
proof the Feynman rule for the photon propagator.
with the canonical momentum π(x, t) = φ̇(x, t). Finally, instead of discretizing the
trajectory only in time, we discretize all four spacetime directions and integrate over the
field value at each spacetime point. The corresponding integration measure is
Y
Dφ ≡ dφ(x1i , x2j , x3k , tl ) . (6.16)
i,j,k,l
97
Chapter 6 Path Integrals
R
Similarly, also the path integral over the canonical momentum Dπ, is now understood
as
Y dπ(x1i , x2j , x3k , tl )
Dπ ≡ . (6.17)
2π
i,j,k,l
The transition amplitude between an initial field configuration φ(x, 0) and a final field
configuration φ0 (x, t) is then, in complete analogy to eq. (6.2),
Z
0 −iĤt
hφ (x, t)|e |φ(x, 0)i = Dφ hφ0 (x, t)|e−iĤδt |φ(x, tn )i
2πi 1/2
1 4 2
exp i δ x φ̇ (x) . (6.23)
δ4x 2
98
6.2 The Path Integral for a Free Scalar Field
The prefactor can be absorbed into a redefinition of the integration measure Dφ. Then,
the matrix element eq. (6.21) is
Z Z t Z
0 −iĤt 3 1 µ 1 2 2
hφ (x, t)|e |φ(x, 0)i = Dφ exp i dt d x (∂µ φ)(∂ φ) − m φ
0 2 2
Z Z t Z
= Dφ exp i dt d3 x L(φ̇, φ) . (6.24)
0
In QFT, an object of central importance will be the transition amplitude from the vacuum
at t = −∞ to the vacuum at t = +∞. It is called the partition function and is given by
Z Z
Z0 ≡ h0(t = +∞)|e−iĤt |0(t = −∞)i = Dφ exp i d4 x L(φ̇, φ) . (6.25)
Note the allusion to statistical mechanics implied by the term “partition function”. In
statistical mechanics, the partition function Z ≡ n hn|e−Ĥ/T |ni sums up the Boltzmann
P
factors for all microscopic states |ni of a system. In the path integral formalism of QFT,
the states become trajectories in configuration space, the temperature is set to 1, and
there is an extra factor of i in the exponent. Since a non-interacting system starting out
in its ground state will remain there forever, Z0 = 1 in the free theory. Defining the
partition function still makes sense because it will differ from unity as soon as we add
an interaction.
Note that, as in QM, the path integral formalism provides a way of writing time-
ordered correlation functions:
Z Z
4
h0|T φ(x1 ) · · · φ(xn )|0i = Dφ φ(x1 ) · · · φ(xn ) exp i d x L(φ̇, φ) . (6.26)
Before concluding this section, we will add a very particular interaction to the La-
grangian: a source term
Lsource ≡ φ · J . (6.27)
This term can be interpreted as a vertex that generates a particle out of nowhere, or
annihilates a particle into nothingness, without requiring the presence of another parti-
cle. The real coefficient function J(x) gives the strength of such processes and can be
interpreted as a classical (non-quantized) field. The reason for adding the source term
will become clear shortly. But first, let us define
Z Z
4
Z0 [J] ≡ Dφ exp i d x L(φ̇, φ) + φJ . (6.28)
99
Chapter 6 Path Integrals
and so on. Because of these relations, Z[J] is also called the generating functional of the
theory.
Note that we have split the Fourier transform of the source term into two (identical) pieces
and added a factor 1/2. Since the Fourier transform is a linear unitary transformation,
we can replace the path integral over Dφ in the definition of Z0 [J] by a path integral
over Dφ̃. The Jacobian corresponding to this substitution is 1.
Our goal is to simplify eq. (6.34). To this end, let us complete the square (in φ̃) in the
exponent by defining
˜
J(k)
χ̃(k) ≡ φ̃(k) + . (6.35)
k 2 − m2 + i
This substitution, being a constant shift, does not change the integration measure either,
i.e. we can simply replace Dφ̃ by Dχ̃. The action now becomes
1
Z 4
d k ˜ J(−k)
J(k) ˜
2 2
S0 = χ̃(k) k − m χ̃(−k) − 2 (6.36)
2 (2π)4 k − m2 + i
100
6.4 Wick’s Theorem from the Path Integral
i
Z ˜ J(−k)
d4 k J(k) ˜
· exp − . (6.37)
2 (2π)4 k 2 − m2 + i
The first line is just Z0 [J = 0] = 1. The second line does not depend on χ̃ any more—we
have gotten rid of the path integral. We transform the second line back to coordinate
space using the obvious definition
Z
˜
J(k) ≡ d4 x eikx J(x) . (6.38)
Note that the k-dependent term obtained that way is just the Feynman propagator
DF (x − y). We thus obtain
Z
1 4 4 0 0 0
Z0 [J] = exp − d x d x J(x )DF (x − x )J(x) . (6.39)
2
We knew from section 2.4.3 that this had to be the outcome, but it is reassuring to see
that it can be obtained in the path integral formalism as well.
Consider next the 4-point function. By applying the functional derivative four times,
we obtain
1 2 1 2 1 2
= + + . (6.41)
3 4 3 4 3 4
101
Chapter 6 Path Integrals
By expanding the first exponential (but not the second), we recover a series of time-
ordered correlation function that coincides precisely with the perturbation series in the
numerator of our master formula, eq. (4.40). Wick’s theorem then tells us that, to
evaluate the correlation function, we have to consider all possible ways of connecting
the φ factors by propagators. We thus immediately recover the Feynman rules from
section 4.3.
We have thus rederived the formalism for evaluating correlation functions in terms of
Feynman diagrams without ever having to invoke creation and annihilation operators.
From here on, the remaining step towards the computation of physical observables are
the same as before: the LSZ formula relates correlation functions to matrix elements,
which are then integrated over phase space.
Note that all correlation functions we encountered in the path integral formalism so
far were correlation functions of the free theory, as they must be for our master formula
to be applicable. As we have seen in eq. (4.40), going from the free theory to the full
theory leads to an additional normalization factor, and this factor is the reason we have
written “∝” instead of “=” in eq. (6.44).
102
6.6 Quantization of the Photon Field
d4 p −ig µν e−ip(x−y)
Z
µν
∆ (x − y) ≡ . (6.45)
(2π)4 p2 + i
In the second step, we have integrated by parts. As we did for the scalar field in eq. (6.34),
we now transform the path integral to Fourier space:
Z Z 4
i d k
Z0 [J] = Dà exp − õ (k)(k 2 g µν − k µ k ν )Ãν (−k)
2 (2π)4
˜µ ˜µ
− J (k)õ (−k) − J (−k)õ (k) . (6.48)
We have again split up the source term into two identical pieces and included a factor
1/2 to compensate for the implied double counting.
As in section 6.3, we would like to complete the square in the exponent to rewrite
Z0 [J] as Z0 [0] = 1, multiplied by a term not involving a path integral any more, but
containing J and the photon propagator. Doing so would require us to invert the matrix
k 2 g µν − k µ k ν ≡ k 2 P µν , so that we could define B̃ µ (k) ≡ õ (k) − [(k 2 P )−1 ]µν J˜µν (k).
Unfortunately, the matrix P is singular because P µν kν = 0.
The reason for the singular nature of P is gauge invariance. L is invariant under
A (x) → Aµ (x)+(1/e)∂ µ α(x), which in momentum space translates to õ (k) → õ (k)−
µ
(i/e)k µ α̃(k). This implies that the component of Aµ (k) proportional to k µ is unphysical,
but it also suggests a solution to our difficulties defining the photon propagator: we
simply drop the unphysical component of the photon field and redefine DÃ to mean
103
Chapter 6 Path Integrals
path integration only over components of à orthogonal to k, i.e. k µ õ (k) = 0. In other
words, we fix the gauge to the Lorenz gauge. Within this subspace, P µν is just the
identity matrix since
kµ kν kν k λ
µν λ µν λ
P (k)Pν (k) = g − 2 gν − 2
k k
µ
k k λ
= g µλ − 2
k
= P µλ (k) (6.49)
and
The first relation implies that P µν (k) is a projection operator, i.e. its eigenvalues can
only be zero or one. The second relation gives just the trace of P µν and implies that three
of the eigenvalues are 1 and the fourth one is zero. We have already argued that the zero
eigenvalue corrresponds to the direction of k µ , which we have removed from the path
integral. Therefore, in the remaining subspace, all three eigenvalues are 1 and P µν (k)
must be the identity matrix. Therefore, in this subspace, (k 2 P µν )−1 = (1/k 2 )P µν . We
define
In the second step, we have used that Z0 [0] = 1 to eliminate the path integral over DB̃.
We have also exploited the fact that k µ Jµ (x) = 0, so that replacing P µν = g µν − k µ k ν /k 2
by g µν leaves Z0 [J] invariant. Finally, we emphasize that the integration measure DB̃
implies integration only over the components of B̃ that are orthogonal to k µ .
104
6.7 Path Integrals for Fermions
θη = −ηθ . (6.53)
For finite sets of Grassmann numbers, one can construct matrix representations of G.
The above definition is, however, more general. One immediate consequence of eq. (6.53)
is that
θ2 = 0 . (6.54)
f (θ) = A + Bθ . (6.55)
In other words, the Taylor expansion of f (θ) terminates after the linear term. Note
that A is an ordinary number and B is a Grassmann number, so that Bθ is an ordinary
number.
R We will also want to integrate over Grassmann numbers, and we therefore define
dθ f (θ) by requiring the properties
Z Z
dθ c · f (θ) = c · dθ f (θ) , (6.56)
Z Z Z
dθ f (θ) + g(θ) = dθ f (θ) + dθ g(θ) , (6.57)
Z Z
dθ f (θ + η) = dθ f (θ) , (6.58)
Z
dθ 1 = 0 . (6.59)
for θ, η ∈ G and c ∈ C. These properties, inRparticular the invariance under shifts of the
integration variable, can be satisfied only if dθ f (θ) is a constant times B. We therefore
define
Z Z
dθ f (θ) = dθ (A + Bθ) ≡ B . (6.60)
For multidimensional Grassmann integrals, the ordering of the integrals and the Grass-
mann numbers in the integrand is important. We choose the convention
Z
dθ dη η θ = +1 . (6.61)
105
Chapter 6 Path Integrals
∗
Compare this to the corresponding integral with only real numbers, dx dx∗ e−x bx =
R
2π/b.
We first note that, in the Taylor expansion of the exponential, Q the only term that
contributed is the one in which all θi , θi∗ appear exactly once: i θi∗ θi . The strategy
is to transform to an eigenbasis of B by applying a unitary transformation θi → θi0 =
106
6.7 Path Integrals for Fermions
In the last step we have used the unitarity of U . We can now write
Z P 0∗ 0
I = dθ1∗ dθ1 · · · dθn∗ dθn e− k θk bk θk (6.75)
Z Y
= (−1)n dθ1∗ dθ1 · · · dθn∗ dθn bk θk∗ θk (6.76)
k
Y
= bk (6.77)
k
= det B . (6.78)
This should again be compared to the corresponding integral over ordinary num-
∗
bers, dx∗1 dx1 · · · dx∗n dxn e−xi Bij xj = (2π)n / det B.
R
Remember that for the free scalar field, we were able to rewrite the partition function in
a form without the path integral as
Z
1 4 4 0 0 0
Z0 [J] = exp − d x d x J(x )DF (x − x )J(x) , (6.81)
2
107
Chapter 6 Path Integrals
(see eq. (6.39)). To do the sameR for Z0 [η, η̄], we follow completely analogous arguments.
We transform the action S0 = d4 x L0 into Fourier space, using
Z 4
d k −ikx
ψ(x) ≡ e ψ̃(k) , (6.82)
(2π)4
Z 4
d k −ikx
η(x) ≡ e η̃(k) . (6.83)
(2π)4
This yields
d4 k h ˜
Z i
S0 = ψ̄ (k)(k/ − m) ψ̃(k) + ˜(k)ψ̃(k) + ψ̄˜(k)η̃(k) .
η̄ (6.84)
(2π)4
η̃(k)
χ̃(k) ≡ ψ̃(k) + . (6.85)
k/ − m + i
This leads to
Z 4 h
d k 1 i
S0 = ˜(k)(k/ − m)χ̃(−k) − η̄˜(k)
χ̄ η̃(k) (6.86)
(2π)4 k/ − m + i
0
d4 k eik(x −x)
Z Z
= d4 x χ̄(x)(i∂/ − m)χ(x) − d4 x d4 x0 η̄(x) η(x0 ) . (6.87)
(2π)4 k/ − m + i
from which it follows that
h Z i
Z0 [η, η̄] ≡ Z0 [0, 0] exp − d4 x d4 x0 η̄(x)SF (x − x0 )η(x0 ) . (6.88)
Note that Z0 [0, 0] = 1 since a field that is initially in the vauum state will remain there
forever in the absence of external sources.
We can again obtain time-ordered correlation functions by taking functional derivatives
of Z0 [η, η 0 ]. For instance
1 δ 1 δ
h0|T ψ(x1 )ψ̄(x2 )|0i = − Z0 [η, η̄]η=η̄=0 . (6.90)
i δ η̄(x1 ) i δη(x2 )
Note the extra minus sign in the functional derivative with respect to η(x2 ). Its origin
is the fact that, in the last term in the Lagrangian eq. (6.79), η appears to the right
of ψ̄, so that the derivative operator δ/δη(x2 ) needs to be anticommuted past ψ̄ before
evaluating it.
108
6.8 The Quantum Equations of Motion: Schwinger–Dyson Equations
In quantum mechanics,
R we are used to deriving equations of motion by demanding that
the action S = dt L is invariant under infinitesimal variations of the system’s phase
space trajectory. Let us consider now infinitesimal variations of the field φ(x):
Since the path integral means that the field value φ(x) is integrated over at each spacetime
point, this shift is nothing but a simple substitution of variables in the path integral.
At each spacetime point, eq. (6.92) is just a constant shift, which leaves the integration
measure Dφ invariant. Therefore,
Z Z
i d4 x L[φ] 0
R R 4
Dφ φ(x1 ) · · · φ(xn ) e = Dφ φ0 (x1 ) · · · φ0 (xn ) ei d x L[φ ] . (6.93)
This is just the expression that vanishes in the classical theory according to the Euler–
Lagrange equations of motion. In the terms in the second line of eq. (6.94), we can
write
Z
(xj ) = d4 x (x) δ (4) (x − xj ) . (6.96)
Noting now that eq. (6.94) has to hold for any (x), and writing the path integral in in
terms of correlation functions, we find
Z E
D δ 4 0 0
ΩT φ(x1 ) · · · φ(xn ) d x L[φ(x )] Ω
δφ(x)
109
Chapter 6 Path Integrals
n
X
ΩT φ(x1 ) · · · φ(xj−1 ) iδ (4) (x − xj ) φ(xj+1 ) · · · φ(xn )Ω .
= (6.97)
j=1
These equations are called the Schwinger–Dyson equations. The terms involving δ-
functions on the right hand side are called contact terms because they are non-zero
only when x equals one of the xj . The Schwinger–Dyson equations can be interpreted as
the equations of motion of the quantum theory: they tell us that the classical equations of
motion (which would dictate that eq. (6.95) vanishes) still hold inside a correlation func-
tion, provided the coordinate x does not equal any of the xj appearing in that correlation
function.
with α(x) infinitesimal. This is just an infinitesimal gauge transformation on ψ(x). Note,
however, that we do not apply the corresponding gauge transformation to Aµ (x), i.e. the
Lagrangian is not invariant under eq. (6.98). Instead, it transforms as
LQED [ψ, ψ̄, A] → LQED [ψ 0 , ψ̄ 0 , A] ≡ LQED [ψ, ψ̄, A] − e [∂µ α(x)] ψ̄γ µ ψ . (6.99)
The crucial point is that the Lagrangian changes only by a term containing a deriva-
tive of α(x). The reason is that LQED is invariant under global (i.e. x-independent)
transformations ψ(x) → (1 + ieα)ψ(x).
Consider now the transformation of the path integral
Z R 4
Dψ̄ Dψ DA ei d x LQED [ψ,ψ̄,A] ψ(x1 )ψ̄(x2 )
Z
0 0
R 4
= Dψ̄ Dψ DA ei d x LQED [ψ ,ψ̄ ,A] ψ 0 (x1 )ψ̄ 0 (x2 ) . (6.100)
The identity follows from the fact that Dψ and Dψ̄ are invariant under eq. (6.98). Ex-
panding the right hand side of eq. (6.100) in α(x), and introducing the notation
we obtain
Z Z
i d4 x LQED [ψ,ψ̄,A]
R
0= Dψ̄ Dψ DA e d4 x [∂µ α(x)] j µ (x) ψ(x1 )ψ̄(x2 )
−i
+ ieα(x1 )ψ(x1 ) ψ̄(x2 ) + ψ(x1 ) − ieα(x2 )ψ̄(x2 )
110
6.9 The Ward–Takahashi Identity
Z Z
i d4 x LQED [ψ,ψ̄,A]
R
= Dψ̄ Dψ DA e d x iα(x) [∂µ j µ (x)] ψ(x1 )ψ̄(x2 )
4
+ ieδ (4) (x − x1 )α(x1 )ψ(x1 )ψ̄(x2 ) − ieδ (4) (x − x2 )ψ(x1 )α(x2 )ψ̄(x2 ) .
(6.102)
In the last step, we have integrated by parts in the first term and pulled the last two
terms under the integral by introducing appropriate δ functions. SinceR eq. (6.102) has
to hold for any infinitesimal α(x), it must hold also when the integral d4 x is omitted.
Written in terms of correlation functions, this implies
i∂µ h0|T j µ (x)ψ(x1 )ψ̄(x2 )|0i = − ieδ (4) (x − x1 )h0|T ψ(x1 )ψ̄(x2 )|0i
+ ieδ (4) (x − x2 )h0|T ψ(x1 )ψ̄(x2 )|0i . (6.103)
This identity, which is called the Ward–Takahashi identity, can be expressed diagram-
matically as
x1 x1 x1
i∂µ · x
(4) (4)
= δ (x − x1 )
ie −δ (x − x2 ) .
x2 x2 x2 (6.104)
Note that the left hand side is just the full QED vertex (including all quantum correc-
tions), but with the photon field Aµ (x) replaced by a derivative ∂µ . The contact terms
on the right hand side involve, besides the δ-functions, the full propagator.
Let us elaborate more on the connection between the left hand side of the Ward–
Takahashi identity and the correlation function h0|T Aµ (x)ψ(x1 )ψ̄(x2 )|0i (the full QED
vertex), and also on its connection to a corresponding scattering amplitude hf |ii obtained
by applying the LSZ reduction formula to it. We use the equations of motion for the
photon field. Working in Lorenz gauge ∂µ Aµ = 0, they read
We know from the previous section that these equations of motion still hold inside a
correlation function, up to contact terms. Therefore, eq. (6.103) is equivalent to
The double derivative ∂ 2 on the left hand side appears also in the LSZ reduction formula.
We now put in all the remaining pieces of that formula except the polarization vector
µ (pk). (Since Aµ (x) is already contracted with ∂µ , we would not have anything to
111
Chapter 6 Path Integrals
contract it with.) On the left-hand side of the LSZ formula, we write the scattering
amplitude as
←−
× h0|T ∂µ Aµ (x)ψ(x1 )ψ̄(x2 )|0i (−i ∂/ x2 − m)u(p) e−ipx2 .
(6.108)
Z Z Z
d4 x d4 x1 d4 x2 e−ikx eiqx1 ū(q)(i∂/x1 − m)
= −i
←−
× (contact terms) (−i ∂/ x2 − m)u(p) e−ipx2 .
(6.109)
Let us first consider eq. (6.108). Integrating by parts over x1 , it turns into
Z
−ikµ Mµ (p, k, q) = lim (−i) d4 x1 eiqx1 ūs (q)(/q − m) . . . h0|T ψ(x1 ) . . . |0i
q 2 →m2
We have written out only the pieces corresponding to ψ(x1 ) here. The dots indicate the
other pieces. The limit is introduced to avoid the otherwise infinite expression 1/(q 2 −
m2 ), which would appear in the correlation function according to the Feynman rules. In
the
R 4 second line of eq. (6.110), we have introduced the Fourier-transformed field ψ̃(q) =
iqx
d x e ψ(x). Since the scattering matrix element hf |ii must be finite (up to the overall
energy-momentum conserving δ function), the correlation function h0|T ψ̃(q) . . . |0i must
include a pole proportional to 1/(/q − m) to cancel the factor /q − m from the LSZ formula,
and similar poles also for all other external particles. For the specific case of the full
QED vertex, the requisite pole structure is
1 1 1
. (6.111)
/ − m k2
/q − m p
When we discussed the Feynman rules for correlation functions, we saw that these pole
corresponded to the propagators attached to the external vertices. It is the coefficient of
all these poles that gives hf |ii. Terms that are lacking one or several of the poles do not
contribute since they go to zero in the limit q 2 → m2 , p2 → m2 , k 2 → 0.
Consider now the contact terms (right hand side of eq. (6.106) and last expression in
eq. (6.109)). The δ-function in a contact term eliminates one of the coordinate integrals
and simplifies the momentum dependence. For instance, consider the term containing
the delta function δ (4) (x − x1 ). After eliminating the x or x1 integral using this delta
function, the resulting term depends only on q−k, but not on the orthogonal combination
112
6.9 The Ward–Takahashi Identity
q + k. Therefore, it cannot have separate 1/k 2 and 1/(/q − m) poles. It thus does not
have the structure of eq. (6.111) and therefore does not contribute to hf |ii.
These arguments can be generalized to arbitrary correlation functions and also to
theories other than QED. Therefore, we conclude that contact terms never contribute to
scattering matrix elements.
Coming back to the electron vertex function, we can now conclude from eqs. (6.108)
and (6.109) that
k µ Mµ = 0 . (6.112)
In other words, when we take the matrix element describing the interaction of a photon
with two fermions and replace the photon polarization vector µ (k) by k µ , the result is
zero.
Of course, we could have carried out a fully analogous derivation also for QED diagrams
with more external vertices. The result would be the same: any QED scattering matrix
element µ (k)Mµ satisfies kµ Mµ = 0.
Remember that our starting point in this section was the invariance of LQED under
global gauge transformations (the statement that under a local gauge transformation
LQED changes only by a term containing a derivative of α(x)). The Ward–Takahashi
identity and the relation kµ Mµ = 0 are direct consequences of this. Another way of
phrasing eq. (6.112) is to say that the amplitude for the production of a (hypothetical)
longitudinal photon vanishes. It is often useful when computing complicated Feynman
diagrams to check the calculation by verifying that eq. (6.112) is satisfied.
113
Chapter 6 Path Integrals
114
Weyl and Majorana Fermions
7
In section 3.4.1, we have already argued that the left-chiral and right-chiral components
of a 4-component spinor belong to different representations of the Lorentz group. (We
have argued that the transformation matrices are block diagonal, thus not mixing the
upper and lower components of a 4-component spinor.) In the Lagrangians we have
encountered so far, the only term that mixes the left-handed and right-handed pieces is
the mass term. For instance, consider the QED Lagrangian. Using
1 − γ5 1 + γ5
ψ(x) = ψ(x) + ψ(x)
2 2
= PL ψ(x) + PR ψ(x)
≡ ψL (x) + ψR (x) , (7.1)
it can be written as
1
LQED = ψ̄(i∂/ − m)ψ − eψ̄γ µ ψAµ − Fµν F µν
4
/ L − eψ̄L γ µ ψL Aµ + ψ̄R (i∂)ψ
= ψ̄L (i∂)ψ / R − eψ̄R γ µ ψR Aµ
1
− mψ̄L ψR + h.c. − Fµν F µν . (7.2)
4
We see that, indeed, apart from the mass term, this looks like the Lagrangian of two
independent fields, each coupled to the photon. Group theory allows one to classify
all possible representations of the Lorentz group SO+ (1, 3) and give them labels.1 By
convention, the representation corresponding to ψL is called ( 12 , 0), an the representation
1
When we write “Lorentz group” in the following, we consider for simplicity only the proper or-
thochronous Lorentz group, not the more general Lorentz group O(1, 3), which includes also time
reversal and parity transformations.
115
Chapter 7 Weyl and Majorana Fermions
where S(Λ) is the 4 × 4 block-diagonal matrix given in eq. (3.74) (for infinitesimal tran-
formations). We write
s(Λ)
S(Λ) = (7.4)
(s̄T )−1 (Λ)
(The transposition and the inverse in the second row are conventions that will turn out
to be useful later in this chapter. For now, this is just how we name the lower 2 × 2 block
of S(Λ).) The upper 2 × 2 block, which is relevant for the transformation of left-handed
spinors, has the form
i
s(Λ)αβ = δα β − (σµν )αβ ω µν , (7.5)
4
where in the 2-component world, σ µν is understood to mean
i µ ν β
(σ µν )αβ = σ σ̄ − σ ν σ̄ µ α . (7.6)
2
116
7.1 Spinor Indices
As we hinted at above, the position of the spinor indices used here is not arbitrary. The
convention is that an upper index is always contracted with a lower index and vice-versa.
(The normalization is a mere convention.) The salient feature of αβ is its invariance
under Lorentz transformations. To see this, note first that det s(Λ) = 1, as befits a
symmetry transformation of a quantum field (after all the “O” in SO+ (1, 3) stands for
“orthogonal”). According to the definition of the determinant, this means in component
notation
This proves the invariance of αβ under Lorentz transformations, making it the equivalent
of gµν for spinor indices. We define αβ (with upper indices) as the inverse of αβ (with
lower indices). By defining
we see that
ψ α ≡ αβ ψβ . (7.13)
117
Chapter 7 Weyl and Majorana Fermions
This can be modified further by using the following relation, which follows from eq. (7.10):
η
θβ s(Λ)β δ αδ ακ = θβ s−1 (Λ)α ηβ ακ
θ
⇔ θβ s(Λ)β κ = s−1 (Λ)α ακ
α
⇔ αβ s(Λ)β γ = s−1 (Λ)β βγ . (7.15)
However, the antisymmetry of αβ introduces complications in the form of extra minus
signs sometimes:
The last equality follows from the anticommutation property of spinor fields. In the
following we will adopt the following convention for contractions of two spinors written
without spinor indices:
ψχ ≡ ψ α χα . (7.20)
ψχ = χψ . (7.21)
118
7.1 Spinor Indices
where
α̇ i
s̄−1 (Λ)β̇ = δβ̇ α̇ − (σ̄µν )β̇ α̇ ω µν , (7.23)
4
with
i µ ν α̇
(σ̄ µν )β̇ α̇ = σ̄ σ − σ̄ ν σ µ β̇ . (7.24)
2
Also in the (0, 21 ) representation, we can define an invariant tensor α̇β̇ and its inverse α̇β̇
by exactly the same arguments as for the ( 12 , 0) representation. The relations defining
these tensors are identical to eq. (7.11). Lowering and raising dotted indices works
in exactly the same way as for undotted indices, i.e. there are relations identical to
eqs. (7.12), (7.13) and (7.17) to (7.19), but with all undotted indices replaced by dotted
ones. The conventions for contractions without explicitly written indices is, however,
different: for right-handed spinors, we define
Remembering that the γ matrices are composed of Pauli matrices, we can conclude
immediately that, in two-component notation
119
Chapter 7 Weyl and Majorana Fermions
Since the first index on σ ν transforms with s(Λ), it must be an undotted index. Since the
second one transforms with s̄(Λ), it must be a dotted index. Putting in spinor indices,
eq. (7.27) can be rewritten as
Here, the index structure of s(Λ) can be read off from eq. (7.5). That of s̄T (Λ) must be
the same as that of (s̄T )−1 (Λ) from eqs. (7.22) and (7.23).
In the same way, we can also show that σ̄ µα̇α is invariant under Lorentz transforma-
tions. Alternatively, we can also show this by using that
Here, according to our conventions χ and ψ are left-handed and thus have undotted
indices, while ψ † ≡ ψ̄ is right handed and has a dotted index.
Since the spinor indices in eq. (7.30) are all contracted, the vector current transforms
like a Lorentz vector:
Let us compute the hermitian conjugate of this vector current. This is something we
need to do for instance when computing th squared matrix element |M|2 for a scattering
process. We find
= χ†β̇ σ̄ µβ̇α ψα
= χ† σ̄ µ ψ . (7.32)
The third equality follows from the hermiticity of σ̄ µ . Remember that, for 4-component
spinors we had [ψ̄γ µ χ]† = χ̄γ µ ψ. We have now derived the 2-component version of this
identity.
Another useful relation is
ψ † σ̄ µ χ = ψα̇† σ̄ µα̇β χβ
120
7.2 The QED Lagrangian in 2-Component Notation
where in the third step we have used eq. (7.29) as well as the anticommutation property
of fermion fields.
In the last line, we have used eq. (7.33) followed by an integration by parts on the kinetic
term of χ. We have also applied eq. (7.33) to the gauge interaction term of χ, where it
introduces a minus sign compared to the corresponding term for ψ. This makes sense:
remember that χ̄ is the right-handed component of Ψ, i.e. it has the same charge as ψ.
Consequently, χ must have opposite charge. As expected, we see that, apart from the
mass terms, ψ and χ are decoupled.
Note that eq. (7.35) is invariant under the transformation
121
Chapter 7 Weyl and Majorana Fermions
C ≡ −iγ 2 γ 0 . (7.38)
σ2
1
C = −i (7.39)
σ̄ 2 1
−iσ 2
= (7.40)
−iσ̄ 2
0 −1
1 0
= , (7.41)
0 1
−1 0
and therefore
One conclusion we can draw from this is that charge conjugation transforms a left-handed
4-spinor (ψα , 0)T into a right-handed 4-spinor (0, ψ †α̇ ).
Note that the matrix C has the following useful properties:
C T = C † = C −1 = −C and C∗ = C . (7.46)
This can be used for instance to show that applying charge conjugation twice is the
identity:
and
Ψ̄Θc = Ψ† γ 0 Cγ 0 Θ∗
122
7.3 Majorana Fermions
= −Θ† (γ 0 )T C T (γ 0 )T Ψ∗
= Θ̄Ψc . (7.49)
In the second equality, we have used the fact that all spinor indices are contracted, i.e.
the expression is just a number and we can without loss of generality replace it with its
transpose. We gave to introduce a minus sign, however, since the fermion fields Θ and Ψ
change places. (Write out the indices to see this!) In analogy to eq. (7.49), we also have
Ψc Θ = χc Ψ . (7.50)
Plugging this into eq. (7.35) and using eqs. (7.21) and (7.32) leads to
1 1
LQED = − Fµν F µν + iξ¯1 σ̄ µ ∂µ ξ1 + iξ¯2 σ̄ µ ∂µ ξ2 − mξ1 ξ1 + mξ2 ξ2 + h.c.
4 2
− eξ¯1 σ̄ µ ξ2 − eξ¯2 σ̄ µ ξ1 . (7.53)
The kinetic and mass terms now describe two completely independent 2-component
spinor fields. The gauge interaction term is instead off-diagonal.
For uncharged particles (i.e. particles without electromagnetic gauge interaction terms,
for instance neutrinos), the decoupling of ξ1 and ξ2 is complete. Removing one of them
from the theory should therefore still lead to a valid model, with a Lagrangian of the
form
¯ µ ∂µ ξ − 1 mξξ − 1 mξ † ξ † .
LMajorana = iξσ̄ (7.54)
2 2
A field of this type is called a Majorana fermion. It is a massive fermion with only
two degrees of freedom. This is possible because the right-handed state of a Majorana
fermion is identical to the antiparticle of the left-handed state. To see this, define
ξ
Ψ(x) ≡ †αα̇ , (7.55)
ξ
which implies that Ψc = Ψ. We can also rewrite eq. (7.54) in 4-component notation:
1 1
LMajorana = Ψ̄γ µ i∂µ Ψ − mΨ̄Ψ . (7.56)
2 2
123
Chapter 7 Weyl and Majorana Fermions
W Z
ν ν
where W and Z are the gauge bosons mediating weak interactions, and `j are the charged
lepton fields. But let’s come back to the mass term in eq. (7.57). One may ask the
question whether the “charge” of the neutrinos under the weak interaction does not
forbid this mass term, just like the electric charge forbids Majorana mass terms for the
other SM fermions. The truth is, eq. (7.57) is forbidden in the SM. In fact, all fermion
mass terms (Dirac and Majorana) are forbidden. Then, why are fermions obviously
massive? The solution is the Higgs mechanism, which we will discuss in detail in the
second part of this course, the essentials of which we can however already understand
now. The SM symmetries allow for the following Yukawa couplings:
L ⊃ y ij Li H̃N j + · · · + h.c. , (7.58)
124
7.4 Application: Majorana Neutrinos and the Seesaw Mechanism
where Li = (νLi , `iL )T consists of the left-handed neutrino and charged lepton fields of
the i-th generating, arranged here in a 2-vector, and H̃ is the Higgs field, which is also
a 2-vector. (The reason for arranging these particles in 2-vectors is that they form a
2-dimensional representation of SU (2), one of the symmetry groups of the SM.) The
field N in eq. (7.58) describes the right-handed neutrinos, but as before, we write a right
handed field as the hermitian conjugate of a left-handed one, i.e. N are the left-handed
antiparticles of the RH neutrinos, and N̄ are the RH neutrinos themselves. Note that
in the original form of the SM, right-handed neutrinos are not included. Therefore,
the original SM predicted neutrinos to be exactly massless. With the discovery of neu-
trino oscillations, which prove that neutrinos have nonzero masses, the model had to be
amended. The dots in eq. (7.58) represent similar couplings for the charged leptons and
the quarks. The crucial point is now that hΩ|H̃|Ωi = (v, 0)T 6= 0, i.e. the Higgs field has
a non-vanishing vacuum expectation value. In other words, even when no Higgs particles
are present in a given process, the field is non-zero. Expanding H̃ around this vacuum
expectation value turns the Yukawa couplings in eq. (7.58) into
Here, it is implied that νL = (νLe , νLµ , νLτ )T contains the three left-handed SM neutrino
flavors, and N = (N 1 · · · N n ) contains the singlet neutrinos. mD is thus a 3 × n matrix
and M is an n × n matrix.
We see that the neutrino mass matrix is off-diagonal, i.e. there is mixing between the
νL and N fields. For instance, an electron neutrino does not have a definite mass, but
is a superposition of several mass eigenstates. Similarly, a neutrino of definite mass is a
mixture of different flavor eigenstates. To transform the mass matrix in eq. (7.61) to the
mass basis, we diagonalize it by applying a unitary transformation to the neutrino state.
125
Chapter 7 Weyl and Majorana Fermions
To obtain the form of this transformation, let us consider for simplicity a simplified model
with only one νL flavor and one N flavor. The desired transformation can be written as
νmL cos θ sin θ νL
= . (7.62)
Nm − sin θ cos θ N
and requiring that the off-diagonal elements vanish, we find for the mixing angle
2mD
tan 2θ = − . (7.64)
M
The eigenvalues of the mass matrix are
1
q
m1,2 = M ∓ M 2 + 4m2D ,
(7.65)
2
and the corresponding mass terms read
1 1
L2-flavor ⊃ − m1 νmL νmL − m2 Nm Nm + h.c. . (7.66)
2 2
There are theoretical reasons to believe that M mD . Namely, while mD is determined
by the vacuum expectation value (vev) of the Higgs field v ∼ 174 GeV according to
eq. (7.59), the scale of M is completely free. (Note that y ij cannot be much larger than
O(1) in a consistent QFT, and values much smaller than O(1) are considered unnatural
by many theorists.) If M is determined by some underlying, more complete, theory, the
typical mass scales of that theory should be v, otherwise we would have discovered
it already. If M mD , we see from eq. (7.65) that one neutrino mass eigenstate is of
order M , while the other is
m2D
m1 ' . (7.67)
M
This provides a natural explanation of the smallness of neutrino masses. For instance, for
mD ∼ 100 GeV, M ∼ 1013 GeV, we obtain neutrino masses of order 0.1 eV, consistent
with experimental constraints.
The above mechanism is called the type-I seesaw mechanism because the fact that
the N have a are very large mass term reduces the effective mass of the light neutrinos,
which are mostly νL , with a small admixture of N .
Lifting the restrictions of the 2-flavor toy model and treating mD and M as matrices
again, the seesaw formua eq. (7.67) generalizes to
126
7.4 Application: Majorana Neutrinos and the Seesaw Mechanism
Note that the matrices mD and M can be off-diagonal. mD can in fact be an arbitrary
complex matrix, while M has to be complex symmetric. This follows from the form of
the N mass term, M ij N i N j + h.c. (see eq. (7.60)), which shows that any antisymmetric
contribution would cancel and thus be unphysical. The off-diagonal nature of the mass
matrices implies that also the SM neutrinos can mix among each other. This means, for
instance, that the electron neutrino does not have a definite mass, but is a mixture of
the three light mass eigenstate (plus a tiny, usually negligible admixture of order mD /M
of the heavy mass eigenstates). Conversely, the light neutrino mass eigenenstates are
mixtures of νe , νµ and ντ (and a tiny admixture of the heavy flavor eigenstates). This
mixing is at the core of the so-called neutrino oscillations.
where Q = mH − mHe is the decay energy and mj are the neutrino mass eigenstates.
The spectrum of electron energies from β decay is given by
dNe (Ee ) p q
∝ F (Z, Ee ) Ee2 − m2e Ee (Q − Ee ) (Q − Ee )2 − m2ν̄e θ(Q − Ee − mν̄e ) .
dEe
(7.71)
Here, the Fermi function F (Z, Ee ) accounts for the interaction of the produced electron
with the Coulomb field of the nucleus. By θ(x) we denote the Heaviside function.
To make the dependence on the neutrino mass more visible, one defines the Kurie
functions
s
dNe (Ee )/dEe
K(Ee ) ≡ . (7.72)
F (Z, Ee )pe Ee
For mν̄e = 0, it should be a straight line ∝ Q − Ee , while for mν̄e 6= 0, it has a cutoff,
see fig. 7.1.
127
Chapter 7 Weyl and Majorana Fermions
�ν = �
�ν = �� ��
������ = ��� ��� �� ��
It is important that ν̄e does not have a definite mass, i.e. when we write mν̄e above, this
is actually ill-defined. We should therefore regard the process (7.69) as the combination
of the processes
3
H → 3 He + e− + ν̄1 , (7.73)
3 3 −
H → He + e + ν̄2 , (7.74)
3 3 −
H → He + e + ν̄3 .. (7.75)
Since we can only detect the electron, we do not know on an event-by-event basis which
of these processes has occured, hence we always have to take all three of them into
account. Thus, in particular, the maximum energy of the electron Eemax is given by
the maximum of the kinematic endpoints of the three processes. For neutrino mass
mj maxk |mj − mk |, this subtlety is negligible.
For smaller neutrino masses, the Kurie function is
p
K(Ee ) = |Ue1 |2 [K1 (Ee )]2 + |Ue2 |2 [K2 (Ee )]2 + |Ue3 |2 [K3 (Ee )]2 , (7.76)
where Uej are elements of the leptonic mixing matrix that determine the admixture of
the mass eigenstate νj to the flavor eigenstate νe , and
r q
Kj (Ee ) ≡ C (Q − Ee ) (Q − Ee )2 − m2j θ(Q − Ee − mj ) . (7.77)
Here, C is a j-independent normalization constant. For Ee not too close to the endpoint,
we can expand
q m2j
(Q − Ee )2 − m2j ' Q − Ee − . (7.78)
2(Q − Ee )
128
7.4 Application: Majorana Neutrinos and the Seesaw Mechanism
At energies below Q − maxj (mj ) (where all the θ functions are 1), this can be written as
r q
K(Ee ) ' C (Q − Ee ) (Q − Ee )2 − m2ν̄e , (7.80)
Very close to the endpoint, this description breaks down. Instead, several kinks are
expected in the spectrum.
is
d u
W e−
ν̄e
ν̄e
W
e−
d u
129
Chapter 7 Weyl and Majorana Fermions
Sn (10100)
40000 A=136
NDS 52, 273(1987)
36000
NDS 71, 1(1994)(U) Sp (600)
Sn (11800) (3+) 3.7 s
EC
32000
(7+) 3.3 s
%p=0.09 (3.7 s) 136
%p=0.09 (3.3 s) EC 63Eu
p
28000 Sn (3500)
Sp 12010 Sn (9200)
Sp (3900) QEC(10400)
24000 0.82 s
136 0+ 47 s
51Sb β− %n=24.0
136
EC 62Sm
n+2n Sn (11070)
20000 Sp (2400)
Sn 4670
Qβ−(9300) Sp 8960 5(+),6– 107 s
EC
(2+) 47 s QEC(4500)
136
16000
Sn 8530 EC 61Pm
Sp 5540
0+ 17.5 s n
136
52Te β−
Sn 3780 QEC7850
1.1%
Sn 9940
12000 Sp 9926 Sp 4030
Qβ−5070
Sp 7197
(6–)
46.9 s –
640 β− Sn 7440
10000 (1 ) 0 Sn 8060 Sn 6828.1 Sn 9107.74 Sp 7130 0+ 50.65 m
83.4 s Sp 8594.0 136
136 Sp 5460 EC 60Nd
8000
53I β−
2+ 13.1 m
136
EC 59Pr
QEC2211
Qβ−6930
6000
5000 QEC5126
4000
IT
230+x
8– β− 114 ms
IT
0+x 1+ 0
3000 19 s 9.87 m
0+ 5+ 0
13.16 d 136 0+
EC 57La
>2.36×1021 y
136 136
55Cs β−
136
2000 54Xe 58Ce
IT
100
0
0+ 0 Evaluator: J.K. Tuli
136
56Ba
4632
Figure 7.2: Level scheme for A = 136 nuclei. Note that 136 Xe cannot undergo direct
β − decay to 136 Cs, but 0ν2β decay to 136 Ba is allowed. Similarly, 136 Ce cannot decay
via β + decay to 136 La, but 0ν2β + decay to 136 Ba is allowed.
130
7.5 Twistors
W
e−
M
e−
W
d u
This diagram is sometimes called “lobster diagram”. Since my artistic skills are not
sufficient to explain why, I have to appeal to the reader’s imagination . . .
The crucial point is that the rate of neutrinoless doble beta decay is proportional to
the neutrino masses. To see this, we must know that the W boson couples only to left-
handed particles and right-handed antiparticles. For instance, the upper vertex in te
lobster diagram can be viewed as emitting a LH electron and a RH antineutrino. The
lower vertex absorbs a LH neutrino and emits a LH electrons. This is only possible if the
neutrino experiences a chirality flip along the vertical propagator, i.e. if there is a mass
insertion (denoted by a little cross in the Feynman diagram).
While in two-neutrino double beta decay, part of the decay energy is carried away by
the neutrinos, in neutrinoless double beta decay it is all carried by the electrons and thus
visible to a detector. The telltale signature of neutrinoless double beta decay is thus a
monoergetic peak at the Q value of the decay.
7.5 Twistors
7.5.1 Unifying spinors and momentum 4-vectors
In high energy physics, one often deals with particles that are approximately massless.
For instance, at LHC energies of multiple TeV, the up, down and strange quark masses
of order MeV are completely negligible. We now introduce a formalism that allows
scattering amplitudes involving massless particles to be written in a very
P compact form.
s r r
Consider a massless spinor u (p), with s = ±. Using the spin sum r u (p)ū (p) = p /
131
Chapter 7 Weyl and Majorana Fermions
(see eq. (3.25)), we can write its outer product with itself as
1 + sγ 5 X r 1 − sγ 5 1 + sγ 5
us (p)ūs (p) = u (p)ūr (p) = p
/, (7.85)
2 r
2 2
where the projection operators (1 ± sγ 5 )/2 make sure that only the term with r = s
contributes in the spin sum. Similarly, for v spinors:
1 − sγ 5 X r 1 + sγ 5 1 − sγ 5
v s (p)v̄ s (p) = v (p)v̄ r (p) = p
/, (7.86)
2 r
2 2
Note that for massless spinors we have v s (p) = u−s (p). This follows from the explicit
expressions for the u and v spinors when taking the limit m → 0, see eqs. (3.107)
and (3.108). Therefore, in the following, it is sufficient to focus on u spinors only.
According to the Feynman rules, scattering amplitudes for massless particles depend
on spinors and on 4-momenta. We will now write the momenta in terms of spinor-like
objects as well. This will lead to some simplifications. We define
We have already mentioned in eq. (7.29) that σ̄ µα̇α = αβ α̇β̇ σβµβ̇ . Therefore, we also have
Remembering that
0 σµ
µ 1 1 0
γ = , (1 − γ 5 ) = , (7.89)
σ̄ µ 0 2 0 0
we can write
− − 0 pαα̇
u (p) ū (p) = . (7.90)
0 0
The 2-component object φα (p) is called a “twistor”. It carries a spinor index and trans-
forms accordingly under Lorentz transformation. Since twisters are not fields, but just
a numerical quantity, they are not anticommuting, though.
132
7.5 Twistors
We can repeat the above arguments also for positive helicity spinors
+ 0
u (p) ≡ , and ū+ (p) ≡ (φαT (p), 0) , (7.93)
φ∗α̇ (p)
with φα (p) = αβ φβ (p). Note that the φ(p) appearing in u+ (p) is the same as the one
appearing in u− (p) (up to the raising and lowering of indices). This follows again from
the explicit expressions eq. (3.107). We can also write p in terms of u+ (p):
+ + 0 0
u (p) ū (p) = . (7.94)
pα̇α 0
or
We have thus seen that both the particles spinor and its momentum can be expressed
in terms of the twistor. It is therefore possible to express a Feynman amplitude for
massless particles entirely in terms of twistors.
Given two 4-momentum vectors p and k, and the corresponding twistors φα (p) and κα (k),
we thus have
[p k] ≡ φα (p) κα (k)
(7.100)
hp ki ≡ φ∗α̇ κ∗α̇ .
[p k] = −[k p] . (7.101)
Similarly,
hk pi = − hp ki (7.102)
Note that these identities appear to be at odds with eq. (7.21), which stated that ψχ =
χψ. Recall, however, that eq. (7.21) holds for anticommuting spinor fields, while we
133
Chapter 7 Weyl and Majorana Fermions
are now dealing with commuting twistors. Hence the extra minus sign. The definitions
eq. (7.100) imply
hp ki = [k p]∗ . (7.103)
We can also rewrite the dot products of 4-momenta in Minkowski space in terms of the
twistor products hp ki and [p k] by using the identity
ū+ (p)P
/ u+ (k) = [p|P |ki , (7.110)
ū− (p)P
/ u− (k) = hp|P |k] , (7.111)
Pµ σαµα̇ = |P ]hP |
(7.112)
Pµ σ̄ µα̇α = |P i[P | ,
and thus
[p|P |ki = [p|α̇ Pα̇α |kiα = [pP ] hP ki ,
(7.113)
hp|P |k] = hp|α P αα̇ |k]α̇ = hpP i [P k] .
134
7.5 Twistors
7.5.3 Examples
As a first example, consider the annihilation of two massless fermions into two scalars in
Yukawa theory:
e− , p1 φ, p3 e− , p1 φ, p3
e+ , p2 φ, p4 e+ , p2 φ, p4
/1 − p
i(p /3 ) /1 − p
i(p /4 ) s
2 s0
= g v̄ (p2 ) + u (p1 )
(p1 − p3 )2 (p1 − p4 )2
ip ip
0 /3 /4
= −g 2 v̄ s (p2 ) + us (p1 ) . (7.114)
2p1 p3 2p1 p4
Note that, for brevity, we have replaced pj → j here. Moreover, we have used eq. (7.107)
to rewrite the contraction of the 4-momenta in the denominators in terms of twistors,
eq. (7.113) to rewrite the momenta contracted with gamma matrices, and eqs. (7.101)
and (7.102) to reorder the momenta inside twistor products.
The asmplitude for s = s0 = − is obtained in a similar way by replacing square brackets
by angle brackets and vice-versa. Using this observation, we can directly read it off from
eq. (7.115):
−− 2 h23i h24i
A = −ig + . (7.116)
h13i h14i
If, on th other hand, s = s0 , we would obtain mixed products of left-chiral and right-chiral
spinors like hpk], which vanish.
135
Chapter 7 Weyl and Majorana Fermions
136
Acknowledgments
I would like to thank the students attending my lectures, in particular Moritz Breitbach,
for their critical reading of these lecture notes, and for their comments which helped to
improve them.
137
Chapter 7 Weyl and Majorana Fermions
138
Bibliography
139