Ap 3
Ap 3
Ap 3
This course gives an introduction to the ingredients of gauge theories which are necessary
to calculate cross sections for particular processes. The section headings are given below:
Outline of Lectures:
2. Spin
3. Relativistic Electromagnetism
4. Coulomb Scattering, eµ → eµ
5. Compton Scattering, eγ → eγ
6. Colour
7. Renormalisation
This course runs in parallel with the Quantum Field Theory course, from which we will
use some results. Some topics mentioned in this course will be covered in more detail in
the Standard Model and Phenomenology courses next week.
Textbooks:
These notes are intended to be self-contained, but only provide a short introduction to a
complex and fascinating topic. You may find the following textbooks useful:
The first two are more practical and closer to the spirit of this course while the other contain
many more mathematical details. The last one is very recent. If you are particularly
interested in (or confused by) a particular topic, I encourage you to take a look at it. If
there are other textbooks which you find particularly helpful, please tell me and I will
update the list.
These notes are based heavily on the content of previous versions of this course, in partic-
ular the 2013 version by Jennifer Smillie. Throughout, we will use “natural units” where
~ = c = 1 and the metric signature (+ − −−).
Please email any comments, questions or corrections to a.banfi@sussex.ac.uk.
Andrea Banfi
May 25, 2016
1 Relativistic Quantum Mechanics
In order to describe the dynamics of particles involved in high-energy collisions we must
be able to combine the theory of phenomena occurring at the smallest scales, i.e. quantum
mechanics, with the description of particles moving close to the speed of light, i.e. special
relativity. To do this we must develop wave equations which are relativistically invariant
(i.e. invariant under Lorentz transformations). In this section we will derive relativistic
equations of motion for scalar particles (spin-0) and particles with spin-1/2.
p2
E= + V (x) . (1)
2m
To convert this into a wave equation, we make the replacements E → i∂t and p → −i∇,
so that a plane-wave solution
has the energy-momentum relation given in eq. (1). Applied to a general wavefunction φ,
a linear superposition of plane waves, this gives
1 2
i∂t φ(t, x) = − ∇ + V (x) φ(t, x) = H φ(t, x) , (3)
2m
where H is the so-called Hamiltonian. We recognise this as the Schrödinger Equation, the
cornerstone of Quantum Mechanics. From this form, we can deduce that eq. (3) cannot
be relativistically invariant because time appears only through a first-order derivative on
the left-hand side while space appears as a second-order derivative on the right-hand side.
Yet we know that if we make a Lorentz transformation in the x direction for example, this
would mix the x and t components and therefore they cannot have different rôles.
The problem with the Schrödinger Equation arose because we started from a non-relativistic
energy-momentum relation. Let us then start from the relativistic equation for energy.
For a particle with 4-momentum pµ = (E, p) and mass m,
E 2 = m 2 + p2 . (4)
Again we convert this to an operator equation by setting pµ = i∂µ so that the corresponding
wave equation for an arbitrary scalar wavefunction φ(x, t) gives
However, the Klein-Gordon equation requires that the right-hand side is equal to [−∇2 +
m2 ]ψ(t, x) and therefore α and β must satisfy
αi αj + αj αi = {αi , αj } = 2δ ij , βαi + αi β = {αi , β} = 0, β2 = 1 . (11)
If αi and β are just numbers, these equations cannot be solved. Dirac solved them by
instead taking αi and β to be n × n matrices, and ψ(t, x) to be a column vector. Even
now, the solution is not immediate. One can show that the conditions in eq. (11) require
Tr αi = 0 = Tr β, (12)
and further that the eigenvalues of the above matrices are ±1. This in turn means that n
must be even (do you understand why?). In 2-dimensions, there are still not enough lin-
early independent matrices to satisfy eq. (11). There do exist solutions in four dimensions.
One such solution is
0 σ 12 0
α= , β= , (13)
σ 0 0 −12
where σ are the usual Pauli matrices and 12 represents the 2 × 2 identity matrix. Now we
have formed an equation which may be thought of as a square-root of the Klein-Gordon
equation, but which is not obviously Lorentz invariant. To show that, we first define the
new matrices
γ 0 = β, γ = βα . (14)
Then we form γ µ = (γ 0 , γ) where the µ is a Lorentz index. Each component is a 4 × 4
matrix. In terms of the γ-matrices, one can write the conditions in eq. (11) in a Lorentz
covariant form
{γ µ , γ ν } = γ µ γ ν + γ ν γ µ = 2g µν . (15)
This is an example of a Clifford algebra. Any matrices satisfying this condition in eq. (15)
may be used to construct the Dirac equation. The representation in eqs. (13) and (14) is
just one example, known as the Dirac representation. Note, for example, that any other
matrices satisfying
αi0 = U αi U −1 , and β 0 = U βU −1 , (16)
where U is a unitary matrix, will also be suitable.
Multiplying through by γ 0 , we may rewrite the eq. (9) in a covariant form as
(iγ µ ∂µ − m14 )ψ(t, x) = (i∂/ − m)ψ(x) = 0 , (17)
where a, a vector with a slash, is a short-hand notation for γ µ aµ . The equation above
is known as the Dirac equation. In momentum space, i.e. after a Fourier transformation,
∂µ → −ipµ , and the Dirac equation becomes
(γ µ pµ − m14 )ψ̃(p) = (p − m)ψ̃(p) = 0 , (18)
where ψ̃(p) is the Fourier transform of a solution of the Dirac equation ψ(x).
We mentioned in passing that ψ(t, x) is a column vector rather than a scalar. This
means that it contains more than one degree of freedom. Dirac exploited this property to
interpret his equation as the wave equation for spin-1/2 particles, fermions, which can be
either spin-up or spin-down. The column vector ψ is known as a Dirac spinor.
Comparing eq. (9) to the Schrödinger equation in eq. (3) gives the Hamiltonian for a free
spin-1/2 particle:
HDirac = −iα · ∇ + β m . (19)
E ..
.
+m
−m
..
.
Figure 1: The energy levels in the Dirac sea picture. They must satisfy |E| > m, but
negative-energy states are allowed. The vacuum is the state in which all negative-energy
levels are filled.
The trace of the Hamiltonian gives the sum of the energy eigenvalues. The condition
that the matrices α and β are traceless therefore means that the eigenvalues of HDirac
must sum to zero. Therefore, like the Klein-Gordon equation, also the Dirac equation has
negative-energy solutions.
Dirac himself proposed a solution for this problem which became known as the “Dirac
sea”. He accepted the existence of negative-energy states, but took the vacuum as the
state in which all these states are filled, see fig. 1. There is a conceptual problem with
this in that the vacuum has infinite negative charge and energy. However, any observation
relies only on energy differences, so this picture can give an acceptable theory.
As the negative-energy states are already full, the Pauli exclusion principle forbids any
positive-energy electron to fall into one of the negative-energy states. If instead energy
is supplied, an electron is excited from a negative-energy state to a positive-energy state
and an “electron-hole” pair is created. The absence of the negative-energy electron, the
hole, is interpreted as the presence of of state with positive energy and positive charge,
i.e. a positron. Dirac predicted the existence of the positron in 1927 and this particle was
discovered five years later.
However, Dirac’s argument only holds for spin-1/2 particles which obey the Pauli exclusion
principle. A consistent solution for all particles is provided by Quantum Field Theory
in a picture developed by Feynman and Stückelberg, in which positive-energy partices
travel only forward in time, whereas negative-energy particles travel only backwards in
time. In this way, a negative-energy particle with momentum pµ , travelling backward in
time, is re-interpreted as a positive energy anti-particle with momentum −pµ travelling
forward in time. Let us see how this picture naturally arises by considering two processes,
the scattering e− µ− → e− µ− , and Compton scattering e− γ → e− γ. In non-relativistic
quantum mechanics, the scattering e− µ− → e− µ− corresponds to the scattering of an
electron from an external Coulomb potential. This is represented on the left-hand side of
fig. 2. The horizontal axis represents the time at which a give elementary process occurs.
x x x
e− e− e− e−
e− e−
µ− µ− µ− µ− µ−
t t1 t2 t t1 t2 t
t1= t 2
2 Spin
In the previous section, we introduced a Dirac spinor as a solution to the Dirac equation
in the form of a column vector. In this section, we will discuss the explicit form of
the solutions to the Dirac equation, and verify that they indeed correspond to the wave
functions for particles with spin-1/2.
x x
e− γ e− γ
γ
γ e−
t1 t2 t t1 t2 t
where r = 1, 2 and N is√a normalisation conventionally chosen such that u†r (p)us (p) =
2E δ rs , which gives N = E + m. The spinors χ1 and χ2 cover the two (spin) degrees of
freedom:
1 0
χ1 = , and χ2 = . (26)
0 1
2.2 Spin
Each Dirac spinor has two linearly independent solutions which we stated earlier corre-
sponded to the two possible spin states of a fermion. In this subsection we will define the
corresponding spin operator. If we again consider a particle at rest we have
1 0
0 1
u1 =
0 ,
and u2 =
0 .
(29)
0 0
This operator satisfies h(p)2 = 1, and hence its eigenvalues are ±1.
with S(Λ) a suitable 4 × 4 matrix. Its explicit form is derived by imposing that the Dirac
equation is Lorentz invariant:
so that ψ 0 (x0 ) is a solution of the transformed Dirac equation, provided ψ(x) is a solution
of the original one.
Eq. (40) is enough to construct the matrices S(Λ). By direct inspection one observes that
The fact that S −1 (Λ) 6= S † (Λ) is not surprising, and is due to the fact that the Lorentz
group is non-compact, and therefore it does not admit unitary finite-dimensional repre-
sentations.
One can construct bi-linear products ψ̄Γψ, with Γ a 4 × 4 matrix. We now show that Γ
can be decomposed into a set of bi-linears, each having a definite transformation property
under the Lorentz group. Since Γ is 4 × 4 matrix, we expect to find 16 such bi-linear
products, constructed out of linearly independent matrices. Already we can find 5 such
bi-linears:
ψ̄ ψ → ψ̄S −1 (Λ)S(Λ)ψ = ψ̄ ψ (scalar) ,
(43)
ψ̄ γ µ ψ → ψ̄S −1 (Λ)γ µ S(Λ)ψ = Λµν ψ̄ γ ν ψ
(vector) ,
which satisfies
(γ 5 )2 = 1, {γ 5 , γ µ } = 0, (γ 5 )† = γ 5 . (47)
The factor of i is to make then matrix Hermitian. Using γ 5 , we can construct 5 more
bi-linears
ψ̄ γ 5 ψ → ψ̄ S −1 (Λ)iµνρσ γ µ γ ν γ ρ γ σ S(Λ)ψ
= i µνρσ Λµα Λνβ Λργ Λσδ ψ̄ γ α γ β γ γ γ δ ψ
(48)
= det(Λ) ψ̄ i αβγδ γ α γ β γ γ γ δ ψ = det(Λ) ψ̄ γ 5 ψ (pseudo-scalar) ,
ψ̄ γ 5 γ µ ψ → det(Λ) Λµν ψ̄ γ 5 γ ν ψ
(pseudo-vector) .
We have then found a set of 16 linearly independent matrices (check that they are linearly
independent!)
i
1, γ 5 , γ µ , γ µ γ 5 , Σµν = [γ µ , γ ν ] , (49)
4
so that any bi-linear ψΓψ can be written as a sum of terms with definite transformation
properties, i.e. transforming in a clear way as a scalar, pseudo-scalar, vector, pseudo-
vector and tensor. (This is why the Feynman rule for a pseudo-scalar interacting with a
particle-anti-particle pair has a γ 5 for example.)
The most common use of γ 5 is in the projectors PL = (1 − γ 5 )/2 and PR = (1 + γ 5 )/2.
You can check explicitly that these behave like projectors (ie. P 2 = P and PL PR = 0).
When these act upon a Dirac spinor they project out either the component with “left-
handed” chirality or with “right-handed” chirality. These projectors therefore appear
when considering weak interactions, for example, as W bosons only couple to left-handed
particles. One has to take care when defining the handedness of antiparticles because
3 Quantum Electro-Dynamics
In this section, we will develop the theory of quantum electro-dynamics (QED) which
describes the interaction between electrically charged fermions and a vector field (the
photon Aµ ).
3.1 The QED Lagrangian
In this course, we have so far considered spin-0 and spin-1/2 particles. We will postpone a
detailed discussion of spin-1 particles until section 5.1. For the time being, we start from
the Maxwell’s equations in the vacuum in relativistic notation:
∂µ F µν = J ν , where F µν = ∂ µ Aν − ∂ ν Aµ , (51)
The Dirac equation for ψ and its equivalent for ψ can be derived from the Lagrangian
The starting point for the QED Lagrangian is then the sum of Lem and LDirac . However,
in order to make the theory describe interactions, we must include a term which couples
Aµ to ψ and ψ. If we wish Maxwell’s equation to be valid, this term has to be of the
form Lint = −J µ Aµ , with J µ a conserved vector current. We then observe that the vector
current J µ = ψ̄ γ µ ψ is conserved if ψ is a solution of Dirac equation. In fact
←
−
∂µ J µ = ψ̄
∂ ψ + ψ̄ (∂ψ) = (−mψ̄) + ψ̄ (mψ) = 0 . (55)
ψ(x) → ψ 0 (x) = e−ieα(x) ψ(x) , Aµ (x) → A0µ (x) = Aµ (x) + ∂µ α(x) . (58)
Notice that the addition of the interaction term Lint is equivalent to the replacement
∂µ → Dµ = ∂µ − ieAµ . (59)
This prescription is known as “minimal coupling” and automatically ensures that the
Lagrangian is gauge invariant. The use of gauge invariance to introduce interactions will
be covered in detail in the Standard Model course next week. This gives
1
L = − F µν Fµν + ψ(iγ µ (∂µ + ieAµ )ψ . (60)
4
The fact that L is invariant under the gauge transformations in eq. (62) means that Aµ
contains unphysical degrees of freedom. This is clear in view of the fact that a massless
vector field contains two physical polarisations, whereas Aµ has four degrees of freedom. In
order to eliminate this degeneracy, a “gauge-fixing” condition is imposed. A possible choice
of a gauge condition is the so-called Coulomb gauge, in which ∇ · A = 0. Although this
condition eliminates the two additional degrees of freedom, it breaks Lorentz covariance.
A common choice that preserves Lorentz covariance is the Lorentz gauge:
∂µ Aµ = 0. (61)
This corresponds to choosing the gauge parameter α such that α = −∂µ Aµ above. In
this gauge, the Maxwell equations become Aν = 0.
Notice that the Lorentz gauge condition reduces the number of degrees of freedom in A
from four to three. Even now though Aµ is not unique. A transformation of the form
Aµ → A0µ = Aµ + ∂µ χ , χ = 0 , (62)
will also leave the Lagrangian unchanged. At classical level we can eliminate the extra
polarisation “by hand”, but at quantum level this cannot be done without giving up
covariant canonical commutation rules. The way out, which can only be summarised, is
to add a gauge-fixing Lagrangian Lgf , so that the full QED Lagrangian becomes
1
LQED = Lem + LDirac + Lint + Lgf , Lgf = − (∂µ Aµ )2 . (63)
2ξ
Using this Lagrangian as a starting point, and an extra condition on physical states, only
the two physical polarisations propagate on-shell. Notice that setting ξ = 0 corresponds
to enforcing the Lorentz gauge condition ∂µ Aν = 0, otherwise the equations of motions
give ∂µ Aν = 0, i.e. ∂µ Aν is a free field.
Outgoing
p
i(p + m) u(p)
p →
p2 − m2 + iε
v(p)
p →
µ ν
−i pµ pν
µν
g − (1 − ξ) 2 ε∗µ (p)
p
2
p + iε p p → µ
Figure 4: The Feynman rules for QED. Wavy lines represent a photon and straight lines
represent any charged fermion. The arrow on the straight line tells you it is a particle or
anti-particle depending on whether it is with or against momentum flow. The polarisation
vectors εµ (p) will be discussed in section 5.2.
In the quantum field theory course at this school, you learn how to derive the “Feynman
rules” for scalar φ4 theory. The principles are the same here so in this course we will state
the Feynman rules for QED and learn how to work with them. The Feynman rules are
shown in figure 4. The left-hand column represents internal parts of the diagram while
the right-hand column gives the rules for external fermions and photons.
A few comments are necessary here:
2. The photon propagator term has a free parameter ξ. This is due to the gauge
freedom we discussed in the previous section. It does not represent a physical degree
of freedom and therefore any calculation of a physical observable will be independent
of ξ. We will most commonly work in Feynman gauge ξ = 1.
3. The propagators come with factors of iε in the denominator, otherwise they would
have poles on the real axis and any integral over p would not be well-defined. The
e(p) e(p′ ) e(p) e(p′)
↓q
µ(k) µ(k ′ )
µ(k) µ(k ′ )
Figure 5: Building the leading-order Feynman diagram for Coulomb scattering. We start
from the initial and final states on the left-hand side. The diagram on the right is the only
way to connect these with up to two vertices.
factor of iε prescribes which direction to travel around the poles. This choice corre-
sponds to the “Feynman prescription”, which ensures causality.
4. The interaction vertex contains only one flavour of fermion. We know that the
emission of a photon does not change an electron to a quark for example.
We start by drawing the external particles, see left-hand side of fig. 5. We now want to
find all possible ways to connect these. There is no direct interaction between an electron
and a muon but both interact with a photon, so a possible connected diagram is the one
shown on the right-hand side. In fact, this is the only possible diagram with no more than
two vertices. The number of vertices is directly related to the powers of the coupling e
and therefore the diagram shown on the right is the leading-order (or tree-level) process.
If we consider e(p) e(k) → e(p0 ) e(k 0 ) or e+ (p) e− (k) → e+ (p0 ) e− (k 0 ) instead, there are two
diagrams with two vertices, i.e. at O(e2 ) (try this!). Both have to be added before squaring
the amplitude to have the tree-level contribution to the cross section.
If we allow ourselves more than two vertices, there are many more diagrams we can draw.
Since the number of external particles doesn’t increase, these must contain closed loops
and, therefore, they represent higher-loop processes. In this course, we will limit ourselves
to tree-level processes. Loop-diagrams will be covered in the phenomenology course.
Now we will construct the tree-level amplitude for Coulomb scattering from the rules in
Fig. 4. Keeping in mind the earlier warning about the ordering of matrices and spinors,
we take each fermion line in turn. The electron line gives
[ur0 (p0 )γ ρ ur (p)]∗ = u†r (p)γ ρ† γ 0† ur0 (p0 ) = u†r (p)γ 0 γ ρ ur0 (p0 ) = ur (p)γ ρ ur0 (p0 ) , (70)
where we have used γ ν† = γ 0 γ ν γ 0 , which you showed on the problem sheet. We now use
eq. (37) to find
X X
[ur0 (p0 ) γ µ ur (p)][ur0 (p0 ) γ ρ ur (p)]∗ = ur0 (p0 ) γ µ ur (p)ur (p)γ ρ ur0 (p0 )
r,r0 r,r0
X (71)
= ur0 (p0 )γ µ (p + m) γ ρ ur0 (p0 ) .
r0
We will use m for the electron mass and M for the muon mass. It is now useful to add a
component index in spinor-space like you would do in normal linear algebra. Schematically
we have
X
ur0 i Γij ur0 j , (72)
r0
where Γ represents the chain of γ-matrices in eq. (71). Now that we are explicitly labelling
the components, we can swap the order of the terms to get
X
Γij ur0 j ur0 i = Γij (p 0 + m)ji = Tr(γ µ (p + m) γ ρ (p 0 + m)) . (73)
r0
We could have anticipated that we would get a trace as we need to get a single number
from a series of matrices. Working from the anti-commutation relations, one can readily
show the following identities (see problem sheet):
Tr(odd number of γ matrices) = 0 , Tr(γ µ γ ν ) = 4g µν ,
(74)
Tr(γ µ γ ν γ ρ γ σ ) = 4(g µν g ρσ − g µρ g νσ + g µσ g νρ ) .
Therefore, eq. (73) equals
4pν p0σ (g µν g ρσ − g µρ g νσ + g µσ g νρ ) + 4m2 g µρ . (75)
The same series of steps gives
X
[us0 (k 0 ) γµ us (k)][us0 (k 0 ) γρ us (k)]∗ = 4k α k 0β (gµα gρβ − gµρ gαβ + gµβ gαρ ) + 4M 2 gµρ .
s,s0
(76)
Substituting these results into eq. (69) gives
2 8e4
(pk) (p0 k 0 ) + (pk 0 )(p0 k) + 2m2 M 2 − M 2 (pp0 ) − m2 (kk 0 ) .
|M| = (77)
(q 2 )2
We will now rewrite the invariants which appear in the above equation in terms of the
centre-of-mass energy squared, s and the exchanged momentum-squared, q 2 = t. We have
2(pk) = (p + k)2 − m2 − M 2 = s − m2 − M 2 , 2(p0 k 0 ) = s − m2 − M 2
2(pp0 ) = −(p − p0 )2 + 2m2 = −q 2 + 2m2 , 2(kk 0 ) = −q 2 + 2M 2 (78)
2(pk 0 ) = 2p·(p + k − p0 ) = s + q 2 − m2 − M 2 , 2(p0 k) = s + q 2 − m2 − M 2 ,
which finally gives
2 2e4
(s − m2 − M 2 )2 + (s + q 2 − m2 − M 2 )2 + 2q 2 (m2 + M 2 ) .
|M| = (79)
(q 2 )2
This expression can be further simplified by introducing the further invariant u = (p −
k 0 )2 = (p0 − k)2 :
2 2e4
(s − m2 − M 2 )2 + (u − m2 − M 2 )2 + 2t (m2 + M 2 ) .
|M| = (80)
t2
The above equation gives the probability that the corresponding process occurs at a given
point in phase space. In the next section, we will derive how to calculate a total cross
section (or a total decay width) from amplitudes squared.
q
where Ef = p2f + m2 . Although the final expression explicitly separates the depen-
dence on E and p, it is still Lorentz invariant as the original expression is clearly Lorentz
invariant. Eq. (81) is frequently referred to as the Lorentz Invariant Phase Space mea-
sure (LIPS). The factors of 2π correspond to the conventions used for momentum space
integrations in QFT.
We now need to normalise this expression to the flux of incoming particles. This is done
by multiplying by the flux factor, F. For the scattering of two incoming particles, this is
usually written as
1
F= , (82)
4Ea Eb |v a − v b |
where Ei and v i are the energy and velocity of each incoming particle.1 A neater, equivalent
form which explicitly demonstrates the Lorentz invariance of this quantity is
1
F= p . (83)
4 (pa pb )2 − m2a m2b
In the massless limit s m1 , m2 , this simplifies to F ' 1/(2s). Finally, we must impose
total conservation of momentum to find
! !
Y Z d3 pf 2 X
σ=F 3 (2E )
|M| (2π)4 δ 4 pf − p1 − p 2 . (84)
f
(2π) f
f
If you wish to calculate a total decay width instead, the expression is very similar. The
only difference is that the flux factor becomes
1
F= , (85)
2M
where M is the mass of the decaying particle. The total decay width, Γ, is therefore given
by
! !
1 Y Z d 3 pf 2 4 4
X
Γ= |M| (2π) δ pf − pM . (86)
2M f
(2π)3 (2Ef ) f
d3 p 0 d3 k0
Z
2
σ=F 3 0 3 0
|M| (2π)4 δ 4 (p0 + k 0 − p − k) . (87)
(2π) (2Ep ) (2π) (2Ek )
As this expression is Lorentz invariant, we are free to choose which frame to evaluate it
in. This is an extremely powerful tool to evaluate these integrals, as a careful choice can
lead to considerable simplifications. We will choose the centre-of-mass frame here so that
p = −k. We can easily do the the k0 integration using three of the δ-functions to give
Z 3 0
dp 1 2
|M| (2π)δ Ep0 + Ek0 − Ep − Ek .
σ=F 3 0 0
(88)
(2π) 4Ep Ek
1
You can find a motivation for the flux factor in Aitchison and Hey and a more complete derivation in
Peskin and Schröder chapter 4.5.
We will proceed by transforming to spherical polar coordinates, d3 p0 = |p0 |2 d|p0 |dΩ, where
we have written the solid angle, sin θ dθ dφ, as dΩ:
F |p0 |2
Z
0 2 0 0
σ= dΩd|p | 0
|M| δ E p + Ek − Ep − Ek . (89)
(2π)2 4Ep0 Ek
We now make the change of variable |p0 | → E = Ep0 + Ek0 , which has Jacobian factor
∂E E|p0 |
= (90)
∂|p0 | Ep0 Ek0
to get
F |p0 | √ F |p0 |
Z Z
2 2
σ= dΩdE |M| δ E − s = dΩ √ |M| , (91)
(2π)2 4E (2π)2 4 s
√
where it is understood that k0 = −p0 with |p0 | determined from E = s. The only
undefined variables are the angles which remain to be integrated over. We could now
2
substitute the expression for |M| explicitly in terms of these angles but it is actually
informative to instead study the differential cross section
dσ F |p0 | 2
= 2
√ |M| . (92)
dΩ 16π s
We will now consider the high energy limit where s m2e , m2µ . In this limit, the three
Mandelstam invariants are given by
which gives
2
+ u2 2e4
2 4s 4 θ
|M| ' 2e = 1 + cos . (94)
t2 sin4 (θ/2) 2
Note that this amplitude squared has no dependence on the azimuthal angle φ. Using the
conventional notation α = e2 /(4π), we obtain
dσ α2 1 + cos4 (θ/2)
' . (95)
dΩ 2s sin4 (θ/2)
θ
k = (Ek , k)
dσ α2
1 − v 2 sin2 (θ/2)
= 2 2 4
dΩ 4k v sin (θ/2)
(96)
dσ
1 − v 2 sin2 (θ/2) ,
=
dΩ R
α2
dσ
= (97)
dΩ R 4k2 v 2 sin4 (θ/2)
is the Rutherford cross section which was calculated in preschool problem 9. The extra
v 2 -term in eq. (96) then gives the relativistic correction to this. This result is entirely due
2
to the electron being a spin-1/2 particle. If it were spin-0 instead, |M| would look much
simpler as there are no fermion traces to be performed and in that case we would find that
there is no relativistic correction.
4.4 e+ e− Annihilation
The calculation we have just performed is almost identical to e+ (p0 ) e− (p) → µ+ (k) µ− (k 0 ).
Although this now involves anti-particles, there is still one single diagram at leading-order
and the trace algebra is very similar. Indeed we can re-interpret the incoming e+ as an
outgoing e− with momentum −p0 , and the outgoing µ+ as an incoming µ− with momentum
−k. Then we do find explicitly that
2 2
|Me+ (p0 )e− (p)→µ+ (k0 )µ− (k) | = |Me− (p)µ− (−k)→e− (−p0 )µ− (k0 ) | . (98)
plus higher-order corrections, where there are Nc colours in each of the nf massless flavours
of quarks with charge Qi . Therefore the ratio
σ(e+ e− → µ+ µ− )
R= (105)
σ(e+ e− → hadrons)
has been used to measure the number of colours to be Nc = 3.
5 Photon Scattering
In this section we will calculate the scattering amplitude for eγ → eγ. In order to do that
we need first to consider how to treat incoming and outgoing photons.
Aµ = 0 , (107)
and is automatically satisfied by a solution of the form in eq. (106), provided k 2 = 0. The
Lorentz gauge condition gives an additional constraint on the polarisation vector
k·ε(k) = 0 . (108)
However, there is still freedom here because, given a polarisation vector ε which solves this
equation, any other vector of the form ε0 = ε + λ k will also be a solution, which corre-
sponds to the propagation of an extra unphysical longitudinal photon, with a polarisation
proportional to kµ . This freedom is usually used to set ε0 = 0 such that k·ε = 0 so that
the two physical polarisations εα , with α = 1, 2, are in the transverse direction, and are
chosen to be orthonormal. A useful relation we will use in the following is
2
X ki ki
εαi (k)εαj (k) = δ ij − k̂ i k̂ j , where k̂ i = = 0. (109)
α=1
|k| k
The Feynman rule for an incoming photon is simply εµ (k) while for an outgoing photon
it is ε∗µ (k), as shown in Fig. 4.
As for fermion spins, for unpolarised processes you compute the total cross section by
averaging over incoming polarisations and summing over outgoing polarisations. Let us
consider the case of a general process with one external incoming photon. The matrix
element would have the form
iM = Aµ εµ (k) . (110)
The left-hand side is a physical quantity, hence it should give the same result for any
choice of the gauge. Had we chosen ε + λ k instead, this implies that Aµ kµ has to vanish.
This is a “Ward Identity” for QED, and is therefore a test of gauge-invariance.
Squaring the scattering-amplitude over the physical polarisations gives
2
X 2
X
|Aµ εαµ (k)|2 = Aµ A∗ν εαµ (k) ε∗α
ν (k)
α=1 α=1
(111)
= A A (δ ij − k̂ i k̂ j ) ,
i j
p p′
k k′
k p′
Figure 7: The two tree-level diagrams for e(p) γ(k) → e(p0 ) γ(k 0 ).
This could be done for each photon in turn if there were more in the process, and we find
the general result that
2
X
εαµ ε∗α
ν → −gµν . (113)
α=1
We have used the → notation of Peskin and Schröder here as the result is not an exact
equality in the absence of the rest of the matrix element, but the result is nonetheless true
in any practical calculation.
k0 + m
p −
2 ∗0µ 0 ν 0 p +
k+m
iM = −ie ε (k ) ε (k) u(p ) γµ γν + γν γµ u(p) . (114)
(p + k)2 − m2 (p − k 0 )2 − m2
You can check explicitly that the above amplitude does indeed satisfy the appropriate QED
Ward Identities, i.e. replacing εν (k) with k ν gives M = 0, and similarly when replacing
ε∗0µ (k 0 ) with k 0µ (see tutorial sheet).
We now square the amplitude to get
2 1X 1 X
|M| = |M|2
2 γ pol 2 e spin
2 ! (115)
(pk 0 )
4 (pk) 1 1 1 1
= 2e + + 2m2 − +m 4
− .
(pk 0 ) (pk) (pk) (pk 0 ) (pk) (pk 0 )
The calculation of the spin traces in this case requires the identities
γµ γ µ = 4 , γµ γ ρ γ µ = −2γ ρ , (116)
k ′ = (ω ′ , k′ )
p = (m, 0)
θ
k = (ω, k)
p′ = (E ′ , p′ )
Figure 8: The Compton scattering process in the rest frame of the incoming electron.
from the problem sheet. We will again choose a suitable reference frame to simplify the
calculation. In this case, it is convenient to work in the rest frame of the incoming electron
as shown in fig. 8. We can use energy conservation to compute ω 0 :
m2 = p02 = (p + k − k 0 )2 = m2 + 2m(ω − ω 0 ) − 2ωω 0 (1 − cos θ)
ω (117)
⇒ ω0 = .
1 + (ω/m)(1 − cos θ)
In this frame, we therefore have
ω0
2 4 ω
|M| = 2e + − sin2 θ . (118)
ω0 ω
The explicit dependence on the electron mass cancels with the factors of m in ω 0 . It is
however present in the flux factor F = 1/(4mω). We now compute the integral over the
phase space to get
d3 p 0 d3 k 0 ω0
1 ω
Z
σ= 2e 4
+ − sin θ (2π)4 δ 4 (p0 + k 0 − p − k) .
2
4mω (2π)3 (2E 0 ) (2π)3 (2ω 0 ) ω0 ω
(119)
We can again do the integral over d3 k0 using the spatial parts of the δ-function. Then we
transfer to spherical polars and find
0 2
α2 ω0
dσ ω ω 2
= + − sin θ . (120)
dΩ 2m2 ω ω0 ω
A nice check of this result is to take the low-energy limit where ω m. Then ω ' ω 0 and
we find
dσ α2 2
= 1 + cos θ . (121)
dΩ 2m2
This is the Thomson cross section for the scattering of classical electromagnetic radiation
by a free electron. In the other limit, the high-energy limit where ω m, we have
0 m dσ α2 1
ω ' ⇒ ' . (122)
1 − cos θ dΩ 2mω 1 − cos θ
and the cross section is strongly peaked for small angles. This leads to a logarithmic
enhancement when you perform the angular integration. These “collinear” logarithms
arise whenever massless particles are emitted; this will be discussed in more detail in the
phenomenology course.
Note that, since ω > ω 0 , eq. (122) holds strictly for (1 − cos θ) > m/ω. For smaller angles,
eq. (121) holds and
dσ α2
' 2. (123)
dΩ m
The forward (small scattering angle) Compton scattering cross section is then a valuable
method to measure the QED coupling α.
6 Strong Interactions
In this section we will develop the theory of the strong interactions, quantum chromo-
dynamics (QCD). The major difference between QED and QCD is that the gluons are
self-interacting because they also carry colour charge (unlike the charge-neutral photon).
Spin-1/2: six families of quarks (up, charge and top with electric charge +2/3;
down, strange and bottom with electric charge -1/3)
For each flavour, there are Nc = 3 of these.
The a, i and j indices are gauge group indices which are discussed further below. The
sum over these is implicit in eq. (124). Each ta is a 3 × 3 matrix in colour space. The ta
matrices do not commute with each other, but obey the following algebra
The QCD Lagrangian LQCD is invariant under the infinitesimal “gauge” transformations
ψi (x) → δij − igs θa (x)taij ψ(x) ,
(129)
Aaµ (x) → Aaµ (x) + Dµab θb (x) ,
where Dµab is the covariant derivative in the “adjoint” representation, the one under which
the gluon fields transforms under SU (3), as opposed to the “fundamental” representa-
tion, which rules the transformation of quark fields. In particular, the adjoint covariant
derivative is given by
Dµab = ∂µ δ ab + igs Acµ (T c )ab , (T c )ab = if acb = −if abc . (130)
The matrices T a , as needed for any generator of a representation of SU (3), satisfy the
same commutation rules as ta :
[T a , T b ] = if abc T c . (131)
These are nothing else than the Jacobi identity satisfied by the structure constants f abc :
Notice that the gauge transformation for Aaµ involves the strong coupling gs :
and only at lowest order in gs does it reduce to the analogous transformation for QED.
As in QED, in order to quantise the QCD Lagrangian, we need to introduce a “gauge-
fixing” term, for instance
1
Lgf = − (∂µ Aµa )2 . (134)
2ξ
We now describe the Feynman rules for QCD. The quark and gluon propagators are
identical to those for QED except they are also accompanied by the appropriate delta-
function in colour space (see fig. 9). The coupling between two quarks and a gluon is now
i(p + m)
p→ δkj
j k
p2 − m2 + iε
µ p→ ν −ig µν
a
δ ab
b p2 + iε
Figure 9: The Feynman rules for the quark and gluon propagators (the latter is in Feynman
gauge ξ = 1).
given in terms of colour matrices, taij as shown in fig. 10. Notice that the Dirac matrix
γ µ also still appears as it must for spin-1/2 particles. The colour matrices and Dirac
matrices do not interact with each other (they act on different vector spaces). The ‘a’
is the “adjoint” index and is associated with the gluon. The k and j are “fundamental”
indices associated with the outgoing and incoming fermion line respectively.
j k
−igs γ µ takj
µ igsγ µ Tkj
a
qν
gs f abc (g µν (p − q)ρ + g νρ (q − r)µ + g ρµ (r − p)ν )
pµ
rρ
c
−igs2 f abe f cde (g µρ g νσ − g µσ g νρ )
a b
+f ace f bde (g µν g ρσ − g µσ g νρ )
+f ade f bce (g µν g ρσ − g µρ g νσ )
d c
Figure 11: Three and four gluon vertices which arise from eq. (124). All momenta are
taken to be incoming.
a
Returning to the Lagrangian, in QCD Fµν has an extra term compared to QED, as required
by gauge invariance. (Technically this term is present for QED too, but QED is an
“Abelian” gauge theory which means that the structure constants are zero). Multiplying
out F aµν Fµν
a
give extra terms with 3 and 4 gauge fields. These correspond to new three-
and four-gluon vertices as shown in fig. 11.
p k p k p k
p’ k’ p’ k’ p’ k’
for this process can be obtained from that of Compton scattering via crossing.
One immediate effect is obvious – there is now a third diagram including the three-gluon
vertex. If we sum the contributions from the first two diagrams we find
M(a)+(b) = A(a)+(b)
µν ε∗µ (k)ε0∗ν (k 0 ),
− −k0 (135)
(a)+(b) 2 0 b p k a a p b
Aµν = −igs v(p ) γν t γµ t + γµ t γν t u(p) ,
(p − k)2 (p − k 0 )2
where we have implicitly assumed that gluon k has colour a and polarisation index µ, and
gluon k 0 has colour b and polarisation index ν. At this order, see eq. (133), gauge invari-
ance corresponds to testing whether the replacement µ → µ + λkµ leaves the amplitude
(a)+(b) µ
invariant. This is equivalent to testing the condition for the Ward Identity, Aµν k = 0:
A(a)+(b)
µν k µ = −igs2 [ta , tb ]v(p0 )γν u(p) 6= 0 . (136)
The non-zero commutator makes these diagrams alone not gauge-invariant. Adding di-
agram (c) gives a contribution which exactly cancels this (try this!) but yields another
term proportional to kµ0 . This vanishes when we remember the whole expression is con-
tracted with ε0∗ν (k 0 ), and so gauge invariance is only obeyed once we project onto physical
polarisations. This wasn’t necessary in QED.
Recall in the QED case in section 3.2, we used Aµν k µ = 0 to show that, in practical
calculations, we can always make the replacement
2
X
εαµ ε∗α
ν → −gµν . (137)
α=1
Although the right-hand summed all polarisations and not only the physical transverse
ones, in actual calculations the unphysical longitudinal gluon polarisations automatically
cancelled. This is no longer the case in QCD, where one has to sum strictly over physical
polarisations. However, this can make calculations more cumbersome, so it might still be
useful to sum over all polarisations, and to cancel in some way the unphysical degrees of
freedom. How this cancellation is performed depends on the gauge. In covariant gauges,
like the Feynman gauge, this is done by introducing extra fields, called the ghost fields.
The alternative is to use the so-called physical gauges, that ensure that that only physical
degrees of freedom propagate on shell.
Ghost Fields
2 Im + =
Figure 13: Unitarity relation for the process e+ e− → γγ. The shaded blob represent the
sum of all possible subdiagrams that can give rise to two photons in the final state at the
lowest order in perturbation theory.
states that, at the considered order in perturbation theory, give a non-zero imaginary part,
namely two virtual photons. Furthermore, it is possible to show that the imaginary part of
any Feynman diagram is obtained by putting on shell in all possible ways the intermediate
propagators (i.e. cutting the diagram) by replacing i/(p2 − m2 + iε) in each of them by
(2π)Θ(p0 )δ(p2 − m2 ). This divides each Feynman diagram into two subdiagrams on either
side of the cut. On one side of the cut, one uses standard Feynman rules. On the other
sides, one needs to apply complex conjugation to all Feynman vertices and propagators.
Cuts of a diagram that conflict with energy-momentum conservation do not give any con-
tribution to the imaginary part. The result of this cutting procedure for the present case
is illustrated in Fig. 14. The dashed line on the right-hand side of the figure represents
the only cut of the diagram that gives a non-zero imaginary part, obtained by putting on
shell the intermediate photon propagators. If the amplitude is computed in the Feynman
2 Im =
Figure 14: Pictorial representation of the cutting rules needed to compute the imaginary
part of the forward amplitude e+ e− → e+ e− mediated by two virtual photons. The
shaded blob represents the sum all possible subdiagrams that can give two photons in the
intermediate state, that is, the two diagrams on the left-hand side of Fig. 13.
b c
q
q→ i
gs f abc q µ δ ab
µ a b q2 + iε
Figure 15: The Feynman rules for ghost fields, which are constructed explicitly to cancel
unphysical degrees of freedom.
we now consider the imaginary part of the forward q q̄ amplitude, at the lowest order in
perturbation theory we need to include not only gluons as intermediate states, but ghosts
as well, as pictorially illustrated in Fig. 16. The ghost-antighost loop contributes to the
imaginary part of the forward amplitude with a factor (−1), just like a normal fermion
loop, so as to cancel the contribution of the unphysical longitudinal gluon polarisations
when summing over all diagrams. The resulting imaginary part equals the amplitude
squared for the process q q̄ → gg, integrated over the gluon phase space and summed over
physical gluon polarisations, as required by unitarity of QCD.
Physical Gauges
Alternatively, we can impose a so-called “physical gauge” condition on the gluon fields
to eliminate unphysical polarisations from the start. This eliminates the need for ghosts,
which do not interact with gluons anymore, but complicates the gluon propagator. In
place of the Lorentz gauge condition ∂ µ Aaµ = 0, we impose
Aaµ nµ = 0 , (140)
2
+ =
Figure 16: Pictorial representation of the unitarity constraint for QCD discussed in the
text. Longitudinal polarisations for on-shell gluons in the cut amplitude on the left-hand
side of the equality are cancelled by the contribution of the ghost-antighost loop. Each
blob represents the sum of the three diagrams in Fig. 12.
for some arbitrary reference vector nµ . This is done by adding the gauge-fixing Lagrangian
1 a µ 2
Lgf = − (A n ) , (141)
2ξ µ
and taking the limit ξ → 0, thus enforcing the gauge condition in eq. (140).
The new expression for the propagator (for ξ = 0) is shown in fig. 17. When we use a
q→ q µ nν + q ν nµ qµqν
µ ν ab i µν
δ 2 −g + − n2
a b q + iε (qn) (qn)2
Figure 17: The gluon propagator when working in a physical gauge, Aaµ nµ = 0.
physical gauge, whenever we sum over polarisations, we can make the replacement
2
X q µ nν + q ν nµ qµqν
εαµ (q)ε∗α
ν (q) → −gµν + − n2 . (142)
α=1
(qn) (qn)2
The different choices of reference vector nµ correspond to different choices of the gauge.
One can explicitly check that results for physical quantities, such as cross sections, are
independent of this choice.
A relevant example of a physical gauge is the light-cone gauge, in which n2 = 0. In such a
gauge, if we have an on-shell gluon q = (ω, q), we can choose n = (1, −q/ω). In this case
2
q µ nν + q ν nµ X α
−gµν + = εµ (q)ε∗α
ν (q) , (143)
(qn) α=1
so that the replacement gives exactly the sum over the physical polarisations introduced
in section 5.1. The expression in eq. (143) is the one that must be used in covariant gauges
if one does not want to introduce unphysical amplitudes squared with ghosts in the final
state.
(a) (b)
Figure 18: Sample “loop” Feynman diagrams: (a) one of the one-loop corrections to
Coulomb scattering and (b) a one-loop correction to the photon propagator.
7 Renormalisation
with p the photon momentum and d the number of space-time dimensions. As the integral
runs over all values of k, it includes very large values of k. Counting the powers of k,
there are six of them in the numerator and four in the denominator, which implies that
this integral diverges. In general, for any integral of the form
dd k N (k)
Z
(145)
(2π)4 M (k)
we define the superficial degree of divergence, D, to be the result of the naı̈ve power-
counting:
If D ≥ 0, then the integral is said to be superficially divergent. Such divergences are called
ultra-violet (UV) because they arise whenever loop R momenta get large. The boundary case
of D = 0 is a logarithmic divergence (think of dk 1/k). The term “superficial” is used
because there can be other factors which can affect the actual degree of divergence. In
the example above, gauge invariance actually implies that the final result of the integral
in eq. (144) must be proportional to (p2 gµν − pµ pν ). Therefore the divergence is only
logarithmic, and not quadratic as it appears from naı̈ve power counting.
The main point, though, is not the degree of divergence, but the fact that one finds
divergences at all. These higher-loop corrections were supposed to be corrections in the
perturbative series, hence smaller than those appearing at the previous perturbative order.
For many years, this caused a major problem for the development of perturbation the-
ory. However, there exists a well-defined procedure to “remove” these divergences which
is called renormalisation. The basic idea behind renormalisation is that the parameters
appearing in the Lagrangian do not need to be physical quantities, but their value is deter-
mined by comparing perturbative predictions to actual experimental data. For instance,
the value of e can be extracted by measuring the Compton differential cross section at
small angles. Therefore, infinities that eventually appear in perturbative calculations can
be in principle reabsorbed in a redefinition of the parameters entering the Lagrangian. In
practice, this amounts to rescaling all quantities in the Lagrangian by a “renormalisation
constant”, Z. For instance, for a field φ we have
φ −→ φ0 = Zφ φR . (147)
The field φ0 is called “bare” field, as opposed to the “renormalised” field φR , and Zφ is
called renormalisation constant. This procedure has to be repeated for all fields, masses
and coupling constants. Provided that all infinities in the theory can be removed with
a finite number of renormalisation constants Z, then the theory is said to be renormalis-
able. After the renormalisation constants have been fixed, we can calculate all physical
quantities in terms of the renormalised quantities and the results will be both finite and
unambiguously defined.
The renormalisation constants are calculated according to some procedure that is called
“renormalisation scheme”. This consists in computing a suitable set of correlation func-
tions, and imposing that these functions are finite at any order in perturbation theory.
In this procedure one finds divergent integrals, which have to be regularised in some way.
The regularisation actually provide means to parameterise the divergence. One approach
is to implement a momentum cut-off, Λ, so as to artificially remove the region with large
momentum. The most common approach though is called “dimensional regularisation”.
Here we decrease the term d in eq. (146) to a lower value, so that we calculate all integrals
in d = 4 − 2 dimensions instead of d = 4. The integration measure becomes
d4 k d4−2 k
−→ , (148)
(2π)4 (2π)4−2
and for each dimensionsless coupling gR one performs the replacement
4−d
gR → µ 2 gR (µ) = µ gR (µ) . (149)
The factor of µ is essential to preserve the correct dimensions of the bare coupling in d
dimensions. The renormalised coupling gR stays dimensionsless and depends now on the
scale µ. The latter quantity is the famous renormalisation scale and it is the price that
we pay for renormalisation as our finite calculations are now all dependent upon µ.
To summarise, the steps to perform renormalisation within dimensional regularisation are:
and a similar expression holds for all couplings and masses. Both φ0 and Z are infinite for
→ 0, whereas φR (µ) stays finite, but depends on the unphysical renormalisation scale µ.
In a renormalised theory then, even tree-level diagrams depend on the renormalisation
scale, through the coupling for example. The dependence on the renormalisation scale
would dissappear only if we were able to calculate physical quantities to all orders in
perturbation theory. Although this is unpractical, calculating one or two extra orders in
perturbation theory can reduce the dependence considerably. However, this does mean
that any theoretical calculation now depends on a free parameter, and it is exactly this
parameter which leads to a way to estimate the “theory uncertainty”. In fact, consider an
observable O(αR (µ), µ, {Qi }), where {Qi } is a set of characteristic scales for the process.
If we know O(n) , the perturbative expansion of O at order n in perturbation theory, we
have
O(n) (αR (µ0 ), µ0 , {Qi }) = O(n) (αR (µ), µ, {Qi }) + O(αR
n+1
(µ)) , (151)
so that the variation of µ around some central value µ0 produces automatically a higher-
order term. Notice that O(n) (αR (µ), µ, {Qi }) might contain ln(Qi /µ). This is why the
central scale µ0 is normally chosen of the order of the typical value that the scales Qi can
assume. For example in gg → H, one would typically take µ0 ∼ mH .
The obvious way to gauge how the strength of the dependence on the scale in a calculation
is to vary the scale and see how the result varies. If the dependence is very weak, the
result will be negligible. If the dependence is very strong, the variation will be large.
The consensus of the community is to quote the theoretical uncertainty when the central
scale is varied by a factor of 2 in each direction. One should remember that this is only an
uncertainty of the dependence on the renormalisation scale and not a strict error bar. This
is illustrated by the plot in fig. 19, which is taken from Gehrmann-De Ridder, Gehrmann,
Glover & Pires, arXiv:1301.7310. It shows the scale dependence for inclusive jet production
in the gluon-gluon channel at LO, NLO and NNLO. Indeed the variation decreases each
time indicating that the sensitivity to the scale is decreasing. The fact that the lines do
not overlap is a clear sign that these uncertainty bands are not error bands.
dσ/dp (pb)
LO
s=8 TeV
T
NLO
80 anti-kT R=0.7 NNLO
MSTW2008nnlo
70 µ =µ =µ
R F
80 GeV < pT < 97 GeV
60
50
40
30
20
1 µ/p
T1
Figure 19: Plot showing the scale dependence for inclusive jet production at LO, NLO and
NNLO, taken from Gehrmann-De Ridder, Gehrmann, Glover & Pires, arXiv:1301.7310.
order in perturbation theory, and compare the obtained number with experimental data:
O(n) (αR (µ0 ), µ0 , {Qi }) = Oexp ⇒ αR (µ) . (152)
By doing this for various observables, characterised by different typical scales µ, one can
actually measure the dependence of the coupling on the renormalisation scale µ. This
dependence can be predicted theoretically, and the comparison of the predicted dependence
with the one that is actually observed represents one of the most stringent tests of the
validity of a given QFT. This is illustrated for QCD in fig. 20, where one sees an astonishing
agreement between the predicted “running” of the QCD coupling with the renormalisation
scale Q, and what is observed in experimental data.
The theoretical object that dictates how a coupling evolves with the renormalisation scale
is the beta function β(αR ), defined as
∂αR
µ2 2
= β(αR ) = −β0 αR 3
− β1 α R + .... (153)
∂µ2
There are various ways to compute the beta function, which in general depends on the
renormalisation scheme used. However, one can show that the first two coefficient of the
beta function, β0 and β1 , are independent of the renormalisation scheme. If we consider
a scheme tied to dimensional regularisation (e.g. the so-called MS scheme), one has the
relation
α0 () = µ2 Zg2 (, µ2 )αR (µ2 ) , (154)
where α0 = g02 /(4π) and αR = gR2 /(4π). The crucial observation is that the bare coupling
α0 does not depend on µ. Therefore, its logarithmic derivative with respect to µ2 is zero:
µ2 ∂Zg2
2 ∂α0 2 2 2 2 ∂αR
0=µ = µ Zg (, µ ) + 2 2 αR + µ . (155)
∂µ2 Zg ∂µ ∂µ2
Figure 20: The QCD coupling αs as a function of the renormalisation scale Q, in theory
and experiment, taken from arXiv:1512.0519.
This gives
µ2 ∂Zg2 µ2 ∂Zg2
∂αR
µ2 2 = − + 2 2 αR ≡ β(, αR ) → β(αR ) = − lim 2 2 αR , → 0.
∂µ Zg ∂µ →0 Zg ∂µ
(156)
In any scheme based on dimensional regularisation we have
αR (µ2 ) (1)
Zg (, µ2 ) = 1 + Zg + . . . . (157)
Therefore the first term of the beta function is just obtained from the 1/ pole of Zg , as
follows
(1)
µ2 ∂Zg2 2Zg ∂αR
β(αR ) = − lim 2 2
αR = − lim µ2 2 αR = −β0 αR
2
⇒ β0 = −2Zg(1) . (158)
→0 Zg ∂µ →0 ∂µ
| {z }
=−αR
(1)
The calculation of Zg can be performed using any quantity that involves an interaction
vertex. A way that is common to both QED and QCD is to consider the renormalised
interaction Lagrangian
p Z1
Lint → Zg Z2 Z3 (gR ψ̄R AR ψR ) = Z1 (gR ψ̄R AR ψR ) , ⇒ √ .
Zg = (159)
Z2 Z3
√ √
Here we have used the ubiquitous notation Zψ = Z2 and ZA = Z3 . The function
Z1 contains all UV divergences associated with loop corrections to the interaction vertex,
whereas Z2 and Z3 contain UV divergences arising in the calculations of the fermion
and gauge-boson propagators respectively. In QED, a powerful Ward identity implies
Z1 = Z2 , so that the beta function can be calculated just from all the loop corrections to
the propagator in the unrenormalised theory. For the case of QED
(1) 1
β0 = −2Zg(1) = Z3 = − . (160)
3π
Inserting this expression in the beta function we obtain
1 2
βQED (α) = α . (161)
3π
which means that the QED coupling, at least until the beta function is dominated by its
first term, becomes stronger with energy.
In QCD instead the Ward identity Z1 = Z2 does not hold any more. However, it holds
at least for the part of these renormalisation functions that depends on CF . Since, at
(1)
one loop, Z2 is proportional to CF , its contribution to the beta function cancels exactly
(1)
with the abelian contribution to Z1 . Therefore, the only contributions to the QCD beta
(1)
function at one loop come from the renormalisation function of the gluon Z3 and the
(1) (1)
non-abelian part of Z1 , which we call Z1 |n.a . The two depend on the gauge, but this
gauge dependence cancels in the combination
(1) (1)
β0 = −2Zg(1) = Z3 − 2Z1 |n.a . (162)
Acknowledgements
It has been a pleasure to give this course at the 2016 HEP Summer School. I would like
to thank Nikos Konstantinidis for directing this successful school, the other lecturers and
tutors for making this such an interesting and entertaining fortnight, and the students for
engaging with the courses.
Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.
Alternative Proxies: