You are on page 1of 200

Quantum Field Theory II

Prof. Luciano Vanzo1


Physics Department of Trento University
Via Sommarive, 14
I-38123 Povo (Trento)

Fall Semester 2016

1 luciano.vanzo@unitn.it
Abstract

The subject of the present course is the study of non abelian


gauge theories, BRST symmetry and cohomology, spontaneous
symmetry breaking phenomena and renormalization group meth-
ods. The path integral is introduced along the way as the
main tool to quantize gauge theories. The material in the
course should provide the fundamental tools to understand
more advanced topics, such as the study of the standard model
of particle physics and theories of grand unification. Some
books and papers are listed in the references. There are
countless of them, the papers being mostly referred to in
the cited textbooks. We only indicate a restricted sample
of them. Starred ones are recommended.
Contents

1 Classical gauge fields . . . . . . . . . . . . . . . . 4


1.1 Gauge invariance . . . . . . . . . . . . . . . . . 4
1.2 Gauge field dynamics . . . . . . . . . . . . . . . 9
1.3 Field equations and conservation laws . . . . . . 12
1.4 Fiber bundle description [incomplete] . . . . . . 14
1.4.1 Fiber bundles and the transition functions 14
1.4.2 Linear connections as gauge fields . . . . 15
1.5 Tree level processes . . . . . . . . . . . . . . . 15
2 The path integral method . . . . . . . . . . . . . . . 20
2.1 Solving functional equations . . . . . . . . . . . 20
2.1.1 Bosonic fields . . . . . . . . . . . . . . 20
2.1.2 Fermions . . . . . . . . . . . . . . . . . . 25
2.2 The Hamiltonian path integral . . . . . . . . . . 26
2.3 The Lagrangian path integral . . . . . . . . . . . 32
2.4 Path integral for constrained systems . . . . . . 34
2.4.1 Second class constraints . . . . . . . . . 37
2.4.2 First class constraints and the Faddeev
formula . . . . . . . . . . . . . . . . . . 40
2.5 Functional calculus . . . . . . . . . . . . . . . 44
2.5.1 Correlation functions and generating func-
tionals . . . . . . . . . . . . . . . . . . 45
2.5.2 Commutation relations . . . . . . . . . . . 50
2.5.3 Equations of motion and Ward identities . 51
2.6 Quadratic Lagrangians . . . . . . . . . . . . . . 52
2.7 The vacuum persistence amplitude . . . . . . . . 57
2.8 Transition to quantum field theory . . . . . . . 61
2.8.1 Scalar fields . . . . . . . . . . . . . . . 62
2.8.2 The massive vector field . . . . . . . . . 64
2.8.3 Spinor fields . . . . . . . . . . . . . . . 70
3 Quantum gauge fields . . . . . . . . . . . . . . . . . . 73
3.1 The Faddeev-Popov-De Witt path integral . . . . . 73

1
3.2 The quantum Lagrangian and the ghosts . . . . . . 80
3.3 BRST symmetry and cohomology . . . . . . . . . . . 84
3.4 Physical states and ghost decoupling . . . . . . 90
3.5 BRST Cohomology in the Fock representation . . . 93
3.6 Slavnov-Taylor identities . . . . . . . . . . . . 101
4 The effective action formalism . . . . . . . . . . . . 104
4.1 Preliminaries . . . . . . . . . . . . . . . . . . . 104
4.2 Proper vertices . . . . . . . . . . . . . . . . . . 106
4.3 The effective potential . . . . . . . . . . . . . 110
4.4 Symmetry and symmetry breaking . . . . . . . . . 113
4.5 The one-loop approximation . . . . . . . . . . . . 116
4.6 Convexity . . . . . . . . . . . . . . . . . . . . . 119
5 Spontaneous symmetry breaking . . . . . . . . . . . . . 120
5.1 Classical theory . . . . . . . . . . . . . . . . . 120
5.2 Decoupling of vacua . . . . . . . . . . . . . . . 127
5.3 The Goldstone theorem . . . . . . . . . . . . . . 130
5.4 The case of broken gauge symmetries . . . . . . . 133
5.5 The abelian Higgs mechanism . . . . . . . . . . . 136
5.6 The general Higgs mechanism . . . . . . . . . . . 140
5.7 The Standard Model Higgs . . . . . . . . . . . . . 144
6 The renormalization group . . . . . . . . . . . . . . . 150
6.1 Bare and renormalized functions . . . . . . . . . 150
6.2 The renormalization group equations . . . . . . . 153
6.3 Renormalization group trajectories . . . . . . . 154
6.4 Asymptotic behavior and fixed points . . . . . . 157
6.4.1 Infrared free theory . . . . . . . . . . . 158
6.4.2 UV free theory . . . . . . . . . . . . . . 160
6.5 General results of the RG analysis . . . . . . . 161
6.5.1 Prescription dependence . . . . . . . . . . 161
6.5.2 Physical Parameters . . . . . . . . . . . . 162
6.5.3 Dynamically generated masses . . . . . . . 163
6.5.4 Dimensional transmutation . . . . . . . . 165
6.5.5 The effective potential . . . . . . . . . . 166
6.6 Calculation of the beta function . . . . . . . . 168
6.7 Other renormalization schemes . . . . . . . . . . 170
6.7.1 The RG equation of Gell-mann and Low . . 170
6.7.2 The Georgi-Politzer RG equation . . . . . 171
6.7.3 The Callan-Symanzik RG equation . . . . . 172
6.8 The QED renormalization group equations [incom-
plete...] . . . . . . . . . . . . . . . . . . . . . 176

2
6.9 The case of QCD [incomplete...] . . . . . . . . . 177
6.9.1 Renormalization . . . . . . . . . . . . . . 177
6.9.2 Asymptotic freedom . . . . . . . . . . . . 179
6.9.3 The running coupling . . . . . . . . . . . 182
6.10 Appendix . . . . . . . . . . . . . . . . . . . . . 188
6.10.1 A - Operator insertions . . . . . . . . . . 188
6.10.2 B - The Berezin integral . . . . . . . . . 190
Bibliography . . . . . . . . . . . . . . . . . . . . . . 195

3
Chapter 1

Classical gauge fields

1.1 Gauge invariance


Gauge fields and gauge invariance are the leitmotiv of mod-
ern high energy physics. We shall therefore begin with a
formulation of these important concepts in purely classical
terms. The story is well known from classical electrodynam-
ics: in order that a matter Lagrangian, say L = Lm (ψn , ∂µ ψn ),
which is invariant under constant phase, or U (1), transfor-
mations

ψn (x) → eiα ψn (x), ∂µ ψn (x) → eiα ∂µ ψn (x) (1.1)

be made invariant also under space-time dependent phase


transformations, in which α = α(x), it is only necessary to
introduce a new field Aµ and a new derivative operator, the
”covariant derivative”,

∇µ ψn = ∂µ ψn − ieAµ ψn (1.2)

such that under the combined substitution rules


′ ′
ψn (x) → ψn (x) = eieα(x) ψn (x), Aµ → A (x) = Aµ + ∂µ α (1.3)

it transforms like ψn 1 . More explicitly,


′ ′ ′
∂µ ψn − ieAµ ψn = eieα(x) (∂µ ψn − ieAµ ψn ) (1.4)
1
The constant e is quite arbitrary at this stage but in the full
theory it will take the role of a coupling parameter.

4
These rules are dubbed local gauge transformations in the
physics literature, and the corresponding invariance is ac-
cordingly called local gauge invariance. Then the modified
Lagrangian density

L = Lm (ψn , ∇µ ψn ) (1.5)

will obviously be invariant under the local transformations


(1.3), if it was invariant under the global transformation
(1.1). In itself this has no dynamical content: we can
always render the theory locally gauge invariant without
changing its physical content provided one introduced a
gauge vector field, except that to explain why this symmetry
is not observed the field must have vanishing strength,
or Fµν = ∂µ Aν − ∂ν Aµ = 0, (otherwise we have a system in an
external field with a different physics, which is not bad
in any case).
Instead of considering Aµ as an external trivial field (a
pure gauge in the physics jargon) just to explain why the
local symmetry is unobserved, it is much more interesting
to consider fields with Fµν ̸= 0 and think of Aµ as a new
dynamical field with his own dynamics and Lagrangian, to
be added to Lm . The unique one with mass dimension four,
which is also Lorentz and gauge invariant, is the standard
electromagnetic Lagrangian

1 1
LF = − Fµν F µν = − Aµ (□ηµν − ∂µ ∂ν )Aν + boundary terms (1.6)
4 2

As is well known, the other possibility is κϵαβµν Fαβ Fµν , which


is a total derivative. If matter is described by a set of
spinor fields, one for each flavor f , then we can take

Lm = − ψ̄f (∇
/ + m)ψf (1.7)
f

where ∇/ = γ µ ∇µ (the Feynman’s slash), and the total La-


grangian will just be the sum L = Lm + LF . Note that the
gauge symmetry is based here upon the abelian group U (1).
Modern non abelian gauge theories are obtained by generaliz-
ing this procedure to general groups. For reasons to be soon

5
clarified it is convenient to restrict attention to compact
connected simple Lie groups, the most representative being
perhaps the unitary groups SU (N ) and the orthogonal groups
O(N ) . The matter fields will then furnish a representation
of the group and will be subject to transformation laws of
the kind (a sum over repeated indices is understood)

δψ = iϵa Ta ψ. (1.8)

The matrices Ta form a representation of the Lie algebra of


the group, and share with it the same commutation relations
(again with repeated indices summed over)

[Ta , Tb ] = iC cab Tc , C cab = −C cba (1.9)

The real structure constants C cab are constrained by the


Jacobi identity
[[Ta , Tb ], Tc ] + [[Tc , Ta ], Tb ] + [[Tb , Tc ], Ta ] = 0 (1.10)
and obey the algebraic equations

C cab C fcd + C cda C fcb + C cbd C fca = 0 (1.11)

Any solutions of this equation defines a possible Lie al-


gebra. It can be checked that the matrices
[TaAd ]bc = −iC bca (1.12)
define a special representation of the algebra (1.9) called
the adjoint representation.
Now for a compact simple or semi simple Lie group it is always
possible to choose the linearly independent representation
matrices Ta so that the structure constants are real and
completely anti-symmetric. In this case we simply write

Cabc = C abc

and avoid any distinction between upper and lower indices.


For example, we shall write the adjoint representation as
[TaAd ]bc = −iCbca = −iCabc . The matrices TaAd are then hermitian
and the corresponding group representation is unitary. To-
tal antisymmetry is a powerful constraint: for instance,

6
there is no compact simple Lie algebra in dimensions two,
since a three index anti-symmetric tensor cannot exist in
2D; and there is only one in dimension three, because any
anti-symmetric 3-index tensor must be proportional to the
Levi-Civita volume form ϵabc .
A necessary condition for the matrices Ta in any represen-
tation to be Hermitian (and thus the representation of the
group to be unitary) can be simply derived. The quadratic
form Q(u, v) = Tr(Ta Tb )ua v b is real and positive because

Q(u, u) = Tr(Ta Tb )ua ub = Tr[T (u)2 ], T (u) = Ta ua

and the eigenvalues of hermitian matrices like T (u) are real.


Neither it can vanish because then T (u) = 0 implies ua = 0,
since the Ta form a basis for the given representation of
the algebra. This then defines a positive symmetric metric
tensor on the vector space structure underlying the Lie
algebra

gab ∝ Tr(Ta Tb ) (1.13)

which is known as the Cartan metric. Furthermore, this


metric satisfies the so called invariance condition

gab C bcd = −C bca gdb (1.14)

which means that the Lie-algebra map u → [u, T ] is an isometry


of the metric.
Proof: the following chain of equalities is easily estab-
lished

gab C bcd = −i Tr[Ta Tb ]C bcd = −i Tr[Ta [Tc , Td ]] (1.15)


= −i Tr[[Ta , Tc ]Td ] = −C bca gbd

Conversely, the existence of a metric satisfying the in-


variance condition (1.14) implies that the algebra must be
a compact simple Lie algebra or a direct sum of such alge-
bras and U (1) subalgebras2 , for which the finite dimensional
representations are all unitary, and the matrices Ta all
Hermitean.
2
See S. Weinberg book “The Quantum Theory of Fields: Applications”,
Ch. 15.

7
Now we want to gauge the symmetry. Starting again with a
Lagrangian invariant under (1.8), and making the parameters
ϵa space-time dependent functions, we introduce a new set
of fields Aaµ (x) (one for each independent parameter) and a
covariant derivative

∇µ ψ = ∂µ ψ − iAaµ Ta ψ (1.16)

(which is well defined since the group acts on ψ) such


that under gauge transformations involving also the gauge
fields,

δ∇µ ψ = iϵa (x)Ta ∇µ ψ (1.17)

This uniquely fixes the transformation rule of the gauge


field itself; with a simple computation we get

δAaµ = ∂µ ϵa + C abc Abµ ϵc ≡ ∇µ ϵa (1.18)

In terms of the matrix-valued one-form A = Aaµ Ta dxµ and ϵ = ϵa Ta


we have the elegant formula

δA = dϵ + i[ϵ, A] (1.19)

This is recognizable as the infinitesimal version of the


finite transformation rule

Aµ = −i∂µ U U −1 + U Aµ U −1 , U = exp(iϵa (x)Ta ) (1.20)

The Lagrangian obtained by the substitution ∂µ ψ → ∇µ ψ

L = Lm (ψ, ∇µ ψ) (1.21)

is then invariant under the combined local gauge transfor-


mations
′ ′
ψ (x) = U (x)ψ(x), A = −idU U −1 + U AU −1 (1.22)

provided it was invariant under the global transformations


with constant parameters and no gauge fields.

8
1.2 Gauge field dynamics
As for electromagnetism, any theory with a rigid symmetry
can be gauged trivially by introducing such a vector field
without changing the content of the theory provided the
gauge field has zero strength. But much more interesting
would be not requiring such restriction and consider instead
the gauge fields as dynamical fields, governed by a suitable
gauge invariant Lagrangian.
But first of all, what is now the correct definition
of the field strength, generalizing electrodynamics? The
easiest way to find it is to consider the commutator of
two covariant derivatives; an easy computation leads to an
equation of the form
[∇µ , ∇ν ]ψ = −iFµν
a
Ta ψ (1.23)
where
a
Fµν = ∂µ Aaν − ∂ν Aaµ + C abc Abµ Acν (1.24)
It strongly resembles the electromagnetic field tensor were
it not for the second, non linear term, peculiar to non
abelian groups. It is a field strength characterizing the
presence of a true gauge field, because it turns out that
a
Fµν = 0 if and only if Aaµ = ∇µ ha for some set of functions ha ,
namely, if the potential is “a pure gauge”. This resembles
the flatness conditions of Riemannian manifolds or general
relativity (GR), according to which the metric is flat (that
is there is no true gravitational field) if and only if the
curvature tensor Rµνσρ = 0. Then the metric can be gauged
away to approach the Euclidean or Minkowski metric, as the
case may be.
Using the Eq.(1.17) in (1.23) we can see that F transforms
according to the adjoint representation of the group, or
a
δFµν = iϵb [TbA ]ac Fµν
c
= ϵb C ab c Fµν
c
=⇒ δF = iϵa [F, Ta ] (1.25)
a
in terms of matrices Fµν = Fµν Ta . The finite transformation
′ −1
is Fµν = U Fµν U , showing by the way that Tr Fµν F µν is gauge
invariant. This bring us to define the Lagrangian
1 1 1
L = − Tr Fµν F µν = − gab Fµν
a
F bµν , gab =: Tr Ta Tb (1.26)
2 4 2
9
It is here that the Cartan metric plays a special role. From
(1.25) we see that the gauge invariant of L is equivalent
to the invariance condition (1.14) (which is the reason for
the name)

gab C bcd = −C bca gdb (1.27)

This condition can be written in terms of matrices in the


adjoint representation as

[g, T A ] = 0 (1.28)

Now for a simple Lie algebra (one that has no proper invari-
ant subalgebras) the matrices T A form an irreducible set3 so
that gab must be proportional to the identity: we write it
in the form

gab = g −2 δab (1.29)

for some real number g, and also adopt the fixed normal-
ization of the gauge generators: Tr[Ta Tb ] = 12 δab . Then by a
rescaling Aaµ → gAaµ we can put the gauge field Lagrangian in
the form
1 a aµν 1
L = − Fµν F = − Tr Fµν F µν (1.30)
4 2
The g factor then appears in the covariant derivatives and
the commutation relations as wells. In particular, the
field strength and the covariant derivatives will take the
form
a
Fµν = ∂µ Aaν − ∂ν Aaµ + gC abc Abµ Acν (1.31)

∇µ ψ = ∂µ ψ − igAaµ Ta ψ (1.32)

δAaµ = ∇µ ϵa = ∂µ ϵa + gAbµ Cbc


a c
ϵ (1.33)

and the gauge transformations too will have such g factors,


e.g. δψ = igϵa Ta ψ, and so on. We see that g plays the role of
a universal coupling constant, in the sense that the same g
governs the strength of the interactions of all matter and
3
If there were matrices other than the identity which commute with all
TaA these would span an invariant sub-algebra, by the Jacobi identity.

10
gauge fields simultaneously. In general there will be one
factor g for any simple component of the total gauge algebra
which, as we know, must be the direct sum of compact simple
and U (1) subalgebras.
The gauge Lagrangian (1.30) has the structure of a free
quadratic part, just like N copies of an abelian gauge
field, plus trilinear and quartic couplings
L = L2 + L3 + L4 (1.34)
where L2 is the free (quadratic) part and
g
L3 = − Cabc (∂µ Aaν − ∂ν Aaµ )Abµ Acν (1.35)
2

g2
L4 = − Cabc Cade Abµ Acν Adµ Aeν (1.36)
4
For a fermion multiplet in a representation T r of the gauge
group with Lagrangian
Lf = −ψ̄(∂/ + m − igA/a Tar )ψ (1.37)
we can read off the interaction
Lf 3 = ig ψ̄γ µ Tar ψ Aaµ = Jaµ Aaµ (1.38)
Similarly, for a scalar multiplet ϕ = (ϕ1 , . . . , ϕn ) in some rep-
resentation of the group, we can write the gauge invariant
standard Lagrangian
Lb = −(∇µ ϕ)∗ ∇µ ϕ − m2 ϕ∗ ϕ − V (ϕ∗ ϕ), ∇µ ϕ = ∂µ − igAaµ Ta ϕ (1.39)
from which we can read off the interactions
←→ ←→
Lb3 = −igϕ∗ Ta ∂ µ ϕAaµ , ϕ∗ ∂ µ ϕ = ϕ∗ ∂ µ ϕ − ∂ µ ϕ∗ ϕ (1.40)

Lb4 = −g 2 ϕ∗ Ta Tb ϕ Aaµ Abµ (1.41)


Digression: we may observe here that if the scalar multiplet
has a non vanishing vacuum expectation value, say vn = ⟨0|ϕn |0⟩,
then expanding ϕ around v the interactions apparently give
a mass term to the gauge fields, with masses of order g|v|,
i.e.
1
Lb4 = − Mab Aaµ Abµ + · · · , Mab = −2g 2 vTa Tb v (1.42)
2
11
But of course this vanishes if the vacuum is left invariant
by the symmetry group, i.e. if Ta v = 0 for any a. Thus the
mass term requires one or more broken symmetry generators,
for which in fact Ta v ̸= 0. This then is another vacuum4 , so
that there must be vacuum degeneracy in order to have such a
mass generation mechanism, and we then say that the symmetry
is spontaneously broken. One may wonder what happens to the
original gauge symmetry that required no mass terms in the
Lagrangian to begin with.
Similarly, if there are Yukawa couplings, as for example

LY = gY ϕj ψ̄j ψj + h.c. (1.43)
j

these will generate masses to the fermion fields too, of


order gY |v|. These simple facts will be the key to interpret
the short range weak interactions of elementary particles as
gauge interactions, for which massive gauge fields would be
needed. It was the natural desire of maintaining the same
good ultraviolet behavior of gauge fields like electrodynam-
ics that prompted physicists to investigate new mechanisms
of mass generation like this one, rather than introducing
the masses by hand.

1.3 Field equations and conservation laws


a
The definition of Fµν as the commutator of two covariant
derivatives, Eq. (1.23), has as a very important implication
the Bianchi identities5
∇µ Fσρ
a
+ ∇ρ Fµσ
a
+ ∇σ Fρµ
a
=0 (1.44)
which in electrodynamics correspond to Faraday’s law and
the absence of magnetic charges. It suffices to note that
commutators obey the Jacobi identity, to deduce (1.44).
The Bianchi’s identities are the Yang-Mills analogue of the
”true” Bianchi identities first proved by Bianchi for the
Riemann curvature tensor

∇µ Rναβ
ρ
+ ∇α Rνβµ
ρ
+ ∇β Rνµα
ρ
=0
4
The gauge symmetry commutes with space-time translations, hence with
the Hamiltonian.
5
In electrodynamics these form the first pair of Maxwell’s equations.

12
still another analogy between gauge theories and general
relativity. The Bianchi identities hold for any gauge field
and do not involve sources. In fact they describe a property
of a general mathematical structure which is common to all
gauge theories, that of a linear connection on a principal
fiber bundle.
The equations linking F with the sources, or in the physics
speech the dynamics of the gauge fields, is derived from
the action principle given above. Let us define the gauge
current, which plays the role of source for the gauge field,
as
δIm
Jaµ = (1.45)
δAaµ
where Im is the matter field (i.e. non gauge, like quarks and
lepton fields) Lagrangian. From the Lagrangian we readily
find the following field equations
∇µ F aµν ≡ ∂µ F aµν + gCabc Abµ F cµν = −J aν (1.46)
We see that the total current
jaµ = Jaµ + gCbc
a bµ cµν
A F (1.47)
is exactly conserved, ∂µ jaµ = 0, but gauge dependent. Since a
non abelian gauge field contributes to the currents it is
itself charged, and it couples to everything carries a gauge
charge or, more formally, to everything carries a gauge in-
dex a, b, c, . . . . Taking the covariant derivative of Eq. (1.46)
it is seen that the current Jaµ , though gauge covariant,
obeys instead the covariant balance equation ∇µ Jaµ = 0.
Remember gravitation: since it couples to everything car-
ries energy and momentum it must couple to itself. This
is the reason why the matter stress tensor is not exactly
conserved, the balance equation being in this case,

1 √
√ ∂µ −gT µν + Γµαβ T αβ = 0
−g

whereas the total stress tensor obtained by adding to Tµν


the stress tensor of gravity, say τµν , satisfies the exact
conservation law

∂µ (T µν + τ µν ) = 0

13
The gravitational stress tensor is a functional of the metric
and its first order partial derivatives, so just like the
gauge current it is gauge dependent6 , that is dependent on
the coordinate frame. At least from this point of view
non abelian gauge theories display then some deep analogy
with general relativity, as a result of which it too can be
considered as a gauge theory of the gravitational field.

1.4 Fiber bundle description [incomplete]


The proper mathematical framework fitting the properties
of gauge fields, in particular the somewhat unusual, non
homogeneous transformation law, is the theory of linear
connections on fiber bundles. We shall describe briefly
what this framework is and point out some interesting fact.

1.4.1 Fiber bundles and the transition functions


A principal fiber bundle (p.f.b.) with structural group G
(supposed to be a Lie group), base manifold M and projection
π : P → M , denoted by P (M, G, π), is a differentiable manifold
over which the group acts freely from the right in such a
way that
(a) for g ∈ G and u ∈ P , the right action u → ug is free (i.e.
without fixed points) and differentiable
(b) M = P /G is the quotient space of P under this action
and π(u) = π(ug). For x ∈ M , the set π −1 (x) is a closed
submanifold of P : it is called the fiber of P over x.
(c) P is locally trivial, that is there is an open covering
{Uα }α∈A of M such that each π −1 (Uα ) is diffeomorphic
to Uα × G, the diffeomorphism being of the form ψα (u) =
(π(u), ϕα (u)), where ϕα : π −1 (Uα ) → G satisfies ϕα (ug) = ϕα (u)g.
For u ∈ π −1 (Uα ∩ Uβ ), the functions ψαβ (u) = ϕα (u)ϕβ (u)−1 only
depend on u through π(u), namely ψαβ (ug) = ψαβ (u). Hence they
really define functions

ψαβ : Uα ∩ Uβ → G
6
It means it is not a true space-time tensor.

14
These are known as the transition functions of the bundle
P , and satisfy the characteristic relations

ψαβ · ψβγ = ψαγ (1.48)

on Uα ∩ Uβ ∩ Uγ , whose proof is trivial. The remarkable


fact is that the converse is also true, namely that given
a covering {Uα }α∈A of a differentiable manifold7 M and a
set of functions ψαβ : Uα ∩ Uβ → G such that (1.48) is true,
there is a unique p.f.b. with structural group G whose
transition functions are the ψαβ . The theorem provides also
a constructive procedure for principal fiber bundles. As
we will see, in gauge theories the transition functions
provide the link between different gauge choices.

1.4.2 Linear connections as gauge fields


To proceed we need some basic facts about the action of Lie
groups and algebras. Given a curve in the group t → g(t)
starting at the identity, g(0) = e, the right action of G
on P produces a curve t → u · g(t) in P with initial point u;
the velocity of this curve at u is a vector at u tangent
to the fiber over π(u), so by varying u we get an induced
vector field which is called a vertical vector field. This
can be identified naturally with a right invariant vector
field on the group, or equivalently with an element of the
Lie algebra of the group.

1.5 Tree level processes


It is interesting at this point to investigate some sim-
ple consequences of the gauge theories outlined above in
the quantum domain. We shall understood with this term
a naive reading of the Feynman rules from the Lagrangian
after splitting it into a free and interaction parts, and
investigate whether they give sensible, gauge invariant,
results. An informative example is fermion anti-fermion
scattering into gauge bosons. Propagators and vertices can
7
A manifold of paracompact type, to be precise.

15
be read from the free Lagrangian directly and are, respec-

F
tively, for fermions and the gauge bosons in Feynman’s gauge
∂ µ Aaµ = 0,
a p b −iδ ab
=

g
ip/ + m − iϵ
p iδ ab ηµν
=−
p2 − iϵ

dg
a,µ b,ν

e
i

= −gγ µ [T a ]ij
µ,a

In the last diagram we are suppressing Dirac fermion indices,


hidden as they are into γ µ . We see that free propagation
cannot change the color charge.8 Also, a fermion-anti-
fermion pair can make up a bi-colored gluon, which in fact
belong to the adjoint representation of the group. The
gauge part of the Lagrangian also gives cubic and quartic
interactions, to which there correspond a 3-point vertex

VUG
µ,a
p
r
q = gf abc (ηνρ (q − r)µ + ηρµ (r − p)ν + ηµν (p − q)ρ )
ρ,c

ν,b

and a 4-point vertex (with incoming momenta not indicated )


b,ν c,ρ

a,µ d,σ
[
= −ig 2 f abe f cde (ηµσ ηνρ − ηµρ ηνσ ) + f ade f cbe (ηµν ησρ − ηµρ ησν )
]
+ f ace f bde (ηµσ ηνρ − ηµν ησρ ) (1.49)

One way to obtain these vertices is, for example, by applying


the cubic or quartic operators in the Lagrangian to a three-
particles or four-particles gluon state, and see the result
8
We are using the QCD language as the paradigmatic example of a
realistic non abelian gauge theory, without any implication to real
QCD.

16
after having been stripped away the polarization vectors of
the gluons. Another would be to take the third order or
fourth order functional derivative of the classical action
with respect to the gauge fields, and Fourier transform
the result9 . For still other ways look at textbooks, for
example11 or.12
Now the diagrams for the tree level f f¯ → gluons are

f¯ g f¯ g f¯ g

+ + (1.50)
f g f g f g

The direct and crossed diagram are just the same as in QED
and the gluon exchange diagram is typical of non abelian
gauge theories and absent in electrodynamics. We can apply
the Feynman rules outlined above to find the total amplitude,
the sum of the three diagrams, which must have the form3

M = M µν ϵ∗µ (k1 )ϵ∗ν (k2 ) (1.51)

in terms of the gluon polarization vectors and final momenta.


Gluons are massless particles with helicity s = 1, so the
two physical polarization states should be transverse, or
ϵµ (k)k µ = 0; together with the longitudinal polarization, say
ϵLµ ∝ kµ = (|⃗k|, ⃗k), and the backward polarization, ϵBµ ∝ k̃µ =
(|⃗k|, −⃗k), with normalization fixed by ϵ · ϵ = −1, they form a
B L

complete, not orthonormal, basis



ϵi(µ ϵ∗i
ν) − ϵµ ϵν − ϵµ ϵν = ηµν
B L L B
(1.52)
i=1,2

As in QED, the longitudinal and backward polarizations


should correspond to zero norm states with zero amplitudes
as dictated by the gauge symmetry. Hence the amplitude
9
This is the tree level procedure of taking the same functional
derivatives of the effective action in the external fields, which gives
the exact 1P I vertices.

17
for emission of at least one longitudinal gluon should also
vanishes

M µν ϵ∗µ (k1 )k2ν = 0

We would like to check this. Let (p1 , p2 ) be the 4-momenta


of the incoming fermions and (k1 , k2 ) those of the outgoing
gluons. Using the Feynman rules outlined above we obtain
easily the amplitude10 for the sum of the first two diagrams

M1+2 = −g 2 v̄(p2 )[γ µ T a (p/1 − k/2 )−1 γ ν T b + (1.53)


+ γ ν T b (p/1 − k/1 )−1 γ µ T a ]u(p1 )ϵ∗µ (k1 )ϵ∗ν (k2 )

Now set ϵ∗ν (k2 ) = (k2 )ν and use Dirac’s equations for massless
fermions to write k/2 u(p1 ) = (k/2 − /p1 )u(p1 ) and v̄(p2 )k/2 = v̄(p2 )[k/2 − /p2 ].
The propagators cancel leaving the result

M1+2 = g 2 v̄(p2 )γ µ [T a , T b ]u(p1 )ϵ∗µ (k1 ) (1.54)

Now consider the third diagram with the gluon vertex. It


gives the cumbersome expression
1
M3 = −ig 2 v̄(p2 )γρ T c u(p1 ) f abc [η µν (k2 − k1 )ρ + (1.55)
(k1 + k2 )2
+ η ρν (k̃ − k2 )µ + η µρ (k1 − k̃)ν ]ϵ∗µ (k1 )ϵ∗ν (k2 )

where k̃ = −(k1 + k2 ) = −(p1 + p2 ). Replace as before ϵ∗ν (k2 ) = (k2 )ν ;


after some algebra, assuming the other gluon to be on shell
and physical ϵ∗ν (k1 )k1ν = 0, this gives

M3 = −ig 2 v̄(p2 )γ µ f abc T c u(p1 )g 2 ϵ∗µ (k1 ) (1.56)

Since [T a , T b ] = if abc T c , we see that the total amplitude with one


longitudinal gluon vanishes:M1+2 + M3 = 0, in agreement with
expectations. Note that if the coupling is not universal
(the same g to all vertices) no cancellation occurs.
What’s about if both gluons are unphysical? It is con-
venient to describe them by means of the forward/backward
polarization ϵL,B introduced above. The amplitude to produce
10
The reference here is Peskin & Schroder’s Quantum field theory book,
Ch. [15].

18
such a pair of forward/backward polarized gluons turns out
to be non vanishing
1 ω1
MLB = g 2 v̄(p2 )k/1 T c u(p1 ) f abc , ω1,2 = |⃗k1,2 | (1.57)
2k1 · k2 ω2
This fact has some unpleasant consequences. One is that even
if the states of a single forward or backward polarized boson
have zero norm in the Hilbert space, the state describing
both has positive norm and there seems to be no way to
eliminate it from the spectrum. On the other hand, consider
for example the forward scattering of fermions via gluons
exchange. According to the optical theorem11 , the 1-loop
diagram below has an imaginary part which is proportional
to the scattering cross section for all reactions f f¯ →
gluons.

With the covariant propagator we are using, η µν (k 2 − iϵ)−1 ,


all polarization states circulate within the loop (look at
Eq. (1.52)), so they must also be present in the final two-
gluon states of the annihilation process. Thus we cannot
simply erase the unwanted states from the spectrum without
running into troubles with loop diagrams. Should we give
up the optical theorem? In this case we loose unitarity
since the optical theorem is a necessary consequence of it.
Or is perhaps the propagator we used the wrong one? This
seems even more strange, given that it is the standard form
used in countless computations with massless particles. To
clarify these matters we now reformulate the quantum theory
with the more powerful methods of functional integration.
Along the way we will encounter an unexpected new symmetry
for gauge theories, which is powerful enough to eliminate
the unphysical degrees of freedom, and even to provide an
alternative road to quantization.

11
See Weinberg books, Vol.[1], §(3.6).

19
Chapter 2

The path integral method

Canonical methods to quantize gauge theories are possible


but very inconvenient. As we described in the previous sec-
tion, even the use of a simple gauge fixing like the Lorenz
gauge leads to some troubles with unitarity. Additional
negative norm states are to be postulated in order to can-
cel the unwanted polarization states, as guessed by Feynman.
As showed by L. Faddeev, A. Popov and B. De Witt, the al-
ternative but formal method of path integral quantization
has something more to offer.

2.1 Solving functional equations


One can view functional integration as the inverse opera-
tion of taking the functional derivative of a functional,
and consequently as a tool to solve linear functional dif-
ferential equations. Let us see how this works for the
simplest quantum fields, the scalar and the fermion fields.

2.1.1 Bosonic fields


Let us remind the Symanzik’s computation of the generating
functional of a scalar field theory, say for a real scalar
field, for which

Z[J] = ⟨0| T exp(i Jϕd4 x) |0⟩ (2.1)

20
We have evidently

δZ
−i = ⟨0| T ϕ(x) exp(i Jϕd4 x |0⟩ (2.2)
δJ(x)
Now apply the KG operator
( ) ∫
δZ
(□ − m ) −i
2
= (□ − m ) ⟨0| T ϕ(x) exp(i Jϕd4 x |0⟩
2
(2.3)
δJ(x)
Time derivatives do not commute with the time ordering op-
eration due to the presence of θ functions. For example,
using the canonical commutation relations we easily obtain

∂02 ⟨0| T ϕ(x)ϕ(y) |0⟩ = ⟨0| T ∂02 ϕ(x)ϕ(y) |0⟩ − iδ(x − y)

This has the effect that



(□ − m ) ⟨0| T ϕ(x) exp(i
2
Jϕd4 x |0⟩ =

= ⟨0| T (□ − m )ϕ(x) exp(i Jϕd4 x |0⟩ − J(x)Z[J]
2
(2.4)

as a result of which we obtain the so called Schwinger-Dyson


equation for this case
( ) ∫
δZ
(□ − m ) −i
2
= ⟨0| T (□ − m )ϕ(x) exp(i Jϕd4 x |0⟩
2
δJ(x)
−J(x)Z[J] (2.5)

This is a functional differential equations. To solve it we


must take as the inverse of the KG operator the Feynman’s
propagator

∆F (x − y) = −[□ − m2 + iϵ]−1 (x, y)

because
δ 2 Z[J]
∆F (x − y) = −i (2.6)
δJ(x)δJ(y) |J=0

Moreover we will have a field equation for the Heisenberg


field

(□ − m2 )ϕ = V (ϕ) (2.7)

21
For a free field the solution is trivially found: the
equation reduces to
( )
δZ
(□ − m ) i
2
= J(x)Z[J] (2.8)
δJ(x)
with the well known solution
( ∫ )
i
Z[J] = exp J(x)∆F (x − y)J(y)d xd y
4 4
(2.9)
2
In the interacting case there is not an exact solution like
this and one has to resort to some approximation scheme.
An interesting approach is to solve it by functional Fourier
transform. Let us write
∫ ∫ 4
Z[J] = E[φ]ei Jφd x Dφ (2.10)

with some measure Dφ on the space of fields. The differ-


ential equation for Z[J] becomes a simpler equation for the
functional E[φ]
′ δE
(□ − m2 + iϵ)φ E[φ] = V (φ)E[φ] − i (2.11)
δφ
Solving it we get the remarkable result
[ ∫ ]
i
E[φ] = E0 exp (φ(□ − m + iϵ)φ − V (φ))d x
2 4
(2.12)
2
where E0 is a constant of integration normally chosen so
that Z[0] = 1. So E[φ] is the exponential of the classical
action and we have the functional integral representation
∫ [ ∫ ]
i
Z[J] = exp (φ(□ − m + iϵ)φ − V (φ) + Jφ)d x Dφ
2 4
(2.13)
2
Note the presence of the Feynman’s iϵ-prescription. But does
this makes sense? A formal definition of the functional
measure could be

Dφ = dϕ(x) (2.14)
x

which suggests the replacement of the continuum with a lat-


tice having finitely many points and then taking the limit
as the spacing of the lattice tends to zero. This requires

22
first of all putting the system in a finite box so that
one has actually to perform two limits, the infinite volume
limit and the short distance limit. According to a modern
Wilsonian point of view only the infinite volume limit would
be considered, since at very short distances one does not
really know what are the working degrees of freedom. If
gravity is present even the large volume limit is problem-
atic, because very large spatial distributions of energy
could collapse to form black holes. But from a strictly
formal point of view, lattice regularization could really
provide a good definition of the functional measure.
Another solution of the functional measure problem is
rotating the theory to Euclidean signature, making the
oscillatory integral (2.13) exponentially damped. Actu-
ally, in the theory of random fields it is known after
Borcher’s theorem that if the two-point correlation func-
tion ⟨ϕ(x)ϕ(y)⟩
∫ is a positive distribution, then the functional
Z[J] = ⟨exp(i ϕJ)dx⟩ is the Fourier transform of a positive
measure
∫ ∫ ∫
⟨exp(i ϕJdx⟩ = dµ(φ) ei φJdx

So in this case there really exist a rigorous definition of


the functional integral. One may even take a more formal
point of view: Eq.s (2.1) and (2.13) without potential
energy give

⟨0| T ϕ(x1 , . . . , ϕ(xn ) |0⟩ = Dφφ(x1 · · · φ(xn )eiI0 [φ] (2.15)

where I0 is the free action. One may use, now, the left
hand side to define the right hand side. After this one may
calculate the interacting path integral (2.13) by expanding
the exponential in powers of V (φ) and then using (2.15)
repeatedly.
complex scalars and vector fields
Complex fields are in no way more complicated than the
case studied. One has sources for both ϕ and ϕ̄, and a
generating functional for chronological Green functions

Z[J, J] = ⟨0|T exp(i (Jϕ
¯ ¯ + J ϕ̄)d4 x)|0⟩ (2.16)

23
Proceeding as before we get the functional integral repre-
sentation
∫ [ ∫ ]
Z[J] = exp i (φ̄(□ − m + iϵ)φ − V (φ, φ̄) + Jφ + J φ̄)d x DφDφ̄
2 ¯ 4

More difficult is the case of a massive vector field whose


quantization does not immediately follows from the quan-
tization of the scalar field. The Lagrangian (A. Proca,
1936)
1 1
L = − Fµν F µν − µ2 ϕµ ϕµ , Fµν = ∂µ ϕν − ∂ν ϕµ (2.17)
4 2
provides a consistent description of massive spin-1 rela-
tivistic particles, since it has positive energy and the
field equations do not propagate the spin-0 part of the
field, because
∂µ ϕµ = 0 (2.18)
is a consequence of the equations of motion. Due to this
condition the field equations take the form
(□ − µ2 )ϕµ = 0, ∂µ ϕµ = 0 (2.19)
The equation for ϕ0 can also be written as
(∇2 − µ2 )ϕ0 + ∂i ϕ̇i = 0 (2.20)
This is one case in which the solution of this equation, if
plugged back in the Lagrangian, gives the correct Lagrangian
for the remaining variables ϕi = (ϕ1 , ϕ2 , ϕ3 ); this procedure
gives the following reduced Lagrangian for ϕi
1 [ ] 1 1
L = ϕ̇i δij − ∂i (−∇2 + µ2 )−1 ∂j ϕ̇j − Fik Fik − µ2 ϕi ϕi (2.21)
2 4 2
We may note that L is non local and also not (manifestly,
at least) Lorentz invariant. Its quantization looks very
awkward, to say the least. It follows that the apparently
covariant generating functional
[∫ ]
Z[J] = ⟨0|T exp i J ϕµ d x |0⟩
µ 4
(2.22)

has to be treated with some care, since the 0-component ϕ0 is


a non local functional of the remaining spatial variables ϕi .

24
The external current is taken here to be conserved, ∂µ J µ = 0,
in order to decouple from Z[J] the spin-0 component of the
vector field1 . We shall therefore postpone a full treatment
of the vector field after we have understood better the
functional integral, in Sec. (2.8.2).

2.1.2 Fermions
Fermi fields can be treated along similar lines. We define2
[∫ ]
Z[η, η̄] = ⟨0| T exp i (η̄Ψ + Ψη)d x |0⟩
4
(2.23)

where the arguments are anti-commuting sources. For such


functionals we have to distinguish among left and right
derivatives: writing

δZ = (δ η̄(x)A + Bδη(x))dx

then A is defined to be the left functional derivative of


Z, and B the right derivative, written respectively

δL Z δR Z
A= , B=
δ η̄(x) δη(x)

If Z is bosonic then clearly A and B are fermionic and


δ L = −δ R . Using the canonical anti-commutation relations of
fermion fields,
0
[Ψr (t, ⃗x), Ψs (t, ⃗y )]+ = iγrs δ(⃗x − ⃗y ), [Ψr (t, ⃗x), Ψs (t, ⃗y )]+ = 0 (2.24)
the following Schwinger-Dyson equations are obtained
( ) [∫ ]
δL Z
(∂/ + m) −i = ⟨0| T (∂/ + m)ψ(x) exp i (η̄ψ + ψ̄η)d x |0⟩
4
δ η̄(x)
+η(x)Z[η, η̄] (2.25)
( ) ← ←
[∫ ]
δR Z
−i ( ∂/ −m) = ⟨0| T ψ̄(x)( ∂/ −m) exp i (η̄ψ + ψ̄η)d x |0⟩
4
δη(x)
−η̄(x)Z[η, η̄] (2.26)
1
The coupling Jϕ vanishes for ϕµ = ∂µ χ.
2
The Ψ(x)’s denote field operators, small case ψ(x) are c-number anti-
commuting functions.

25
For a free Dirac field they are integrated exactly, giving
the result
[ ∫ ]
Z[η, η̄] = exp i η̄(x)∆f (x − y)η(y)dxdy (2.27)

where
∆f (x − y) = (∂/ + m − iϵ)−1 (x, y) = (∂/ − m + iϵ)−1 (x, y) (2.28)
is the fermion Feynman’s propagator. With several qualifi-
cations we can still try a functional Fourier transform
∫ [∫ ]
4
Z[η, η̄] = DψDψ̄ E[ψ, ψ̄] exp i (η̄ψ + ψ̄η)d x (2.29)

Using the Schwinger-Dyson equations (one needs only one of


theme) we get again the exponential of the classical action
[ ∫ ]
E[ψ, ψ̄] = E0 exp −i ψ̄(∂/ + m − iϵ)ψdx (2.30)

But in order to make sense out of Eq. (2.29) the inte-


gration variables must be anti-commuting objects3 , so that
the integral appears to be nothing more than a very formal
mathematical symbol. It can actually be given a meaning by
means of Berezin’s integration theory (see Appendix [A]).

2.2 The Hamiltonian path integral


The path integral formula for a system with N degrees of
freedom is based on a limiting form of the group composition
property of the evolution kernel4 in quantum mechanics,

according to which for any n and t < t1 < · · · < tn < t

′ ′ ′ ′
⟨q , t |q, t⟩ = dq1 . . . dqn ⟨q , t |qn , tn ⟩ · · · ⟨q1 , t1 |q, t⟩ (2.31)

Here5 for each j ∈ {1, . . . , n}, qj = (qj1 , . . . , qjN ) is a set of (in


principle) generalized coordinates for N degrees of free-
dom, and (we are using units in which ℏ = 1, which is not
3
Because the sources are anti-commuting to begin with.
4
This is the quantum mechanical analogue of the Markov property of
stochastic processes in probability theory.
5
It is understood that each measure dqk really stands for the N -
dimensional measure dN qk over configuration space, with the same meaning
for momentum integrations later on.

26
good when treating the classical limit) and
|q, t⟩ = eiHt |q⟩ (2.32)
is the eigenstate of the Heisenberg position operator Q(t) =
eiHt Qe−iHt whose eigenvalue is q, namely Q(t) |q, t⟩ = q |q, t⟩, so
that
′ ′ ′ ′
⟨q , t |q, t⟩ = ⟨q |e−iH(t −t) |q⟩ (2.33)

is indeed the transition probability amplitude from q to q .
Now for very large n the time intervals ϵk = tk − tk−1 are very
small, so that we can approximatively write6

dp
⟨qk , tk |qk−1 , tk−1 ⟩ = exp i[(qk − qk−1 )p − ϵH̃(qk , p)] (2.34)
(2π)N
with increasing precision as ϵ → 0. We defined the (q, p)-
symbol by
H̃(q, p) = ⟨q|H(Q, P )|p⟩ ⟨q|p⟩−1 (2.35)
which is the matrix element of the quantum Hamiltonian or-
dered with all Q’s to the left of all P ’s (as in the
Weinberg’s textbook on QFT). Note that H̃ is complex in gen-
eral.
Plugging the matrix element (2.34) into Eq. (2.31) we obtain
now the exact representation
∫ ∏n ∏ dps
n+1
′ ′
⟨q , t |q, t⟩ = dqr N
(2.36)
r=1 s=1
(2π)

n
× exp{i [(qk+1 − qk )pk − ϵH̃(qk+1 , pk )]}
k=1

where by definition q0 = q, qn+1 = q . An interesting inter-
pretation comes out if we take the limit as n → ∞, which
must be possible since the left hand side is independent of
n. Then
∫ ∏n ∏ dps
n+1
′ ′
⟨q , t |q, t⟩ = lim dqr N
(2.37)
n→∞
r=1 s=1
(2π)

n+1
× exp{i [(qk − qk−1 )pk − ϵH̃(qk , pk )]}
k=1
∑N
6
Here and always products like pq really indicate α=1 pα q α .

27
Now construct the broken path with vertices at points Xk =
(qk , pk ) at times tk , and let X(t) = (q(t), p(t)) be a continuous

path in phase space on [t, t ], matching the broken path at
the vertices Xk : then the exponent in Eq. (2.36) is an
approximation to the Riemann’s integral along X(t)
∫ ′
t
I[X] = [pdq − H̃(q, p)dt] (2.38)
t

which is of course the classical action for the tilded


Hamiltonian. Therefore we can loosely interpret Eq. (2.37)
as an integral over the space of continuous paths in phase
′ ′
space from q at time t to q at time t , and write accordingly

′ ′
⟨q , t |q, t⟩ = DX(s) eiI[X] (2.39)

where the (functional) measure is formally given by


∏ dp(s) ′
DX(t) = dq(s) , s ∈ [t, t ] (2.40)
s
(2π)N

Digression 1: in alternative, one can order the Hamiltonian


with all Q to the right of all P . Then

dp
⟨qk , tk |qk−1 , tk−1 ⟩ = exp[i(qk − qk−1 )p − iϵH̄(p, qk−1 )] (2.41)
(2π)N
would be the result, with the (p, q)-symbol given by
H̄(p, q) = ⟨p|H(Q, P )|q⟩ ⟨p|q⟩−1 (2.42)
In that case the path integral formula becomes
∫ ∏n ∏ dps
n+1
′ ′
⟨q , t |q, t⟩ = lim dqr (2.43)
n→∞
r=1 s=1
(2π)N

n+1
× exp{i [(qk − qk−1 )pk − ϵH̄(pk , qk−1 )]}
k=1

and again q0 = q and qn+1 = q . Another popular ordering is due
to H. Weyl, leading to the midpoint rule: one computes the
classical Hamiltonian in the midpoint

qk + qk−1
q̄k =
2
28
so that

dp
⟨qk , tk |qk−1 , tk−1 ⟩ = exp[i(qk − qk−1 )p − iϵHcl (q̄k , p)] (2.44)
(2π)N
where Hcl (p, q) is now the classical Hamiltonian, and the path
integral becomes
∫ ∏n ∏ dps
n+1
′ ′
⟨q , t |q, t⟩ = lim dqr (2.45)
n→∞
r=1 s=1
(2π)N

n+1
× exp{i [(qk − qk−1 )pk − ϵH̄(pk , q̄k−1 )]}
k=1

This is a simple consequence of Weyl’s correspondence rule,


according to which for a broad class of functions Hcl (p, q)
there correspond the integral operator HW on L2 (Rn ) (the
Weyl quantum Hamiltonian) defined by
∫ ( ′ )
′ 1 ′
ip(q −q) q +q
(HW ψ)(q ) = e Hcl p, ψ(q)dp dq (2.46)
2π 2
The kernel of this Weyl transform
∫ ( ′ )
′ 1 ′
ip(q −q) q +q
⟨q |HW |q⟩ = e Hcl p, dp (2.47)
2π 2
is easily seen to to be the matrix element of the quantum
Hamiltonian given by

HW (Q, P ) = h(x, y)eixQ+yP dx dy (2.48)

where

−2
h(x, y) = (2π) dq dp e−ipx−iqy Hcl (p, q) (2.49)

is the Fourier transform of the classical Hamiltonian, gen-


erally interpreted in distributional sense.
Digression 2: instead of the amplitudes considered so far,
′ ′
we may consider the mixed amplitude ⟨p , t |q, t⟩. Clearly

′ ′ ′ ′ ′ ′ ′ ′ ′
⟨p , t |q, t⟩ = ⟨p , t |q , t ⟩ ⟨q , t |q0 , t⟩ dq (2.50)
∫ ′
dq iI[X]−ip q
′ ′
= DX(t) e (2.51)
(2π N /2

29
The functional in the exponent
∫ t′
′ ′ ′ ′
I[X] = I[X] − ip q =
˜ [pq̇ − H(p, q)]ds − ip q (2.52)
t

is the action for the mixed boundary conditions q(t) = q,


′ ′
p(t ) = p , so Eq. (2.50) is a path integral representation of
′ ′
the kernel ⟨p , t |q, t⟩

′ ′ ˜
⟨p , t |q0 , t⟩ = D̃X(t)eiI[X] (2.53)

provided we identify the measure over the space of paths


′ ′
starting at q at time t and ending at p at time t (with any

value for q ) as

dq
D̃X(t) = DX(t) (2.54)
(2π N /2
In the same way
∫ t′
¯
I[X] = [pq̇ − H(p, q)]ds + ipq (2.55)
t
′ ′
is the action for the mixed boundary conditions q(t ) = q ,
p(t) = p and corresponds via path integral to the kernel
′ ′
⟨q , t |p, t⟩, with the measure
dq
D̄X(t) = DX(t) (2.56)
(2π N /2
Finally
∫ ′
t
′ ′
ˆ
I[X] = [pq̇ − H(p, q)]dt + ipq − ip q
t
∫ t

= − [q ṗ + H(p, q)]ds (2.57)


t
′ ′
is for the boundary conditions p(t ) = p , p(t) = p and gives the
′ ′
momentum to momentum amplitude ⟨p , t |p, t⟩, with measure

dq dq
D̂X(t) = DX(t) (2.58)
(2π)N
Digression 3: the formulae we have given made exclusive
use of canonical coordinates (q k , pj ). But one can use other

30
coordinate systems on the phase space. To write the func-
tional integral in general coordinates, say z A = (z 1 , . . . , z 2N ),
let us introduce the symplectic 2-form 2σ = σAB dz A ∧ dz B whose
components are7
∑ ∂pk ∂q k
σAB = 2 [A ∂z B]
= {z A , z B }−1 (2.59)
k
∂z
and the Liouville one-form ω = ωA dz A , such that σ = dω, or
locally σAB = ∂A ωB − ∂B ωA . Then the functional integral takes
the form
∫ ∏ ∏
′ ′
⟨z , t |z, t⟩ = Dz [det(σ)]1/2 exp iI[z], Dz = dz A (t) (2.60)
t t,A

In fact the invariant measure under coordinate changes in


phase space is just the one written, since σAB (z) transforms
as a second rank tensor under diffeomorphisms (and it is
invariant under canonical transformations). Here the action
is

I[z] = (ωA ż A − H(z))dt (2.61)
′ ′
and the integral is over all path from z A to z A in time t − t
that are subject to the constraints
′ ′ ′ ′
ωA (z(t))δz A (t) = ωA (z (t ))δz A (t ) = 0 (2.62)
fixing only N of the 2N coordinates at the boundary points.
Note that these are the boundary conditions necessary for a
well defined variational problem, which states the differen-
tiability of the action in the variables z A (t): δI[z]/δz A (t) = 0.
As is well known this gives the Hamilton equations of motion
in general coordinates
∂H
σAB ż B = A (2.63)
∂z
The given formulas can be directly obtained from the change
of variables (q j , pk ) → z A applied to the Liouville integra-
tion measure which for N degrees of freedom is given by
the 2N -form σ N = σ ∧ σ ∧ · · · ∧ σ/N ! (see Henneaux & Teitel-
boim7 “Quantization of Gauge Systems”, or M. Chaichian and
A. Demichev20 “Path Integral in Physics, Vol. 1”).
7
Recall that ) the Poisson bracket is defined by {F, G} =
∑ ( ∂F ∂G
k ∂q k ∂pk
−∂G ∂F
.
∂q k ∂pk

31
2.3 The Lagrangian path integral
One the most important cases encountered in physics is when
the Hamiltonian is a quadratic function of momenta with q-
independent coefficients, the paradigmatic example being of
course
P2
H= + V (Q) (2.64)
2m
With this Hamiltonian the momentum integrals are easily done
by using the N -dimensional Fresnel integral
∫ ∞ (√ )N
2 πi
eiap dN p = (2.65)
−∞ a
The result is the Feynman’s path integral formula (as always

q0 ≡ q and qn+1 ≡ q )
′ ′
( m )N (n+1)/2 ∫ ∏n
⟨q , t |q, t⟩ = lim dN qr (2.66)
n→∞ 2πiϵ
r=1
{ n+1 [ ]}
∑ m(qk − qk−1 )2
× exp i − ϵV (qk )
k=1


where ϵ = (t − t)/(n + 1) so n → ∞ is equivalent to ϵ → 0. We
see that in the Lagrangian formalism the naïve definition
of the measure is not correct, an additional ϵ-dependent
factor being needed.
For a broken path the exponent in (2.66) is the Riemann
integral of the Lagrangian, namely
∫ ′
t ( )
1 2
mq̇ − V (q) dt
t 2

and for a continuous path passing the points qk at times tk


is an approximation of the same integral. It is customary
to absorb the first factor of (2.66), which does not depend
upon the dynamical variables, into the measure Dq(t). We
shall therefore represent the indicated limit again as a
path integral
∫ [ ∫ ′ ]
t
′ ′
⟨q , t |q, t⟩ = Dq(s) exp i L(q, q̇)ds (2.67)
t

32
Additional corrections appear for Hamiltonians of the kind

1 ∑ ij 1
H= A (q , . . . , q N )pi pj + V (q 1 , . . . , q N )
2 i,j

with q-dependent kinetic energy. Let q = (q 1 , . . . , q N ) and qk =


(qk1 , . . . , qkN ) the k-th point on the broken path, for k = 1, . . . , n:
the momentum integration produces an additional factor in
the measure,
∏√
M (q1 , . . . , qn ) = det[Ak ]
k

where Ak is the N ×N matrix Aij (qk ), coming from the quadratic


Fresnel integral. In the limit n → ∞ this becomes a product
over time of a function of q evaluated at time t and is called
a local measure. The path integral becomes
∫ [ ∫ ′ ]
t
Dq(s)µ[q] exp i L(q, q̇)ds (2.68)
t

∏ √
where the local measure reads µ[q] = t det[A(q(t)]. This
can be exponentiated giving a correction to the classical
Lagrangian8 by noting that, formally
∏ ∑
M (q(t)) = exp log M (q(t)) (2.69)
t t

where M (q(t)) = det[A(q(t)]; but for continuously increasing
number of division points, with ti+1 − ti = ϵ, we have (with
f (i) = f (ti ))
∑ ∫ ∫
1∑ 1
f (t) = lim ϵf (i) = lim f (t)dt = δ(0) f (t)dt
ϵ→0 ϵ ϵ→0 ϵ
t i

The functional integral (2.68) becomes


∫ [∫ ′ ]
t
Dq(s) exp i Lef f (q, q̇)ds (2.70)
t

8
If we restore ℏ the local measure is seen to be independent of it, and
thus it represents a quantum correction to the classical Lagrangian.

33
where (restoring ℏ to emphasize the quantum nature of the
correction)

Lef f = L(q, q̇) + δ(0) log M (q) (2.71)
i
9
In passing by, we used limϵ→0 1/ϵ = δ(0) ∫ ; this can be obtained
by comparing the discretization
∑ of δ(τ − t)f (t)dt
∑ = f (τ ), which
reads f (i) = limϵ→0 k ϵf (k)δ(ti − tk ), with f (i) = k f (k)δik , where
δik is the Kronecker delta. More details are found in many
textbooks, e.g. Weinberg’s first volume book, in Ch.[9].
The quantum correction in (2.71) was shown to be necessary
to correct the naive perturbation theory of certain quantum
field theories like the non linear sigma models or chiral
theories, as was shown in.35

2.4 Path integral for constrained systems


This section resumes very quickly known results on dynamical
systems endowed with constraints, the theory whose funda-
mentals were laid down by Paul Dirac. For a more detailed
description we recommend Ref.7
Several Hamiltonian dynamical systems are subject to a vari-
ety of constraints, generally given by phase space functions
χj (p, q) such that the motion is restricted to a submanifold
S of the phase space, called the constraint surface,

χj (p, q) = 0, j = 1, . . . , J

The most important situation for the appearing of S is when


the Lagrangian is singular, in the sense that the equations
defining the momenta, pn = ∂L/∂ q̇ n , cannot be inverted to give
the velocities q̇ n as uniquely defined functions of the vari-
ables p, q, because the matrix ∂ 2 L/∂ q̇ n ∂ q̇ m is singular. Thus
some of the constraints are given independently on the equa-
tions of motion, they simply result from the structure of
the Lagrangian and are called primary constraints. Others,
called secondary constraints, could result from the consis-
tency condition χ̇j (p, q) = 0, which is not always a consequence
9
Note that in the time domain δ(0) = lim ω/2π = lim ν = limϵ→0 1/ϵ, where
ν is the frequency.

34
of the equations of motion and the primary constraints. We
assume here that we have given a complete list of them so
that no new, terziary or higher order constraints can arise
from fulfilling the equations of motion. We will indicate
the primary constraints with indices l, m, n = 1, . . . , M , the re-
maining for j = M + 1, . . . , J being secondary.
For example, one form of the action of a relativistic par-
ticle is
∫ ∫ √
I[x] = Ldτ = −m −ẋ2 dτ (2.72)

where τ is an arbitrary parameter along space-time curves.


The momentum is
∂L mẋµ
pµ = = √ (2.73)
∂ ẋµ −ẋ2
from which the constraint follows (the mass shell condition)

pµ pµ + m 2 = 0 (2.74)

independently on the equations of motion, so the particle


is constrained to move on the mass shell, a good prediction
indeed! But note the embarrassing fact that the canonical
Hamiltonian vanishes, pµ ẋµ − L = 0. In any case the time
derivative of the constraint vanishes identically so that
there are no new, secondary constraints. A second example,
this time from field theory, is electromagnetism, with the
Lagrangian density
1
L = − F 2 + J µ Aµ (2.75)
4
Here the canonical momentum is
∂L
π µ (t, ⃗x) = = −F 0µ (2.76)
∂∂0 Aµ (x)
so there is a primary constraint, π 0 = 0. Since moreover, by
the field equations,

∂j π j = −∂j F j0 = J 0 (2.77)

and this contains no second order time derivatives, there is


another, secondary, constraint, ∂j π j −J 0 = 0, namely the Gauss

35
law. No other constraints arise from the field equations,
so these are the only ones.
Returning to the general case, the primary constraints
can be handled by the Lagrange multipliers method, using
the Hamiltonian
∑ ∑ ∑
H= pk q̇ k − L + um χm ≡ Hc + um χm (2.78)

and treating all variables as independent. Using Poisson


brackets the equations of motion are10

Ḟ = {F, Hc } + um {F, χm }, χm (p, q) = 0 (2.79)
m

where Hc is the canonical Hamiltonian. From (2.79), the


phase space trajectory will stay on the constraint manifold
S if and only if on S

χ̇j = 0 (2.80)

or11 , with j = 1, . . . , J, m = 1, . . . , M

{χj , H} ≈ {χj , Hc } + um {χj , χm } = 0 (2.81)
m

We are assuming that no new constraints arise from (2.81),


since the functions χj (p, q) form a complete set of them. But
clearly, if for some subset {γa } of the constraints (possi-
bly empty or even for the totality of them) the brackets
{γa , γb } = 0 on S, some of the multipliers ua cannot be deter-
mined by the equations of motion and therefore are arbitrary
functions of time. The presence of arbitrary functions of
time (or of arbitrary functions of space-time coordinates
in field theory) in the equations of motion is the dis-
tinctive sign for the existence of gauge symmetries among
the canonical variables, symmetries whose parameters can be
specified arbitrarily at each instant of time.
Following Dirac, we shall define a first class function

10
No term {F, um }χm is present here, because χm (p, q) = 0.
11
A weak equation, written F (p, q) ≈ 0, is one that becomes an equality
on the surface χa (p, q) =
∑0 jbut is non vanishing elsewhere in phase space.
Equivalently, F (p, q) = λ χj .

36
as any function F (q, p) which weakly commutes with all con-
straints

{F, χj } ≈ 0 (2.82)

or, equivalently,

{F, χj } = uk χk (2.83)
r

Functions which are not first class are called second class.
For example, from ∑ m Eq. (2.81) we see that the total Hamil-
tonian H = Hc + u χm is a first class function. Another
set comes from solutions of Eq. (2.81), since the general
solution of the homogeneous equation

um {χj , χm } = 0 (2.84)
m

just says that there is a combination of constraints which


is of the first class. To be specific, given a linearly in-
dependent complete set of solution for um → Vam , a = 1, . . . , A,
we set γa = Vam
∑χm a and then the general solution of (2.84)
m m
will be u = a v Va ; the functions γa are a complete set
of primary first class constraints. The coefficients v a are
not determined by the consistency condition (2.81) and are
arbitrary as a matter of fact. Thus the constraints them-
selves can be of either type, with first class constraints
always appearing in the presence of gauge symmetries. The
total Hamiltonian (2.78) can now be rewritten in the form

H = Hc + U m χm + v a γa (2.85)

where U m denote a particular solution of (2.81). This split-


ting into first and second class constraints is not unique,
since U m + ua Vam is another possible particular solution.
For the time being we shall first discuss the second class
case.

2.4.1 Second class constraints


Let us assume for the time being that all constraints are
of the second class type. Then Eq. (2.81) can be uniquely
solved for the Lagrange multipliers um . Therefore second

37
class constraints are characterized by the condition that
the Poisson brackets

Cab = {χa , χb } = −Cba (2.86)

are the entries of an invertible matrix, with inverse C ab ,


even on the constraint surface χa ≈ 0. The number of second
class constraints is then necessarily even. A new bracket,
the Dirac bracket, can be defined by

{F, G}∗ = {F, G} − {F, χa }C ab {χb , G}, C ab Cbc = δca (2.87)

which has the noteworthy property of being consistent with


the constraints, namely, {F, χa }∗ = 0. Not only this: the
Dirac bracket is obviously anti-symmetric, {F, G}∗ = −{G, F }∗ ,
and satisfies the Jacobi identity

{E, {F, G}∗ }∗ + {F, {G, E}∗ }∗ + {G, {E, F }∗ }∗ = 0 (2.88)

The constraint surface equipped with the Dirac bracket is


the phase space of the independent degrees of freedom over
which the path integral should be done. To describe this
quantitatively, we may choose coordinates y i on S (not nec-
essarily canonical coordinates) and extend them off the
surface in such a way that

{y i , χa } ≈ 0 (2.89)

This can always be achieved since given any extension, say


ȳ i , the coordinates y i = ȳ i + λia χa will satisfy this if we
choose

λia = −C ab {y i , χb },

We have then, locally, a new system of coordinates on phase


space, (y i , χa ), together with the original canonical variables
xα = (q r , ps ). The inverse of the symplectic 2-form on S is the
anti-symmetric tensor given (in terms of Dirac brackets) by

σ ij = {y i , y j }∗ ≡ {y i , y j } − {y i , χa }C ab {χb , y j } (2.90)

and with σij we will denote the symplectic form itself. By


what we know about the path integral in general coordinates,

38
we can write
∫ √
path integral = Dy det(σij ) exp iI[y] =
∫ ∏ √
= DyDχa δ(χa ) det(σij ) exp iI[y]
t,a

where I[y] is the canonical action restricted on S. It would


be clearly much more convenient to write the path integral
in the original canonical variables, since finding the co-
ordinates y i is equivalent to solve the constraints, which
may be a difficult task (particularly in field theory). Due
to condition (2.89), the symplectic form in the coordinates
(y i , χa ) is
( ij ) ( )
αβ σ 0 σij 0
σ = , σαβ = −1 (2.91)
0 Cab 0 Cab
Therefore, given 2p constraints among 2N canonical coordi-
nates, we have the measure for xα = (q r , ps ),


2N ∏ ∏ ∏ ∏
dxα = dy i dχa det(σαβ )1/2 = dy i dχa det(σij )1/2 det(Cab )−1/2
α=1 i a i a

where i = 1, . . . , 2(N − p), a = 1, 2, . . . 2p, and the path integral for


systems with second class constraints finally reads
∫ ∏ ∏
path integral = Dx det(Cab )1/2 δ(χa ) exp iI[x] (2.92)
t t,α

Due to the presence of δ(χa ) the Hamiltonian in I[x] can be


taken as the usual canonical Hamiltonian without wondering
about the constraints. This formula would be self-evident
if the constraints were simply given, for example, by

p1 ≈ 0, q 1 ≈ 0, {q 1 , p1 } = 1

For in fact, in this case,

∑n ( )
∗ ∂F ∂G ∂F ∂G
{F, G} = −
j=2
∂q j ∂qj ∂pj ∂q j

39
and Eq. (2.92) reduces to the usual canonical path inte-
gral over the independent variables (q 2 , . . . , q n , p2 , . . . , pn ). This
is actually not really a special case: it can be proven
that it is always possible to choose local coordinates
(Q1 , . . . , Qn , P1 , . . . , Pn ) around S such that S itself is given
locally by equations Q1 = · · · = Qk = P1 = · · · = Pk = 0. Eq. (2.92)
reduces then to the usual path integral over the remaining
variables (Qk+1 , . . . , Qn , Pk+1 , . . . , Pn ) endowed with the canonical
Poisson bracket (see1, 7 for lot of details).

2.4.2 First class constraints and the Faddeev


formula
Let us now assume that all constraints are first class
functions, denoted γa ≈ 0, for which the defining condition
is {χj , γa } ≈ 0 for all constraints; in particular

{γa , γb } ≈ 0 ⇐⇒ {γa , γb } = Cab


c
γc (2.93)
c
where the Cab can be functions on phase space, in general.
If they are numerical constant they generate evidently a
certain Lie algebra. The set {γa } contains in general both
primary and secondary constraints. In both cases, due to
(2.93) the constraints generate canonical transformations
on S via the usual formula

δF = {F, ϵa γa } ≈ ϵa {F, γa } (2.94)

where the ϵa are arbitrary functions of phase space variable


and time t.
It is known that the primary first class constraints can be
considered gauge transformations, not changing the physical
state. For in fact, a simple computation with the Poisson
brackets, using the equations of motion and the preservation
in time of the constraints, shows that if δϵ q j ≈ ϵa (t){q j , γa } and
δϵ pj ≈ ϵa (t){pj , γa }, then the functions q j + δϵ q j and pk + δϵ pk will
still obey the equations of motion for arbitrary ϵa (t) if so
do the canonical variables (q j , pk ). And if ϵ(0) = 0 they also
obey identical initial conditions, so there in no unique
solution to the Cauchy problem. The only way out is to
consider the transformations generated by the first class
constraints as gauge transformations, so the physical states

40
correspond to more than one pair of canonical variables
(q, p). An application of the Jacobi identity then shows that
the bracket {γa , γb } and {γa , H} also generate further gauge
transformations.
A symmetry of the equations of motion that can be speci-
fied independently at any time t12 is a gauge symmetry. We
stress that this is a physical interpretation of the for-
malism which is forced on by the mathematical properties of
the constraints. There is, however, a problem: what is the
role of the secondary first class constraints? The Dirac’s
conjecture is that they too correspond to gauge transfor-
mations. This is not a mathematical statement, and in fact
it is false in general13 . However it seems natural that the
time development could be accompanied by arbitrary gauge
transformations. Moreover, the bracket {γa , γb } is a linear
combination of first class constraints involving in general
both primary and secondary constraints, and the bracket too
generates gauge transformations.
With this interpretation the Hamiltonian should contain
all first class constraints, not just the primary ones,
i.e. HE = H + ua γa , where a runs over the full set of first
class secondary constraints and H = Hc + U m χm + v a γa is the
previously used total Hamiltonian including only the primary
first and second class constraints. This HE is known as
the extended Hamiltonian. The extended equations of motion

Ḟ ≈ {F, HE }

are not required by the original Lagrange equations and


represent a true extension of the original Hamiltonian for-
malism. For gauge invariant functions G, such that by
definition {G, γa } ≈ 0, they are fully equivalent to the orig-
inal Lagrange equations.
With the existence of gauge degrees of freedom there are
therefore more than one pair of canonical variables corre-
sponding to a given physical state. We shall assume that the
variables connected by finite gauge transformations gener-
ated by the constraints form an orbit of some gauge group
12
That is the group parameters are functions.
13
The innocent Lagrangian L = ey ẋ2 is an example.

41
G, a subset of phase space over which G acts transitively14 .
The global existence of the orbits is due to the integrabil-
ity conditions satisfied by the transformations generated
by the brackets (2.93). Each orbit corresponds uniquely to a
physical state and viceversa, so it is natural to define the
set of physical states, the so called reduced phase space,
as the quotient space F = S/G, which we assume here to be
a symplectic manifold15 . This last property can actually be
proved, the essential reason being that the null eigenvec-
tors of the symplectic 2-form on the constraint surface are
all tangent to the orbits (for a full proof, see7 ).
The functions defined on the reduced phase space are
all gauge invariant functions, i.e. {z, γa } ≈ 0; given a
complete set of them we can parametrize F (in principle)
with independent coordinates z n (non canonical in general)
so the path integral is taken in the general form
∫ ∏
path integral = Dz n det(σrs )1/2 exp iI[z] (2.95)
t

Here I[z] = [an (z)ż n −H(z)]dt is the action restricted on F, with
H(z) the gauge invariant Hamiltonian, and up to boundary
terms which are needed to accommodate the required boundary
conditions. In is in general an extremely difficult task
to find a complete set of gauge invariant functions Oj ,
satisfying

∀a, {Oj , γa } ≈ 0 (2.96)

which also generate all gauge invariant functions in the


sense that for any function F (p, q) such that {F, γa } ≈ 0 then
F ≈ F (O1 , . . . ), not to mention the essentially local nature of
this procedure. Apart from special cases, it is generally
easier to deal with the path integral over the original phase
space. A widely used method is to fix the gauge symmetry
by choosing a set of “gauge fixing functions” gb (p, q) ≈ 0, one
for each first class constraint, whose aim is to select one
unique element from each gauge orbit.
14
Strictly speaking we are assuming it is a group, partly because this
is the more interesting case.
15
This may be too naïve, and the connected component of the identity
of G should be used.

42
Figure 2.1: The reduced phase space is the set of gauge
orbits

There is a large freedom as to the choice of the functions


ga , but there are two simple requirements that a good gauge
fixing procedure must meet:
(i) the gauge choice

gb ≈ 0

must be reachable by means of gauge transformations, in


order not to change the gauge invariant content of the
theory,
(ii) the chosen functions must fix the gauge, that is given
an arbitrary gauge generator Υ = ϵa γa , the gauge variation

δgb = {Υ, gb }

should not vanish, or ∀ϵa

ϵa {γa , gb } ̸= 0

which implies that {γa , gb } is an invertible matrix. That is


to say, the gauge fixing functions enlarge the set of first
class constraints into a set of second class constraints,

43
for which there is available the Dirac procedure.
One can say that the surface ga = 0 geometrically intersects
each gauge orbit only once, although this is a much more
stringent global condition than the ones locally given here.
In fact there may be topological obstructions to a global
gauge fixing procedure, a problem known in gauge theories
as the Gribov obstruction.
The path integral can be written now as for any other
system with second class constraints χα = (γa , gb ) ≈ 0, now
embracing the constraints γa plus the gauge fixing functions
gb , in the form (see Eq. (2.92))
∫ ∏ ∏
path integral = Dxm det(Cαβ )1/2 δ(χα ) exp iI[x] (2.97)
t t,α

where xm = (pr , q s ) are canonical coordinates on the uncon-


strained phase space and
[ ]
0 {γa , gb }
Cαβ = {χα , χβ } ≈ (2.98)
−{γa , gb } {ga , gb }
The zeroes in the upper left corner correspond to {γa , γb } ≈ 0,
since the γa are first class functions. The square root of
the determinant is then
det(Cαβ )1/2 = |det{γa , gb }| (2.99)
Note that Mab = {γa , gb } is the gauge variation δa gb of the
gauge fixing functions. The equation (2.97) with (2.99)
substituted in is known as the Faddeev formula.

2.5 Functional calculus


The art of integrating functionals is an important aspect of
the path integral formalism. It is based on few important
rules:
(i) the splitting principle: the integral over all paths

from q to q is the same as the integral over all path
′′ ′′
passing through an intermediate point q at some t followed
′′
by integration over dq . This is clearly due to the additive
property of the action integral
∫ t′ ∫ t′′ ∫ t′
(pq̇ − H)dt = (pq̇ − H)dt + (pq̇ − H)dt (2.100)
t t t′′

44
combined with the principle that the measure splits, namely,

′′
Dq[q,q′ ] = Dq[q,q′′ ] Dq[q′′ ,q′ ] dq (2.101)

(ii) translational invariance, i.e. invariance under the



map q(s) → q(s) + ϕ(s), where ϕ(t) = ϕ(t ) = 0
∫ ∫
iI[q]
D[q(s) + ϕ(s)] F [q]e = Dq(t) F [q]eiI[q] (2.102)

(iii) The fundamental theorem of calculus



δ [ ]
Dq(s) F [q]eiI[q] = 0 (2.103)
δq(s)
(iv) insertion of derivative are handled by the rule
∫ ∫
d
Dq(s)q̇(t) · · · eiI[q]
= Dq(s)q(t) · · · eiI[q] (2.104)
dt
In fact, by taking appropriate limits it should be pretty
clear that these commute with the functional integral, pro-
vided the dots indicate arbitrary insertions at times dif-
ferent from t.
Of course we need to learn how to integrate functionals like
F [q]. This is done in principle by inserting into Eq. (2.66)
the numerical function Ft1 ,...,tn (q1 , . . . , qn ), which indicates the
value of the functional F [q] to be integrated for the broken
path passing q1 at t1 ,. . . ,qn at tn , and then taking the limit
n → ∞. But to which matrix element does this correspond?
For many important cases this complicated procedure can be
avoided, but for completeness we will give an explicit com-
putation in the next section for a special functional. With
this question in mind we give here few examples using the
Lagrangian path integral.

2.5.1 Correlation functions and generating func-


tionals
Suppose we want to calculate the integral (we restrict con-
siderations to systems with one degrees of freedom, the
extension to many being straightforward)


C(t1 , . . . , tn ) = Dq(s)q(t1 ) · · · q(tn )eiI[q] , t > tn > · · · > t1 > t(2.105)

45
Here we have a functional which depends on the path only at
finitely many points t1 ,...,tn . We set q(tk ) = qk ; using the
splitting principle, we can first integrate over all paths
passing q1 , . . . , qn at times t1 , . . . , tn (by the way, this is the
reason why we need to time order the sequence of times), and
then integrate over all possible values of the coordinates
qk . But the path integral over all path joining qj at tj
to qj+1 at tj+1 is the probability amplitude ⟨qj+1 , tj+1 |qj , tj ⟩, so
that, by also noting that Q(tj ) |qj , tj ⟩ = qj |qj , tj ⟩, we obtain
∫ ∏
n
′ ′
C(t1 , . . . , tn ) = dqj ⟨q , t |qn , tn ⟩ qn ⟨qn , tn |qn−1 , tn−1 ⟩ qn−1 · · · q1 ⟨q1 , t1 |q, t⟩
j=1
∫ ∏
n
′ ′
= dqj ⟨q , t |Qn (tn )|qn , tn ⟩· · · ⟨q1 , t1 |Q(t1 )|q, t⟩ (2.106)
j=1

Using the completeness relations |q, t⟩⟨q, t| dq = 1

′ ′
Dq(s)q(t1 ) · · · q(tn )eiI[q] = ⟨q , t |Q(tn ) · · · Q(t1 )|q, t⟩ (2.107)

where Q(s) is the Heisenberg position operator at time s.


We see that the result is the matrix element of the time
ordered sequence of operators Q(tn ) · · · Q(t1 ), so we shall write
(2.107) for an arbitrary order of the sequence (t1 , . . . , tn ) as

′ ′
Dq(s)q(t1 ) · · · q(tn )eiI[q] = ⟨q , t |T [Q(t1 ) · · · Q(tn )]|q, t⟩ (2.108)

This is one of the key equations of the functional formalism.


By the rule (v), the functional integral defines actually
a time ordering T ∗ , called covariant time ordered product,
such that when derivatives are present one has16

d
T ∗ [Q̇(t) · · · ] = T [Q(t) · · · ]
dt
We can extend Eq. (2.108) to functionals which admit a

Volterra’s expansion around some fixed path on [t, t ]17 (not
16
Equations valid with arbitrary insertions in the path integral are
also valid as operator equations.
17
They are similar to what Wiener called “analytic functionals”.

46
to be integrated over), for example q(t) = 0
∑ 1 ∫ ∑
F [q] = dt1 . . . dtn Fi1 ,...,in (t1 , . . . , tn )q i1 (t1 ) · · · q in (tn )(2.109)
n
n! i ,...,i 1 n

where supp F ∈ [t, t ] (this is for many degrees of freedom,
n

with q i the i-th coordinate, but also for a single degrees


of freedom if we interpret q i (t) as the i-th power of q(t)).
For a Volterra functional, by definition, we set
∫ ∑ 1 ∫
iI[q]
Dq(t) F [q(t)]e = dt1 . . . dtn
n,i ,...,i
n!
1 n
′ ′
×Fi1 ,...,in (t1 , . . . , tn ) ⟨q , t |T [Qi1 (t1 ) · · · Qin (tn )]|q, t⟩ (2.110)
which can be formally abbreviated as

′ ′
Dq(s) F [q(s)]eiI[q] = ⟨q , t |T {F [Q]}|q, t⟩ (2.111)

It is understood that if F [q] depends on time derivatives


of q(t) one has to do appropriate integration by parts in
Eq. (2.109), but this requires valid boundary conditions,
e.g. the vanishing of F(t1 , . . . , tn ) at the boundary points t,

t . Alternatively, the covariant T ∗ -products must be used,
but only if F [q] is differentiable do the two procedures give
the same result.
The question from which we started is answered now: the
functional integral gives matrix elements of time ordered
products of operators. We did the analysis in configuration
space for simplicity, but the same results are true in the
phase space version of the functional integral.
Let us consider the effect of time translation on the matrix
elements: we are interested in
′ ′ ′ ′
⟨t + s, q |T [Q(t1 ) · · · Q(tn )]|q, t + s⟩ = ⟨t , q |U (−s)T [Q(t1 ) · · · Q(tn )]U (s)|q, t⟩

In the following, without loss of generality, we restrict


transition amplitudes to the interval [0, T ] for paths from
q = 0 to some other value q of q(s) at s = T . The correlations
can be grouped into a single generating functional, defined
by the matrix element (with apologies for the multiple use
of T )
∫ T ∫ ∫T
Z[f ] = ⟨q, T |T exp(i Q(s)f (s)ds)|0⟩ = Dq(s)eiI[q]+i 0 q(s)f (s)ds
(2.112)
0

47
in the sense that functional derivatives of Z[f ] with respect
to f (t) at f = 0 give directly the required chronological
functions (times in ). We would appreciate the power of the
functional integral by computing Z[f ] for a free particle
with unit mass and action

1 T 2
I[q] = q̇ ds (2.113)
2 0
The symmetric Dirichlet Green function
t(T − s)
∆(t, s) = (t − s)θ(t − s) − , ∆(t, s) = ∆(s, t) (2.114)
T
obeys
d2
∆(t, s) = δ(t − s) (2.115)
dt2
with the boundary conditions ∆(0, s) = ∆(T, s) = 0. The function
∫ T ∫ t ∫
t T
q0 (t) = ∆(t, s)f (s)ds = (t − s)f (s)ds − (T − s)f (s)ds(2.116)
0 0 T 0
is then a solution of q̈ = f (t) with the Dirichlet boundary
conditions: q0 (0) = q0 (T ) = 0. Therefore we can change the
variable in the path integral from q(t) to q̄(t), where

q(t) = q̄(t) + q0 (t)

A simple calculation gives the action to be integrated over


as
∫ T ∫
1 q T
I[q] + f (t)q(t)dt = I[q̄] + f ⋆ ∆ ⋆ f + sf (s)ds (2.117)
0 2 T 0
where the ⋆-symbol indicates the convolution
∫ T
[∆ ⋆ f ](t) = ∆(t, s)f (s)ds
0

Upon substitution the path integral becomes


[ ∫ ]∫
1 q T
Z[f ] = exp i f ⋆ ∆ ⋆ f + sf (s)ds Dq̄(t)eiI[q̄] (2.118)
2 T 0

48
or
[ ∫ T ]
1 q
⟨q, T |0⟩f = ⟨q, T |0⟩ exp i f ⋆ ∆ ⋆ f + sf (s)ds (2.119)
2 T 0

where the numerator, the matrix element in the presence of


the source f , is what we previously indicated with Z[f ]. The
chronological correlation functions are now computed by

δ n ⟨q, T |0⟩f
⟨q, T |T [Q(t1 ) · · · Q(tn )]|0⟩ = (−i) n (2.120)
δf (t1 ) · · · δf (tn ) f =0
We have calculated a non trivial path integral without that
awkward discretization.
Examples: let us define

⟨q, T |T F [q]|0⟩
⟨F [Q]⟩ =
⟨q, T |0⟩

Using (2.120) we have for 0 < t, s < T

qt qt
⟨q, T |Q(t)|0⟩ = ⟨q, T |0⟩ =⇒ ⟨Q(t)⟩ =
T T
ts
⟨Q(t)Q(s)⟩ = −i∆(t, s) + q 2
T2
or, using the previous,

⟨(Q(t) − ⟨Q(t)⟩)(Q(s) − ⟨Q(s)⟩)⟩ = −i∆(t, s)

The Green function is the mean (in the above sense) of the
fluctuations, a general result in quantum field theory and
statistical mechanics of Gaussian systems. Similarly

⟨(Q(t) − ⟨Q(t)⟩)(Q(s) − ⟨Q(s)⟩)(Q(u) − ⟨Q(u)⟩)⟩ = 0

and so on.

49
2.5.2 Commutation relations
Consider for simplicity a free particle of unit mass and
the integral

d2 ′ ′
Dq(τ )q̈(t)q(s)eiI[q] = ⟨q , t |T Q(t)Q(s)]|q, t⟩
dt2

Since

δI
q̈ = −
δq(t)

using rule (iii) we get


∫ ∫ ∫
δeiI[q]
Dq(τ )q̈(t)q(s)eiI[q]
=i Dq(τ )q(s) = −iδ(t − s) Dq(τ )eiI[q]
δq(t)

so we get the equation


d2 ′ ′ ′ ′

2
⟨q , t |T Q(t)Q(s)]|q, t⟩ = −iδ(t − s) ⟨q , t |q, t⟩ (2.121)
dt
On the other hand, since T [Q(t)Q(s)] = θ(t−s)Q(t)Q(s)+θ(s−t)Q(s)Q(t)
we have

d
T [Q(t)Q(s)] = T [Q̇(t)Q(s)]
dt

d2
T [Q(t)Q(s)] = T [Q̈(t)Q(s)] + [Q̇(t), Q(t)]δ(t − s) (2.122)
dt2
But Q̈(t) = 0 and Q̇ = P for a free particle with unit mass, so by
comparison with (2.121) we obtain the standard equal time
commutation relation of quantum mechanics, [P (t), Q(t)] = −i.
With few modifications this result can also be obtained for
a particle in a potential, or a system with many particles
or degrees of freedom or in the Hamiltonian version of the
path integral. The path integral is fully equivalent to the
operator method.

50
2.5.3 Equations of motion and Ward identities
Equations of motion: the operator equations of motion have
a counterpart in the functional formalism. Consider rule
(iv) in the following case

δ ( )
Dq(τ ) q(t1 ) · · · q(tn ) · · · eiI[q] = 0
δq(s)

where the dots indicate insertions depending on times def-


initely different from s, but we leave open the possibility
that anyone of the tj can approach s. Then performing the
functional derivatives we obtain the following relation be-
tween matrix elements
[ ]
′ ′ δI[Q]
⟨q , t |T Q(t1 ) · · · Q(tn ) · · · |q, t⟩ = (2.123)
δQ(s)
∑ n
′ ′
= δ(s − tk ) ⟨q , t |T [Q(t1 ) · · · Q̂(tk ) · · · Q(tn ) · · · ]|q, t⟩
k=1

where the hatted Q is stripped out from the matrix element.


This says that the equations of motion (classically δI/δq = 0)
hold in a time ordered matrix element up to contact terms
which vanish if all arguments tk ̸= s. Moreover, since the
insertions in (2.123) (the dots) are arbitrary anyway, we
have the operator equation
[ ] ∑ n
δI[Q]
T Q(t1 ) · · · Q(tn ) = δ(t − tk )T [Q(t1 ) · · · Q̂(tk ) · · · Q(tn )](2.124)
δQ(t) k=1

Symmetries: suppose that both the action and the measure


are invariant under a pointwise variation of the form δq(t) =
ϵF (q(t))

δI
δI = dt F (q) = 0, Dq(t) = D[q(t) + ϵF (q)]
δq(t)

If ϵ is considered a function of t then the variation of the


action must take the form18

δI = ϵ̇(t)J(t)dt (2.125)

18
The dependence of J(t) on q is hidden. On shell holds the conservation
law J˙ = 0.

51
while the measure still stays invariant. Substituting ev-
erywhere in the functional integral the dummy integration
variable q(s) with q(s) + ϵ(s)F (q), using the invariance of the
measure and (2.125) and expanding to first order in ϵ we
obtain the important equation19

dt ϵ̇(t) ⟨f | T [J(t)Q(t1 ) · · · Q(tn )] |i⟩ (2.126)

n
=i ϵ(tk ) ⟨f |T [Q(t1 ) · · · F (Q(tk )) · · · Q(tn )|i⟩
k=1

As it stands it contains the arbitrary function ϵ(t) so it can


only be true if the local identity holds (simply take the
functional derivative of both members with respect to ϵ(t)
and integrate away the time derivative of the δ-function)
d
⟨f | T [J(t)Q(t1 ) · · · Q(tn )] |i⟩ (2.127)
dt
∑ n
= −i δ(t − tk ) ⟨f |T [Q(t1 ) · · · F (Q(tk )) · · · Q(tn )|i⟩
k=1

known in field theory as the Ward identity. The current


is conserved as a consequence of the symmetry (which is
Noether’s Theorem), J˙ = 0, and Eq. (2.127) is the expression
of this conservation law in the quantum domain. In fact
there is no term containing J˙ inside a T -product anywhere
in Eq. (2.127).

2.6 Quadratic Lagrangians


We considered the use of the functional integral in several
situations of interest but without actually computing it
using the defining infinite limit of (2.66). Altough the
functional integral is rarely calculated using the explicit
definition, it is of interest to see how it actually works.
We give now a computation first given by Gelfand and Yaglom
in a classic important paper. Let us consider the quadratic
Lagrangian
1 1
L = q̇ 2 − ω(t)q 2 (2.128)
2 2
′ ′
19
We indicated by |i⟩ the state |q, t⟩, and similarly ⟨f | = ⟨q , t |.

52
We are interested in the path integral (2.39) with this
Lagrangian, given by Eq. (2.66) as

− n+1
J = lim(2πiϵ) 2 dq 1 · · · dq n
n→∞
[ n ( )]
∑ (qk+1 − qk )2 ϵ
× exp i − ω(kϵ)qk2 (2.129)
k=0
2ϵ 2

where for simplicity q0 = 0, qn+1 = q, the time interval is


[0, T ] and ϵ = T /(n + 1). The exponent is

1 ∑ [n] j k qqn q 2
n
A q q − + (2.130)
2ϵ j,k=1 jk ϵ 2ϵ

where the n × n matrix is given by

 
2 − ϵ2 ω1 −1 0 · · 0
 −1 2 − ϵ2 ω2 −1 0 · · 
 
 0 −1 · −1 0 · 
A[n] =


 (2.131)
 · 0 −1 · −1 · 
 · · 0 −1 2 − ϵ2 ωn−1 −1 
0 · · 0 −1 2 − ϵ2 ωn

and ωk = ω(kϵ). Hence we have an integral of the form


∫ ( n ) [ ]
∑ i ∑ [n] j k
n
I = dq · · · dq F
1 n
bj qj exp A q q (2.132)
j=1
2ϵ j,k=1 jk
∫ ∞ ( 2)
(n−1)/2 2 [n] −1/2 iu
= (2πiϵ) (c det A ) F (u) exp du (2.133)
−∞ 2ϵc2
where

n
2
c = [Ajk ]−1 bj bk (2.134)
j,k=1

The last equality can be proved by diagonalizing the matrix


in the exponent and then taking advantage of the rotational
invariance of the measure to rotate the coordinates so that
the vector (b1 , . . . , bn ) → (0, 0, . . . , b). To proceed, let Dn = det A[n] .
Our F is

F = exp(−iqqn / ϵ)

53
so that we have20

q2 [n] −1 q 2 Dn−1
c2 = [([A ) ] nn =
ϵ2 ϵ2 Dn

Finally
∫ ∞ ( ) √ [ ]
iu2 q 2
Dn−1
exp −iu + du = 2πiϵc2 exp −i
−∞ 2ϵc2 ϵ Dn

Using these results into Eq. (2.129) we now have the clean
result
( 2 )
−1/2 iq Dn − Dn−1
J = lim(2πiϵDn ) exp (2.135)
n→∞ 2ϵ Dn
It remains to compute the indicated limit, which since
ϵ = T /(n + 1) is equivalent to ϵ → 0. The determinants Dn
satisfy a very simple recursion relation. Let D0 = 1; we
compute trivially D1 = 2 − ϵ2 ω1 , then

D2 = (2 − ϵ2 ω2 )(2 − ϵ2 ω1 ) − 1 = (2 − ϵ2 ω2 )D1 − D0

and so on and so forth. Thus we write in general


Dn+1 = (2 − ϵ2 ωn+1 )Dn − Dn−1 (2.136)
or more suggestively
Dn+1 − 2Dn + Dn−1
= −ωn Dn (2.137)
ϵ2
Our Dn for fixed T are functions of ϵ which are certainly
divergent21 as ϵ → 0, but we expect that ϵDn will be convergent
(with nϵ fixed) to some function of T , so we define a function
as
f (T ) = lim ϵDT /ϵ (2.138)
ϵ→0

Obviously f (0) = 0 because D0 = 1. From (2.137) we obtain


readily the Gelfand-Yaglom differential equation
f¨ + ω(t)f = 0, f (0) = 0 (2.139)
20
We use the standard formula for the matrix element of the inverse
matrix in terms of minors.
21
The leading divergence is easily seen to be of order a/ϵ for some
constant a depending on T .

54
For example, as ϵ gets very small, with n = T /ϵ, we have
ϵDn+1 − ϵDn f (T ) − f (T − ϵ) df
≃ → (2.140)
ϵ ϵ dT
Similarly
ϵDn+1 − 2ϵDn + ϵDn−1 d2 f
→ (2.141)
ϵ2 dT 2
We still need a boundary condition for f˙(0); but from (2.140)
we have
f˙(0) = lim(D1 − D0 ) = 1 (2.142)
ϵ→0

It follows that f (T ) = limϵ→0 ϵDn is the solution of the differ-


ential equation (2.139) with the boundary conditions f (0) = 0,
f˙(0) = 1. The function

f1 (t) = f (T − t)

has boundary conditions f1 (T ) = 0, f˙1 (T ) = −1 and was actually


the form in which Gelfand and Yaglom first presented their
fundamental result.
The original functional integral was given by Eq. (2.135);
the pre-factor is just (2πif (T ))−1/2 , while the exponent can
be computed as the limit
Dn − Dn−1 f˙(T )
lim = (2.143)
ϵ→0 ϵDn f (T )
In all we have, recalling that J is a transition amplitude
( )
2 ˙
iq f (T )
< q, T |0, 0 >= (2πif (T ))−1/2 exp (2.144)
2 f (T )
Examples: (i) for the free particle with unit mass the
solution of (2.139) with the stated boundary conditions is
f (t) = t; then (2.144) reduces to the well known result
( 2)
−1/2 iq
< q, T |0, 0 >= (2πiT ) exp (2.145)
2T
(ii) the harmonic oscillator has

sin ωt
f (t) =
ω
55
Therefore
( )1/2 ( )
ω iωq 2
< q, T |0, 0 >= exp cot ωT (2.146)
2πi sin ωT 2
(iii) As a less trivial example, in Eq. (2.139) we choose
ω(s) = ω02 e−at . The general solution of (2.139) is
( ) ( )
f (t) = c1 J0 2ωe−at /a + c2 N0 2ωe−at /a

with c1 , c2 determined from the boundary conditions f (0) = 0,


f˙(0) = 1. The kernel follows from Eq. (2.144). In full
generality one can write

f (t) = ρ(t)[a cos θ(t) + b sin θ(t)]

Then the Gelfand-Yaglom differential equation is solved by


∫ t
θ(t) = C ρ(t)−2 dt, C > 0 a constant

with ρ(t) obeying the Pinney differential equation

C
ρ̈ + ω 2 (t)ρ =
ρ3

The path integral for a generic quadratic Lagrangian always


has the form
′ ′ ′ ′
⟨q , t |q, t⟩ = A(t , t)eiB(q ,t;q,t)

An efficient technique to find the exponent is the following.


Let us expand the action in powers of q(t) around the classical
path: with q(t) = qcl (t) + η(t) we have22

1 δ 2 I[qcl ]
I[q] = I[qcl ] + ηj (t) ηk (s)dtds (2.147)
2 δqj (t)δqk (s)
The classical path is the solution of the equations of motion
′ ′
with the boundary conditions qcl (t) = q, qcl (t ) = q , so the fluc-
tuation η(t) will satisfies Dirichlet boundary conditions,
22
The expansion stops at second order because the Lagrangian is
quadratic.

56

namely η(t ) = η(t) = 0. The path integral becomes
∫ [ ∫ ]
′ ′ i δ 2 I[qcl ]
⟨q , t |q, t⟩ = eiI[qcl ]
Dη exp ηj (t) ηk (s)dtds (2.148)
2 δqj (t)δqk (s)
Any quadratic Lagrangian can be reduced to the form

1 1
L = m(t)q̇ 2 − ω(t)q 2 + J(t)q
2 2

Then qcl (s) solves the differential equation

d ′ ′
[m(s)q̇cl ] + ω(s)qcl = J(s), qcl (t) = q, qcl (t ) = q
ds
and
( )
δ 2 I[qcl ] d d
=− m(t) + ω(t) δ(t − s) (2.149)
δqj (t)δqk (s) dt dt
The path integral in (2.148) is formally related to the
determinant of this differential operator, for which a va-
riety of methods exist to compute it. The Gelfand-Yaglom
computation discussed above is actually the computation of
this determinant for the particular case m(t) = 1. This is at

the same time the pre-factor A(t , t) and the action I[qcl ] is the

exponent B(q , t; q, t)23 . As to the determinant, the Gelfand-
Yaglom computation discussed above has been extended to this
more general class of differential operators.40, 41

2.7 The vacuum persistence amplitude


Up to now we have only considered finite time transition in
configuration or phase space. In field theory one is more
interested in vacuum to vacuum transitions in infinite time
when (i), an external source linearly coupled to the field
is perturbing the system, and (ii) there is no external
source but the potential admit scattering states.
Suppose in both cases that the Hamiltonian has a ground
state |0⟩. We obtain from the path integral the desired
23
The classical path depends of course on the undisplayed boundary
data.

57
transition in the presence of a source J(t) from the general
rules of quantum mechanics

⟨0, out|in, 0⟩|J



′ ′ ′ ′ ′
= lim

lim ⟨0, out|q , t ⟩ ⟨q , t |q, t⟩J ⟨q, t|in, 0⟩ dq dq (2.150)
t →∞ t→−∞

Here |in, 0⟩ is the in-vacuum that converges at t = −∞ to the


ground state of the free dynamics (to be defined soon), and
similarly for ⟨0, out| at t = +∞, and
∫ ( ∫ ′ )
t
′ ′
⟨q , t |q, t⟩J = Dq exp iI[q] + i dtJq (2.151)
t

is the usual constrained path integral. Suppose that J(t)


vanishes outside a finite time interval and (to mimick field
theory) the potential is harmonic near the origin in q-space,

1
V (q) = ωq 2 + V1 (q)
2
where the second term is to be considered the interac-
tion energy24 , i.e. we consider the harmonic oscillator as
the ”free dynamics” mentioned above. Thus strictly speaking
there is no scattering in this theory, only transition prob-
abilities from states to states in the harmonic oscillator
spectrum. But of course it is only in field theory that
these states can be interpreted as multi-particle configu-
rations and the transition amplitudes as matrix elements of
the scattering operator.
To proceed, we also assume E0 = 0 (in field theory we always
omit the vacuum energy). Now |q, t⟩ = eiHt |q⟩ is the eigenstate
of the Heisenberg position operator q̂(t) with eigenvalue q
at time t, namely q̂(t) |q, t⟩ = q |q, t⟩, so if the particle is not
trapped in a bound state of V1 (q) we expect that sooner or
later at large |t|, it will escape the support of V1 (q) and
oscillates around the origin where the potential is that of
the harmonic oscillator. Of course lacking a mathematical
proof, this is only a weakly motivated assumption. Thus
24
Which we may assume compactly supported.

58
in the long run the wave functions of the in/out ground
state should become proportional to an harmonic oscillator
gaussian (this is our assumption anyway)
′ ′2
⟨q, t|in, 0⟩t→−∞ ⟨q|0⟩osc ∝ e−ωq ⟨0, out|q , t⟩t→+∞ ∝ e−ωq
2 /2 /2
, (2.152)

One may also note that the path integral is dominated by the
classical path qc (t), which for |t| → ∞ when the source J(t) is
switched off will only feel the potential and perform small
oscillations around the origin.
Now in Eq. (2.150), instead of doing the path integral from
′ ′
q to q followed by the integration over q and q , we may
as well integrate over all unconstrained path over the real
line with finite but unfixed boundary conditions. Thus we
obtain the path integral
∫ ( ∫ )
ω 2
⟨0, out|in, 0⟩|J = Dq exp iI[q] + i dtJq − (q (+∞) + q (−∞))
2
R1 2
where the domain of integration is now the set of all path of
the real line R1 into configuration space, q(t) : (−∞, ∞) → M ,
and the measure Dq includes the integral over q(±∞). Using
the trick25
∫ ∞
2 2
q (+∞) + q (−∞) = lim ϵ q 2 (t)e−ϵ|t| dt
ϵ→0 −∞

we arrive at

⟨0, out|in, 0⟩|J = (2.153)


∫ [ ∫ ( ) ∫ ]
1 2
= Dq exp iI0 [q] − i dt (ω − iϵ)q + V1 (q) + i dtJq
2
R1 2 R1

where I0 [q] is the free action and the limit ϵ → 0 is left


understood. There is an alternative way to look at a this
result. Inserting a complete set of energy eigenstates in
the transition amplitude, replacing the Hamiltonian H with
H(1 − iϵ) and taking the limits we get
′ ′ ′ ′
lim ⟨q , t |q, t⟩J = ⟨q , t |0⟩J ⟨0|q⟩
t→−∞
∫∞ ∫∞ ∫∞
In fact, ϵ −∞ f (t)e−ϵ|t| dt = ϵ 0 (f (t) + f (−t))e−ϵt dt = 0 (f (t/ϵ) + f (−t/ϵ))e−t dt;
25

hence taking the limit ϵ → 0 we get the desired result.

59
because for all positive energy states eiEn t(1−iϵ) → 0 as t → −∞,
save for the vacuum. Similarly, since e−iEn t(1−iϵ) → 0 as t → ∞,
we get
′ ′ ′
lim

⟨q , t |0⟩J ⟨0|q⟩ = ⟨q |0⟩ ⟨0|0⟩J ⟨0|q⟩
t →∞

′ ′
Thus multiplying ⟨q , t |q, t⟩J by the wave functions and inte-
grating we return to Eq. (2.150) in which the Hamiltonian
is multiplied by a factor 1−iϵ. If as in the above treatment
we split the potential into the harmonic oscillator plus a
correction V1 (q), it can be easily seen that this is equiv-
alent to the prescription ω 2 → ω 2 − iϵ, as was given above
(see,4 §6).
When V1 (q) = 0 the quadratic functional integral can be
done exactly. The action is
∫ [ ( ) ]
1 d2
I[q] = q − 2 − ω + iϵ) q + Jq dt
2
(2.154)
2 dt
Define the propagator as
( )−1
d2
∆(t − s) = − 2 − ω + iϵ
2
(t, s) (2.155)
dt
Then the convolution

∆ ⋆ J(t) = ds∆(t − s)J(s)

is a particular solution of the classical equation of motion


q̈ + ω 2 q = J(t); now shifting the integration variable from q(t)
to q(t) − ∆ ⋆ J(t), and using the translation invariance of the
measure, we obtain
( )
i
⟨0, out|in, 0⟩|J = ⟨0|0⟩|J=0 exp J ⋆∆⋆J (2.156)
2
where
∫ ∫ ( )
1 d2
⟨0|0⟩|J=0 = Dq exp i q − 2 − ω + iϵ q dt
2
(2.157)
2 dt
Of course this should be unity, so the measure will be
assumed appropriately normalized. Alternatively, one can
divide so that Z[J] = ⟨0|0⟩|J / ⟨0|0⟩|J=0 .

60
It remains to compute the propagator. A simple computa-
tion in frequency space gives

1 ′ 1 ′
−iω (t−s)
∆(t − s) = dω ′ 2 e (2.158)
2π ω − ω 2 + iϵ
Computing the integral by the method of residues we get
e−iω(t−s) eiω(t−s)
∆(t − s) = θ(t − s) + θ(s − t) (2.159)
2iω 2iω
We mention here that in the scattering case, the S-matrix
elements are computed by taking the limits using the wave
functions of the scattering in/out states

⟨β, out|α, in⟩ =



′ ′ ′
= lim ⟨β, out|q , t⟩ ⟨q , t|q, −t⟩ ⟨q, −t|α, in⟩ dq dq (2.160)
t→∞

Now to make sense of scattering theory one needs to assume


that the wave functions at large times can be computed as if
the system were free, because the classical path dominating
the path integral are asymptotically represented by uniform
motions. We shall leave this interesting topics to the vast
literature, since in quantum field theory one is largely
interested in vacuum-vacuum transitions with a source, as
this is the generating functional of the Green functions.
According to the formalism of LSZ, these functions have poles
on the mass shell whose residues are directly connected to
the scattering amplitude.

2.8 Transition to quantum field theory


After this long detour we come back to field theory. Let
us consider the most popular field theories consisting of
scalar, vector and spinor fields. Among the countless ap-
plications of the functional integral we shall only provide
few significant highlights on the use of functional in-
tegration in field theory, devoted mainly to perturbation
theory. Others will follow during the exposition of gauge
field theories and the effective action formalism.

61
2.8.1 Scalar fields
Boldly extending the definition given for N degrees of
freedom, qi (t), to a continuum q⃗x (t) = ϕ(t, ⃗x) = ϕ(x), which is
now our quantum scalar field, we obtain the generating
functional (2.13), formerly denoted by Z[J]
∫ [ ∫ ]
1
⟨0|0⟩J = exp i ( φ(□ − m + iϵ)φ − V (φ) + Jφ)d x Dφ (2.161)
2 4
2

where Dφ ∝ x dφ(x) and the ϵ-term is the result of selecting
the vacuum to vacuum transition amplitude. It gives the
inverse of the operator □−m2 +iϵ with the Feynman’s boundary
conditions that at large positive times positive energy
modes propagate forward in time, while at negative large
times positive energy modes propagate backward in time. The
proportionality factor is the normalization factor needed
to insure that Z[0] = 1, namely
∏ ∫ [ ∫ ]
i
Dφ = dφ(x)/ exp (φ(□ − m + iϵ)φ − V (φ))d x Dφ (2.162)
2 4

x
2
What have we to do with formulas like this? Recall that
functional differentiation of Z[J] at J = 0 give the chrono-
logical correlation functions, and that these are directly
connected with the scattering operator via the Lehmann-
Zymanzik-Zimmermann reduction formulas. But for general,
even renormalizable, V (ϕ) the integral will not be Gaussian
and must be computed by perturbation theory. One way is the
use of the Gell-mann Low formula, which can be derived from
the path integral in few lines. Simply write

⟨0|T ϕ(1) · · · ϕ(n)|0⟩ = Dφ φ(1) · · · φ(n) (2.163)
[ ∫ ]
i
× exp (φ(□ − m + iϵ)φ − V (φ))d x
2 4
2
∫ ∫
[ ∫ ]
i d4 xV (φ) i
= φ(1) · · · φ(n)e exp (φ(□ − m + iϵ)φ)d x Dφ
2 4
2
[ ∫ 4 ]
⟨0F |T ϕ(1) · · · ϕ(n)ei d x V (ϕ) |0F ⟩
= ∫ (2.164)
⟨0F |T ei d xV (ϕ) |0F ⟩
4

where in the last line the fields are free, the Fock vacuum
|0F ⟩ of the free theory appears because the integral with the

62
free action defines matrix elements in the Fock space and
the denominator is the normalization of the path integral
measure. From this one can compute the chronological cor-
relation functions of the full theory in terms of the known
functions of the free theory, just expanding the exponen-
tial in this formula and making repeated use of the Wick
theorem. But this uses the path integral only indirectly.
A more powerful technique makes use of the identity
( ) ∫ [ ∫ ]
δ i
F −i Z0 [J] = exp (φ(□ − m + iϵ)φ + Jφ)d x F (φ)Dφ
2 4
δJ(x) 2

where Z0 [J] is the free generating functional. Thus from the


path integral we can write
[ ∫ ( ) ]
δ
Z[J] = exp −i V −i 4
dx (2.165)
δJ(x)
∫ [ ∫ ]
× exp i (φ(□ − m + iϵ)φ + iJφ)d x Dφ
2 4

Now expanding the path integral in powers of J we are left


with the task of computing a lot functional derivatives
of expressions of the form (we are using the shorthands
J(k) = J(xk ), etc.)
[∫ ( ) ]v ∫ ∏
(−i)v δ 1
V −i 4
dz J(1) · · · J(2p) ⟨0|T [ϕ(1) · · · ϕ(2p)]|0⟩ d4 xj
v! δJ(z) (2p)! j

in which the free Green functions appears. The v-order


derivative is a sum of terms (by Leibnitz rule) each cor-
responding to a certain Feynman diagram with v vertices and
less than 2p external lines26 . Alternatively, the functional
integral in Eq. (2.165) is that of the free theory and gives
Z0 [J], the field theory analogue of the harmonic oscillator
result (2.156), namely,
[ ∫ ]
i
Z0 [J] = exp − J(x)∆F (x − y)J(y)d xd y
4 4
(2.166)
2
where the Feynman propagator reads

1 eip·x
∆F (x) = d4 p (2.167)
(2π)4 p2 + m2 − iϵ
26
For example, for V ∝ ϕ4 each diagram has e = 2p − 4v external lines.

63
Now expanding (2.165) in powers of the potential energy,
which contains the coupling parameters, and performing the
functional derivatives one generates the usual perturbative
expansion. For the interaction λϕ3 , for example, one would
start perturbation theory by writing the expansion
[∫ ( )3 ]V
∑∞
(−i)V λV δ
Z[J] = VV!
−i d4 z (2.168)
V =0
6 δJ(z)
∑∞ [∫ ]P
(−i)P
× J(x)∆F (x − y)J(y)d xd y
4 4

P =0
2P P !

Fix V and P ; after taking V third order functional deriva-


tives there remain a sum of terms each with E = 2P − 3V
sources (external legs) and coefficients (−i)P −2V = (−i)E+V −P .
We may identify each term with a diagram, so the diagram-
matic expansion and the Feynman’s rules follows from the
rules of functional differentiation. We leave it to the
reader to complete the proof27 . Of course ultraviolet di-
vergences are an issue here, and are to be treated by means
of renormalization theory.
The formalism can be extended straightforwardly to scalar
multiplets as long as there is no spontaneous symmetry break-
ing. For example, the path integral for the complex scalar
field (a doublet of real scalar fields, if we like) reads
∫ [ ∫ ]
⟨0|0⟩J = exp i (φ̄(□ − m + iϵ)φ − V (φ̄φ) + Jφ + J φ̄)d x DφDφ̄
2 ¯ 4

2.8.2 The massive vector field


We noticed that for certain theories the classical La-
grangian is not the correct one to be used in the path
integral (hence to be used to quantize such theories), but
that in these cases one can do better starting from the
Hamiltonian form of the path integral. One interesting
case is the theory of massive vector field, say ϕµ , with
the assumed Proca Lagrangian density
1 1
L = − Fµν F µν − m2 ϕµ ϕµ + Jµ ϕµ , Fµν = ∂µ ϕν − ∂ν ϕµ (2.169)
4 2
27
See, for example, Srednicki’s textbook Quantum Field Theory or Wein-
berg’s The Quantum Theory of Fields, Vol. 1.

64
where Jµ is either an external current or a function of
other fields, but not ϕµ . This Lagrangian is chosen because
it leads automatically to the transversality condition
∂µ ϕµ = 0 (2.170)
which has the effect of eliminating the scalar (spin zero)
particle from the spectrum, leaving one with a pure massive
spin-one particle. Our aim will be to obtain a manifestly
Lorentz invariant path integral.
This theory has two second class constraints (latin indices
i, j, k, l, . . . will take spatial values in {1, 2, 3}), namely

χ1 = π 0 = 0, χ2 = ∂j π j − m2 ϕ0 + J 0 = 0

where the momentum conjugate to ϕj is


π j = −F 0j = ∂0 ϕj + ∂j ϕ0 (2.171)
The canonical Poisson brackets are
{χ1 (t, ⃗x), χ2 (t, ⃗y )} = m2 δ 3 (⃗x − ⃗y ), (2.172)
confirming the second class nature of the constraints. The
dynamical variables can therefore be taken as the three
components vector ϕ ⃗ = (ϕ1 , ϕ2 , ϕ3 ) and its (constrained) con-
⃗ 1 2 3
∫ 3 momentum Π = (π , π , π ). The canonical Hamiltonian,
jugate
H = d x (Π · ϕ̇ − L ) since π0 = 0, is easily worked out and
reads
∫( )
1 j j 1 1 2 j j
H = π π + ϕ ∂j π + Fik Fik + m (ϕ ϕ − ϕ ϕ ) d3 x (2.173)
0 j 0 0
2 4 2

( j )
− Jj ϕ − J 0 ϕ0 d3 x

where ϕ0 is obtained by solving the constraint equation


χ2 = 0, or
1
ϕ0 = (∂j π j + J 0 ) (2.174)
m2
We are here shortening a bit the more pedantic procedure
∫ a
discussed in Sec.(2.4.1) of adding to H a term u χa d3 x
and integrating in the path integral over all phase space
variables including Dπ0 and Dϕ0 , with insertion of the

65
functionals δ[χa ] to enforce the constraints . But these
integrations just force π0 = 0 and ϕ0 to be a solution of the
secondary constraint, which is what we have done.
Finally omitting the inconsequential determinant factor
det[m2 δ 3 (⃗x − ⃗y )] of Eq. (2.172), the path integral takes the
form (cf. Eq. (2.92))
∫ [ ∫ ]
Dϕ Dπi exp i (πj ϕ̇ − H)dt
j j
(2.175)

The part of the action which is quadratic in the momentum


variables is
∫ ( )
1 1 1
I2 = πj ϕ˙j − πj πj − 2
(∂k πk ) − 2 J ∂j π d3 x dt
2 0 j
(2.176)
2 2m m
As with any Gaussian (or Fresnel) integral, the integra-
tion over π with this action I2 gives the inconsequential
determinant factor28 det (δik − m−2 ∂i ∂k ), times the value of the
integrand evaluated at the stationary point of I2 , namely at
the unique solution of the elliptic differential equation

(δij − m−2 ∂i ∂j )πj = ϕ̇i + m−2 ∂i J 0 (2.177)

This can be worked out29 , leaving us with the Lagrangian


path integral
∫ ∏ ∫
Dϕj (. . . ) exp i L d4 x (2.178)
j

where the Lagrangian reads


1 [ ] 1 1
L = ϕ̇i δij − ∂i (−∇2 + m2 )−1 ⋆ ∂j ϕ̇j − m2 ϕi ϕi − Fik F ik
2 2 4
1
+ Jj ϕj − J 0 (−∇2 + m2 )−1 ⋆ J 0 (2.179)
2
It appears to be non local and not manifestly Lorentz in-
variant. The path integral is in fact Lorentz invariant,
but in a very secretive way and perturbation theory with
28
We recall that field independent multiplicative factors can be sorted
out from the functional integral because they give no contribution to
amplitudes or expectation values.
29
The inverse of the operator on the left hand side of (2.177) exists and
is given by the integro-differential operator Cij = δij + ∂i (−∇2 + m2 )−1 ⋆ ∂j .

66
this Lagrangian has to be checked against Lorentz invariance
step by step. The last term is just the Yukawa interaction
between “charges” with density J 0 , since it is well known
that
1
(−∇2 + m2 )−1 (⃗x, ⃗y ) = exp(−m|⃗x − ⃗y |)
4π|⃗x − ⃗y |

However, the lack of locality and Lorentz invariance is only


apparent and can be eliminated, for indeed L in (2.179) is
just the initial covariant Lagrangian in which ϕ0 has been
eliminated by solving the constraint equation

−∇2 ϕ0 + m2 ϕ0 = J 0 + ∂k ϕ̇k (2.180)

which gives ϕ0 as a non local functional of the other fields.


This last observation is the key to recover a covariant for-
malism. Let us isolate the part of the covariant Lagrangian
(2.169) depending on ϕ0 . It is (the second line is valid up
to a total divergence)
1
L = (∂j ϕ0 ∂j ϕ0 + m2 ϕ0 ϕ0 ) + ϕ̇j ∂j ϕ0 + J0 ϕ0 + . . . (2.181)
2
1 0
= ϕ (−∇2 + m2 )ϕ0 + (J0 − ∇ · ϕ̇)ϕ0 + · · · (2.182)
2
Note the absence of time derivatives of ϕ0 and the apparently
wrong sign of the mass term which resembles a imaginary
frequency oscillator. Suppose we integrate exp iI[φµ ] over
all four components of the field. The integral
∫ [ ∫ ( ) ]
1 0
0
Dϕ exp i ϕ (−∇ + m )ϕ + (J0 − ∇ · ϕ̇)ϕ d x
2 2 0 0 4
2

is Gaussian, so the result is proportional to the integrand


computed in the stationary point of the exponent, that is
for the solution of the elliptic equation (2.180). But
this is also the solution of the constraint equation, so
that the path integration over the remaining variables ϕj is
just the same as it was obtained from the phase space path
integral. Therefore, in the end, we may use the original
local and covariant Lagrangian, with the integration measure
involving all four components ϕµ of the vector field, that

67
is we have
∫ ∏ [ ∫ ]
µ µ 4
Z[J] = Dφ exp iI[φ] + i Jµ φ d x (2.183)
µ

and
∫ ( )
1 1 2
I[φ] = − Fµν F + (m − iϵ)ϕµ ϕ d4 x
µν µ
4 2

Now the formalism appears to be fully local and Lorentz in-


variant. However this result could have hardly been guessed
before the Hamiltonian analysis had been properly under-
stood. The quadratic part of the covariant action is

1 ( )
I=− ϕµ η µν (□ − m2 + iϵ) − ∂ µ ∂ ν ϕν d4 x (2.184)
2
so this formulation automatically gives the covariant prop-
agator
∫ (
1 pµ pν ) eip·(x−y)
i ⟨0|T [ϕµ (x)ϕν (y)]|0⟩ = ηµν + 2 d4 p (2.185)
(2π)4 m p + m − iϵ
2 2

The generating functional (or the partition function in


Euclidean field theory) is obtained by the usual trick of
completing the square

i
Z[J] = exp J˜µ (k)∆µν (k)J˜ν (−k)d4 k/(2π)4 (2.186)
2
Note that for a conserved J µ one has k µ J˜µ = 0, so we can write
∫ ˜µ
i J (k)ηµν J˜ν (−k) d4 k
Z[J] = exp (2.187)
2 k 2 + m2 − iϵ (2π)4
This means that in perturbation theory one can use the
simpler propagator

˜ µν = ηµν
∆ .
p2 + m2 − iϵ

which has a well defined zero mass limit and the standard UV
behavior of renormalizable propagators. However, it also
propagates a particle with negative metric.
The naïve propagator i ⟨0|T [ϕµ (x)ϕν (y)]|0⟩ provided by canonical

68
quantization, instead of that provided by path integration,
exhibits the additional non covariant term

· · · + m−2 δµ0 δν0 δ 4 (x − y)

The operator time ordering always is defined up to a distri-


bution concentrated at coincident points, and the new term
should be harmless. Actually, in perturbation theory it is
cancelled by another non covariant∫ term in the Hamiltonian,
the term coming from the coupling d3 xϕ0 J 0 after ϕ0 has been
eliminated using its equation of motion30 .
Note that the zero mass limit of this theory cannot be taken
trivially: either the propagator is singular or there is a
propagating ghost31 . The problem has to do with gauge in-
variance and will be handled exhaustively in the context of
gauge theories. However it can be understood very easily:
the massless action to be integrated is

1
I=− Aµ (η µν □ − ∂ µ ∂ ν ) Aν d4 x (2.188)
2
but the written operator is not invertible. It has a lot of
zero modes, all of the form Aµ = ∂µ ϕ for arbitrary ϕ, so the
integral over such field configurations diverges badly. The
problem is usually evaded by breaking the gauge invariance,
for example by adding a term to the action

α
− (∂µ Aµ )2 d4 x
2

and showing that matrix elements of gauge-invariant opera-


tors are actually independent of the added term, or of the
gauge parameter α. Surprisingly enough, this can also be
done to a massive vector field if coupled to a conserved
current, a fact which is known as the Stuckelberg trick.
It gives a renormalizable theory of massive vectors which
admits a non singular zero mass limit, but unfortunately it
cannot be applied to non abelian massive vectors.
30
See Weinberg’s Book somewhere in Vol.(1).
31
At the Lagrangian level it seems to give the massless vector field
of quantum electrodynamics.

69
2.8.3 Spinor fields
Although the path integral for fermionic fields can be de-
duced from an abstract dynamics involving Grassmann (i.e.
anti-commuting) variables, we take a shorter route and sim-
ply use the Fourier representation (2.29) of the generating
functional
Z[η, η̄] = (2.189)
∫ [ ∫ ∫ ]
= DψDψ̄ exp −i ψ̄(∂/ + m − iϵ)ψd x + i (η̄ψ + ψ̄η)d x
4 4

and additionally we interpret the functional integral as


the continuum version of the Berezin integral described in
Appendix [A]. The measure is therefore

DψDψ̄ = dψα (x)dψ̄α (x) (2.190)
α,x

Eq. (2.189) is the functional integral representation of


the generating functional of time ordered products of the
free spinor field. It can be computed by the ”shifting
variables method”, namely shifting ψ = χ + D /−1 ⋆ η, ψ̄ = χ̄ + η̄ ⋆ D
/−1
in the integral and using the translational invariance of
the measure. Here D / = ∂/ + m − iϵ is the Dirac operator and a
star means performing a convolution. Then in few more steps
one obtains the result
( ∫ )
−1 4 4
Z[η, η̄] = Z[0, 0] exp i η̄(x)D / (x, y)η(y)d xd y (2.191)

where
∫ [ ∫ ]
Z[0, 0] = DψDψ̄ exp −i ψ̄(∂/ + m − iϵ)ψ d x = det iD
4
/

The fermion propagator is well known



−1 1 eip·(x−y)
D
/ (x, y) = (−∂/ + m) d4 p (2.192)
(2π)4 p2 + m2 − iϵ
From the functional integral we then obtain the familiar
equation (note the use of left derivatives)
[ ]
δL2 Z
/−1 ]αβ (x − y)
= ⟨0|T ψα (x)ψ̄β (y)|0⟩ = i[D (2.193)
δηβ (y)δ η̄α (x) |η,η̄=0

70
Apart from the nuisance to distinguish between left and right
derivatives and all matters related to the anti-commuting
nature of the fermion fields, most manipulations with the
path integral are straightforward and pattern similar cal-
culations with bosonic functional integrals.
An instructive example is the external field problem.
Suppose the fermions are disturbed by an external classical
electromagnetic field. The path integral becomes now a
functional of the external field A

Z[A, η, η̄] = (2.194)


∫ [ ∫ ∫ ]
= DψDψ̄ exp −i ψ̄(∂/ − ieA/ + m − iϵ)ψd x + i (η̄ψ + ψ̄η)d x
4 4

where e is the electric charge. Putting the sources to zero,


the functional can be written in the form

Z[A] = exp iS[A] = ⟨out, 0|0, in⟩A (2.195)

so that, provided we normalized Z[0] = 1, the vacuum persis-


tent probability reads

| ⟨out, 0|0, in⟩A |2 = e−2ℑS[A] (2.196)

The functional S[A] is known as the Euler-Heisenberg effec-


tive action. Since

δS
= e−iS[A] ⟨out, 0|Jµ |0, in⟩A ≡ ⟨J µ ⟩A
δAµ

the functional S[A] describes all quantum corrections to


Maxwell equations due to vacuum polarization effects. From
(2.194) we obtain the remarkable formula
[ ]
det[∂/ − ieA/ + m − iϵ) ∂/ − ieA/ + m − iϵ
Z[A] = = det (2.197)
det(∂/ + m − iϵ) ∂/ + m − iϵ
The computation of the determinant is a non trivial task,
but a variety of methods have been devised to compute ratios
of such functional determinants. For an explicit calcula-
tion see Itzykson & Zuber, Quantum Field Theory, Ch.[5].
Besides external fields, other interactions can be intro-
duced by adding to the free action appropriate higher order

71
terms, for example the Fermi-type and Yukawa-type interac-
tions

L1 = gF (ψ̄Γ0 ψ) (ψ̄Γ1 ψ) + gY φψ̄Γ2 ψ

By far the most important are gauge interactions, which we


saw can be introduced via the minimal coupling prescription
∂µ ψ → ∇µ ψ. Of course in this case we need also to add
the action of the gauge fields and perform the functional
integral not only over the matter fields but also over the
gauge fields. We just meet a problem in doing so for the
electromagnetic field. Now we face this issue using the
present functional methods.

72
Chapter 3

Quantum gauge fields

We return to gauge theories and aim to find a path inte-


gral representation of the generating functional Z[J µ ]. Just
like electrodynamics, there are a lot of gauge field con-
figuration with the same action, so the integral over all
configuration (i.e. over all functions Aaµ ) counts gauge
equivalent functions as many times as there are elements in
the gauge group. We now introduce a method to deal with this
feature that exploit the gauge invariance of the action to
”fix a gauge”, thus avoiding the over-counting problem.

3.1 The Faddeev-Popov-De Witt path inte-


gral
The gauge group partitions the space of gauge fields into
equivalent classes, where equivalent fields in each class
are those connected by gauge transformations. These classes
are the orbits of the group action and the set of equivalent
classes is the space of orbits. According to the general
theory of systems with first class constraints the true phase
space of the theory is the gauge invariant reduced phase
space1 . This is the space containing histories over which
the path integrals of the quantum theory should be done,
in order to count each gauge field configuration only once.
For pure Yang-Mills theory the gauge generating constraints
1
As explained in Sec.2.4.2, the reduced phase space is defined as the
quotient of the constraints surface by the gauge group.

73
are
∂LY M
γ1a ≡ π0a (t, ⃗x) = =0 (3.1)
∂ Ȧ0
γ2a ≡ ∂i πia + gC abc Abi πic = 0, where πia = F0ia (3.2)
The constraints are of course of the type required by gauge
theories, namely first class, with Poisson brackets {γ1 , γ2 } =
0 and

{γ2a (x), γ2b (y)} = gf abc γ2c δ(⃗x − ⃗y )

A convenient way to describe the reduced phase space without


having to solve the constraints (which with few exceptions
is very difficult if not practically impossible to do), is by
choosing a “surface” within configuration space (the space
of all matrix valued gauge fields Aµ (x)) that intersects
every orbit only once. This can be specified by a set of
local functionals2 F a [A; x], one for each gauge field, which
define the surface via the equations
F a [A; x] = 0 (3.3)
Such a choice is called a gauge fixing procedure and F a [A; x]
is the set of the corresponding gauge fixing functionals.
Questions of regularity of the surface or differentiability
of the functionals are avoided here, hoping they can be
given in principle. For example, for SU (2) gauge fields on
S 4 it has been proved by Singer that the space of orbits is
a differentiable manifold, but unfortunately he also showed
that it is impossible to find a globally defined gauge fixing
surface for this case.
There are conditions that a good gauge fixing functional
has to satisfy. First of all the condition (3.3) must be
reached by gauge transformations alone. The second condi-
tion is that the gauge fixing really fix the gauge, that
is for x → g(x) in the gauge group and Agµ the transformed
potential
Agµ = −i∂µ gg −1 + gAµ g −1 (3.4)
2
Functionals which are ordinary functions of Aaµ (x) and the derivatives
∂µ1 · · · ∂µk Aaµ (x) up to some finite order k, like for example the axial or
Lorentz gauge, F a [A; x] = Aa3 (x) or F a [A; x] = ∂ µ Aaµ (x), respectively.

74
then there exist only one g such that F a [Ag ; x] = 0, alias,
for which the delta function δ (F a [Ag ; x]) is non zero. This
is because in the contrary case there are gauge equivalent
field configurations on the same gauge fixing surface, that
is the surface intersects a given orbit more than once. This
phenomenon actually occurs for most choices of the gauge
fixing functional, and is known as the Gribov obstruction.
The gauge equivalent configurations on the gauge fixing
surface are called Gribov copies. Actually it may also
happens that for no g will F a [Ag ; x] = 0, or that there is
no globally defined gauge fixing surface at all. Working
locally, in some suitable sense, we can at least check
whether there are also infinitesimally close Gribov copies.
A direction in field space δAaµ will be tangent to the gauge
fixing surface if

δF a [A; x] b
d4 y δAµ (y) = 0 (3.5)
δAbµ (y)
We introduce the kernel of the so called Faddeev-Popov op-
erator (as always, repeated indices are to be summed over)

δF a [A; x]
µ δ (z − y)d z
ab
M (x, y) = (∇z )cb 4 4
(3.6)
δAcµ (z)

acting on Lie-algebra valued functions ha (x) via


∑∫ ∑ ∫ δF a [A; x]
h (x) →
a ab b 4
M (x, y)h (y)d y = c (z)
(∇z )cb b 4
µ h (z)d z
b c,b
δA µ

where

∇cb cb
µ = ∂µ δ + gf
cab a
Aµ (3.7)

If the FP operator has no zero modes then the gauge variation


δAaµ = ∇µ ϵa will take Aaµ away from the surface F a = 0, namely

F a [Aµ + ∇µ ϵ; x] ̸= 0 (3.8)

and therefore there will be no “infinitesimally closed”


Gribov copies. Note that M ab (x, y) is in general the kernel
of a differential operator written in integral form, and
that the functional determinant det M will not vanish. For

75
example, in the Lorentz gauge, where F a [Aµ ; x] = ∂ µ Aaµ (x), the
Faddeev-Popov operator takes the form

abc ∂
( µb c )
h → M ab (x, y)hb (y)d4 y = ∂ µ ∇ab
a b µ a
µ h (x) = ∂ ∂µ h + gf A h
∂xµ
Suppose then that we have a good gauge fixing surface. We
can go through the canonical formalism and use the Faddeev
formula (2.97) in the present field theory context, then
after integration over the momentum variables we would get
the fundamental Faddeev-Popov-De Witt path integral formula
for the generating functional in gauge theories3
∫ ( ∫ )
ab µ a 4
Z[J] = DA δ [F [A]] det[M ] exp iI0 [A] + i Ja Aµ d x (3.9)

where DA = a,µ,x dAaµ (x) and, as usual,

1
I0 [A] = − d4 x Fµν
a
F aµν (3.10)
4
The functional Dirac delta is to be formally interpreted as

δ[F [A]] = δ (F a [Aµ ; x]) (3.11)
a,x

The operator M ab (x, y) is the gauge variation of the gauge


fixing functional (c.f. Sec. (2.4.2))

δF a [A1+ϵ ; x]
ab
M (x, y) = , A1+ϵ = Aµ + ∇µ ϵ (3.12)
δϵb (y) ϵ=0 µ

and is obviously equal to the Faddeev-Popov operator given in


Eq. (3.6). But the Hamiltonian treatment is complicated (see
the book by Faddeev & Slavnov8 for a general discussion). In
fact the Hamiltonian procedure that led us to Eq. (3.9) can
be shortened considerably by what is known as the Faddeev-
Popov trick. Let us consider the functional ∆[A] defined
by4
∫ ∏
1 = ∆[A] Dg δ[F [Ag ]], Dg = dg(x) (3.13)
G x

3
A concise and very clear presentation of the full matter can be found
in.8
4
This is known as the Faddeev-Popov determinant.

76
where Agµ is the gauge transformed potential by the group
element g and the integral is over the gauge group with a
right invariant measure
∏ ∏
Dgh = d[g(x)h(x)] = dg(x) = Dg (3.14)
x x

Such a measure always exists on any compact Lie group and


is known as the right invariant Haar measure (it is also
left invariant as a rule, Dhg = Dg). First we show that ∆[A]
is gauge invariant.
Proof:
∫ ∫
h −1
∆[A ] = Dg δ[F [(A ) ]] = Dg δ[F [Agh ]]
h g

∫G G

= D(gh) δ[F [A ]] = Dg δ[F [Ag ]] = ∆[A]−1
gh
(3.15)
G G

We then insert “1” written∫ in the form (3.13) into the


formal functional integral DA exp iI0 [A]. This is divergent
actually and should be provisionally regularized. In fact
as we noted before it is proportional to the infinite volume
of the gauge group. To extract this volume we then change
the integration variable from A to Ag : using the gauge
invariance of the action, I0 [A] = I0 [Ag ], that of the measure
DA and also of ∆[A], we end up with
∫ ∫ ∫
DA exp iI0 [A] = Dg × DA ∆[A] δ[F [A]] exp iI0 [A] (3.16)
G
The infinite volume of the gauge group has been factored out
from the functional integral, so the remaining factor must
be the integral over the space of orbits. This is known as
the Faddeev-Popov trick. It means also that the functional
integral in (3.16) is largely independent on the choice of
the gauge fixing functional, the obvious restriction being
the existence of ∆[A]. As we shall shortly see, this entails
the absence of zero modes for the Faddeev-Popov operator.
In computing matrix elements of time-ordered products
of gauge invariant operators, we insert the corresponding
classical functional into the integral and we divide by the
integral with no insertions; in this way we obtain
∫ ∫
DA exp iI0 [A]O1 [A] · · · DA ∆[A] δ[F [A]] exp iI0 [A]O1 [A] · · ·
⟨O1 · · ·⟩ = ∫ = ∫
DA exp iI0 [A] DA ∆[A] δ[F [A]] exp iI0 [A]

77
because the infinite volume of the gauge group drops out in
this ratio.
It remains to compute ∆[A]. To this aim we can restrict
the integration over Dg to a neighborhood of the identity
because the measure appears multiplied with the delta func-
tional enforcing the gauge condition F a [A; x] = 0. We suppose
that for a unique g the delta functional is non zero (for an
A satisfying F a [A; x] = 0 this is the identity, but recall the
Gribov problem). Then putting g = 1 + ϵ in (3.13) we obtain5
∫ ∫
1+ϵ
Dg δ[F [A ]] = Dϵ δ[M ab ⋆ ϵb ] ∝ det[M ab ]−1 (3.17)
G G

where M ab (x, y) is given by Eq. (3.12), because (with a concise


notation) A1+ϵ = A + ∇ϵ and expanding

F a [A + ∇ϵ; x] = F a [A; x] + M ab ⋆ ϵb = M ab ⋆ ϵb

the ⋆ denoting the convolution with the kernel M ab (x, y). The
last equality in (3.17) is the functional version of the well
known identity in finite dimensions, δ n (M · x) = | det M |−1 δ n (x).
We may therefore set ∆[A] = det[M ab ], and we recover the
original result of the Hamiltonian path integral method.
At the same time, there must not be any zero mode for M ab ,
otherwise det[M ab ] = 0, and the path integral makes no sense.
Two useful examples can be given. One is the axial gauge
fixing, Aa3 (x) = 0; the operator is
∂ 4
M ab (x, y) = δ ab δ (x − y) (3.18)
∂x3
It does not depend on A and can be sorted out from the func-
tional integral. The path integral for Yang-Mills theory
in axial gauge is then given by
∫ ∏ ( ∫ )
a µ a 4
Z[J] = DA δ(A3 (x)) exp iI0 [A] + i Ja Aµ d x (3.19)
a,x

That Eq. (3.19) is correct can be seen from the canonical


quantization of the theory in axial gauge.2 The second
5
We may absorb any Jacobian for the transformation g → ϵ into the
measure Dϵ.

78
example is the Lorenz gauge condition, ∂ µ Aaµ = 0. The Faddeev-
Popov operator is non trivial now

µ δ (x − y)
M ab (x, y) = (∂x )µ (∇x )ab 4
(3.20)
(the derivatives act on the indicated variables), so the
action on functions χb (y) is

∂ ( b c)
χ (x) →
a
µ χ = □χ + gf
M ab (x, y)χb (y)d4 y = ∂ µ ∇ab b a abc
Aµ χ
∂xµ

The path integral for Yang-Mills theory in Lorenz gauge then


takes the form
Z[J] = (3.21)
∫ ( ∫ )
[ µ ab 4 ]
= DA δ [∂ Aµ ] det ∂ ∇µ δ (x − y) exp iI0 [A] + i Ja Aµ d x
µ µ a 4

where as always

δ [∂ µ Aµ ] = δ(∂ µ Aaµ (x))
a,x

Eq. (3.21) will be one key equation in discussing quantum


gauge theories. We mention here the Coulomb gauge which was
so popular in quantum electrodynamics

∇ · Aa = 0

The quantization in this gauge can be done (a very good


introduction is the review article by Abers and Lee26 ) but
is more complicated and, unlike the axial gauge, suffers
from the Gribov problem.
Digression: it may be of interest to look again at the
Gribov problem in infinitesimal form in Lorenz gauge. Given
a gauge field in this gauge, a gauge equivalent field close
to it can be found which still obeys Lorenz gauge if we can
find a solution of the equation
abc ∂
( b c)
∂ µ ∇ab
µ ϵ = □ϵ + gf
b a
Aµ ϵ = 0 (3.22)
∂xµ
because then

Aµa = Aaµ + ∇ab
µ ϵ
b
(3.23)

79

will still obeys the gauge condition ∂ µ Aµa = 0. Thus the
ghost operator must have a zero mode (as we showed above
this is always true: the condition for the existence of
a Gribov copy in a neighborhood of a gauge field is the
existence of a zero mode of the ghost operator).
The first (yes, there are many) Gribov region is the set

Ω = {Aaµ | ∂ µ Aaµ = 0 ∧ ∂ µ ∇ab


µ > 0} (3.24)

and its boundary is the Gribov horizon. It has very nice


properties: Ω is convex, bounded in all directions in field
space, and has the crucial property that every gauge orbit
intersects it. Also, Aaµ = 0 is in Ω so that one can do pertur-
bation theory. To improve upon the original Faddeev-Popov
procedure Gribov proposed then to restrict the functional
integration over Ω. As shown by himself this can be im-
plemented in the path integral formula, and of course has
the virtue that eliminates the problem of infinitesimally
related Gribov copies. Unfortunately, further copies are
left inside Ω and a further restriction would be necessary.
However if all we are interested in is perturbation theory,
we can ignore the problem, as discussed for example in.32

3.2 The quantum Lagrangian and the ghosts


As it stands there is not much we can do with the path
integral (3.21), or with the more general (3.9). Therefore
we shall manage to put it in a more useful form. This is
not a minor point: beyond computational advantages it will
lead eventually to discover the famous BRST symmetry.
The delta functional can be easily disposed of: for exam-
ple in Lorentz gauge, we can substitute the delta functional
with the oscillatory exponential (up to a field independent
proportionality factor) for arbitrary ξ
( ∫ )
i
δ [∂ Aµ ] → exp −
µ µ b ν b 4
(∂ Aµ )(∂ Aν )d x (3.25)

Then large values of the gradient ∂ µ Aaµ will interfere de-
structively and force the integral to be concentrated around
the stationary point at ∂ µ Aaµ = 0. Inserting this in the path

80
integral we are then integrating with the new local action

1
I[A] = I0 [A] − (∂ µ Abµ )(∂ ν Abν )d4 x (3.26)

which is no more gauge invariant. Hence the new term will
still fix the gauge (up to Gribov copies) for any ξ; with a
small abuse of language we then say that a given ξ corre-
sponds to a choice of gauge, even if it does not corresponds
to a condition on the gauge potentials. It is only in the
limit as ξ → 0 that the oscillatory factor forces the Landau
gauge condition ∂ µ Aaµ = 0. One speaks generically of a family
of Rξ gauges, the R standing for renormalizable.
There is a different (and better) way to look at this. We
have noted before that the path integral (3.9) is largely
independent on the choice of the gauge fixing functional,
the only restriction being the absence of zero modes for
the Faddeev-Popov operator. So imagine to use a slightly
more general gauge condition

∂ µ Aaµ − ω a (x) = 0

and to repeat the Faddeev-Popov trick. Nothing changes and


the determinant ∆[A], in particular, does not depend on ω
despite its presence in the delta-functional. With the
integral independent on the choice of ω a , we can take the
average of the functional integral with any functional B[ω]
without changing its content, in particular with a Gaussian
( ∫ )
i
exp − a
ω ω dx a 4

That is we insert this Gaussian and integrate over Dω a .


This too will not change the functional integral6 . In this
way we get the path integral with the gauge fixed action I
but no delta-function inside.
To convert the Faddeev-Popov determinant in a further
correction to the Lagrangian, because for example we may want
to deduce the Feynman rules of Yang-Mills theory, we could
6
We may also note that large values of ω will not put the gauge fixing
surface outside the safer Gribov region of Landau gauge, because the
oscillatory integral will not allow such large values.

81
trivially exponentiate ∆[A] → exp log ∆[A], but this is not a
good idea because the logarithm is a non local functional
of A. Instead, we can take advantage of the functional
integration over anti-commuting variables and simply write
the functional integral version of Eq. (6.152), which reads
∫ [∫ ]
ab a ab b 4 4
det[M ] = DcDc̄ exp c̄ (x)M (x, y)c (y)d xd y (3.27)

where as usual DcDc̄ = a,x dca (x)dc̄a (x). The anti-commuting
a
fields must be Lorentz scalar, c is known as the ghost and
c̄a as the anti-ghost field. They are not mutually adjoints,
in particular we take them Hermitian so that the ghost
Lagrangian

Lgh (x) = −i c̄a (x)M ab (x, y)cb (y)d4 y (3.28)

is also Hermitian. In particular the anti-ghost, in spite


of the name, will not create the anti-particles of the ghost
field, which has none since it is Hermitian. In Lorenz-
Feynman-’t Hooft gauge this reads, up to a boundary term,
Lgh (x) = −ic̄a ∂ µ ∇ab b µ a a µ a abc b c
µ c = i∂ c̄ ∂µ c + ig∂ c̄ f Aµ c (3.29)
and leads to a new ghost-anti-ghost-gauge interaction
L1gh (x) = igf abc ∂ µ c̄a Abµ cc (3.30)
It corresponds to a new vertex

c
= igpµ f bca
µ

were p is the four-momentum carried by the outgoing c̄-ghost


line. By contrast, in axial gauge the ghost Lagrangian is
Lgh (x) = ic̄a ∂3 ca (3.31)

82
so the ghosts are decoupled from the other fields, and can
be ignored.
Turning back to the covariant Feynman gauge, the Faddeev-
Popov determinant written as a functional integral provides
a further correction to the Yang-Mills actions in the form
of the ghost Lagrangian, leading to the final local, fully
covariant action
∫ ∫
1
I[A, c̄, c] = I0 [A] − (∂ Aµ )(∂ Aν )d x + i ∂ µ c̄a ∇ab
µ b ν b 4 b 4
µ c d x (3.32)

were I0 [A] is the gauge invariant part. As long as the
operators O1 , . . . , On are gauge invariants we have (operators
to the left hand side, c-number functionals to the right)
⟨0|T (O1 (x1 ) · · · On (xn ))|0⟩ = (3.33)

= N DADcDc̄ O1 (x1 ) · · · On (xn ) exp iI[A, c, c̄]

where

−1
N = DADcDc̄ exp iI[A, c, c̄]

is the usual normalization.


Let us look at the propagator of the gauge field. With the
gauge fixing Lagrangian the quadratic part of I takes the
form

1 ( )
I2 = Aaµ η µν □ − (1 − ξ −1 )∂ µ ∂ ν − iϵ Aaν d4 x (3.34)
2

1 ( ) d4 p
= − Ãaµ (−p) η µν p2 − (1 − ξ −1 )pµ pν − iϵ Ãaν (p) (3.35)
2 (2π)4
The inverse of the matrix in the bracket is of course the
propagator, which reads
( )
δ ab pµ pν
˜
∆ab;µν (p) = 2 ηµν − (1 − ξ) 2 (3.36)
p − iϵ p − iϵ
As expected, the choice ξ = 0 corresponds to Landau gauge,
which is transverse, and ξ = 1 is known as the Feynman-
t’Hooft gauge. The corresponding propagator was early used
in our discussion of fermion anti-fermion scattering into
gluons, so it was well motivated after all!

83
Now suppose the vector field has a mass term m2A A2 /2, so
there is no gauge symmetry and we can forget about the gauge
fixing Lagrangian. The propagator reads
( )
1 p µ p ν
˜ µν (p) =
∆ ηµν − 2 (3.37)
p2 + m2A − iϵ mA
and we see it conveys a bad ultraviolet behavior. We showed
previously that the coupling to a conserved current allowed
one to use instead the renormalizable propagator

˜ µν (p) = ηµν

p2 + m2A − iϵ

More generally, we can insert back the gauge fixing La-


grangian. Now the propagator becomes
( )
1 p µ p ν
˜ µν (p) =
∆ ηµν − (1 − ξ) 2 (3.38)
p2 + m2A − iϵ p + ξm2A
which has a good ultraviolet behavior. The original propa-
gator is recovered in the limit ξ → ∞, known as the unitary
gauge. On closer analysis however, renormalizability does
not stand up for massive Yang-Mills theories, a more so-
phisticated approach is needed.

3.3 BRST symmetry and cohomology


There is a curious symmetry left over by the gauge fixing
procedure of the functional integral. It is best displayed
if we write it in terms of Nakanishi-Lautrup fields
∫ [ ∫ ∫ ]
ξ a a 4
Z = DADcDc̄DB exp iI[A, c, c̄] − i B ∂ Aµ + i
a µ a
B B dx
2
where I[A, c, c̄] is the action with no gauge fixing term ξ −1 (∂A)2 .
The integral over B a is Gaussian so it is computed as usual
by substituting for B the stationary point, which occurs at
B a = ξ −1 ∂ µ Aaµ (3.39)
This procedure gives back the original functional integral.
Let us recap the Lagrangian (omitting color indices for
simplicity)
1 ξ
L = − Fµν F µν + Lm (ψf , ∇µ ψf ) + i∂ µ c̄∇µ c + ∂ µ BAµ + BB (3.40)
4 2
84
Consider (somewhat ad hoc) the following transformations
involving gauge, ghost, anti-ghost, B and matter fields

δB Aaµ = ∇µ ca (3.41a)

δB ψm = ig ca [Ta ]mn ψn (3.41b)
n
g
δB ca = − f abc cb cc (3.41c)
2
δB c̄ = iB a
a
(3.41d)
δB B a = 0 (3.41e)

These are the BRST7 transformations. We complete the defi-


nition by extending the operator to the space of polynomials
of the basic fields by linearity and the appropriate gen-
eralization of the Leibnitz rule
δB (Φ1 Φ2 ) = (δB Φ1 )Φ2 + (−1)σ Φ1 δB Φ2 (3.42)
where σ is the grading of Φ1 , namely σ = 0 (σ = 1) for a
commuting (anti-commuting) field, respectively.
To show the invariance of the action under (3.41) we note
that the first two lines in the list (3.41) are just the
ordinary gauge transformations in which the role of the
gauge parameter is played by the ghost field ca . Therefore
the gauge invariant part of the action (the first two terms
in (3.40)) is certainly invariant under BRST. The last term
is trivially invariant. The term iδB (∂ µ c̄)∇µ c is cancelled by
the term ∂ µ BδB Aµ . It remains to show that i∂ µ c̄δB (∇µ c) also
vanishes. But this is equal to i∂ µ c̄δB δB Aµ and now we are
going to show that the BRST transformation is nilpotent:
for whatever field Φj in the theory, δB2 Φj = 0. Then the
claimed invariance will follows.
Obviously δB2 B a = δB2 c̄a = 0 follows directly from the BRST
algebra. Consider now matter fields (in matrix notation).
We have
g2
δB2 ψ = −i f abc cb cc Ta ψ + g 2 ca cb Ta Tb ψ (3.43)
2
Note the sign of the second term: it comes because δB con-
verts bosonic (or even) into fermionic (or odd)variables and
7
The acronym stands for Becchi, Rouet, Stora and Tyutin, who found
the symmetry independently from the first three.

85
vice versa, so we have to complete its definition by re-
quiring that δB should commute (anti-commutes) with bosonic
(fermionic) variables, respectively. This is actually what
Eq. (3.42) says.
Now to the point: since ca cb = −cb ca we can put (one-half)
the commutator [Ta , Tb ] = ifabc Tc in place of Ta Tb . The two terms
then cancel leaving δB2 ψ = 0.
Now consider the ghost. We have after a little algebra
(all repeated indices are summed over as usual), since cb cd ce
is totally anti-symmetric,
g 2 acb cde b d e
δB2 ca = f f cc c (3.44)
2
g 2 ( acb cde )
= f f + f ace f cbd + f acd f ceb cb cd ce
6
But this vanishes by the Jacobi identity. Finally
( g b cde d e )
δB2 Aaµ = δB ∇µ ca = −gf abc ∂µ cb cc + gf abc ∇µ cb cc − A f c c
( ) 2 µ
1 b cde d e
= g f2 abc
f Aµ c c − Aµ f c c
bde d e c
(3.45)
2
We can anti-symmetrize the first term in the last line
(since ce cc = −cc ce ) and rearrange the name and position of
the indices to get at last
( )
δB2 Aaµ = g 2 f abc f bde + f abe f bcd + f abd f bec Adµ cc ce = 0 (3.46)
again by the Jacobi identity. So with regards to the basic
fields the nilpotent character of the symmetry has been
shown. On the other hand, for general (even composite)
fields for which we already know that δB2 Φ1,2 = 0, we have
Hence since δB Φ1 has opposite grading (or statistic) of Φ1 ,
δB2 (Φ1 Φ2 ) = −(−1)σ δB Φ1 δB Φ2 + (−1)σ δB Φ1 δB Φ2 = 0 (3.47)
We can conclude that δB , acting on the polynomial algebra
generated by the local fields (3.41), will be nilpotent.
As with any symmetry that is not spontaneously broken,
in the operator version of the theory there will be a self-
adjoint and conserved BRST charge (in the best of all pos-
sible worlds), say QB , which generates the symmetry in the
sense that
eiθQB Φj e−iθQB = [iθQB , Φj ] = θδB Φj (3.48)

86
where θ is an anti-commuting parameter of the field algebra,
or equivalently
i[QB , Φj ]± = δB Φj (3.49)
But then
δB2 Φj = [iQB , [iQB , Φj ]± ]∓ = −[Q2B , Φj ] = 0 (3.50)
so either Q2B ∝ 1 or Q2B = 0; but the generator of a symmetry
must annihilates the vacuum, so in fact8
Q2B = 0 (3.51)
The conserved current has the form
JBµ = B a ∇µ ca − ∂ µ B a ca + 2i g∂ µ c̄a f abc cb cc − ∂ν (F aνµ ca ) (3.52)
and formally

QB = JB0 d3 x (3.53)

The ghost Lagrangian also has the continuous symmetry,


c → e−λ c, c̄ → eλ c̄ (3.54)
with λ real. The conserved current is found easily from the
Noether’s theorem
J µ = −∂ µ c̄a ca + c̄a ∇µ ca (3.55)

and is Hermitian. The charge, Qg = J 0 d3 x leads to the ghost
quantum number: it takes the value (n − m) on a state with
n ghosts and m anti-ghosts, so this is the ghost quantum
number of the state.
Digression: One may note that
JB0 = ∂ν π aν ca + B a ∇0 ca + · · · (3.56)
where πµa = −Fµa0 is the momentum canonically conjugate to the
gauge field.
Coming back to the Lagrangian (3.40), and using (3.41),
we get the structure
1
L = − Fµν F µν + Lm (ψf , ∇µ ψf ) + δB K (3.57)
4
8
One can also note that QB has non zero ghost number, so it cannot
be proportional to the identity.

87
where K is a functional with ghost number qg = −1
i a a
K = −i∂ µ c̄a Aaµ − ξ c̄ B (3.58)
2
As with any nilpotent operator there is an associated coho-
mology space. In the space of functionals of gauge, ghost
and auxiliary fields, a closed functional G is defined as
one in the kernel of δB , i.e. δB G = 0, while an exact func-
tional is an element in the image: S = δB W for some W.
Obviously since δB2 = 0

Im[δB ] ⊂ Ker[δB ] (3.59)

The quotient space is the BRST cohomology of the symmetry


map, namely the BRST-invariant functionals modulo the exact
ones,

H[δB ] = Ker[δB ]/Im[δB ] (3.60)

Eq. (3.57) says then that the gauge invariant action from
which we started is an element of this cohomology, and this
is true for any gauge invariant functional with ghost number
zero. It can be actually proved that any BRST invariant
functional of gauge, matter, ghost and auxiliary fields
as needed, which has vanishing ghost number is the sum
of a functional of gauge and matter fields alone, having
vanishing ghost number, plus an exact one. Therefore the
cohomology is the space of gauge invariant functionals. Let
us give a proof.
To simplify formulas let us use the condensed DeWitt
notation: a BRST variation of any field is written as
j
δΦj = cA δA Φj = cA RA (Φ) (3.61)

where the indices A = (a, x), j (the field component) subsume


all discrete and continuous labels presents, with the un-
derstanding that the sum is discrete for discrete indices
and an integral for continuous ones. For example, if Φj is
the gauge field Abµ (y) then j = (b, µ, y) and

∑∫
c A
δA Aaµ (y) = [−cb (x)∂xµ δ ab δ 4 (x−y)+f abc cc (x)Abµ (x)δ 4 (x−y)] d4 x = ∇µ ca (y)
b

88
We assume the gauge algebra
C
[δA , δB ] = FAB δC (3.62)
where the F ’s are not necessarily constants (field indepen-
dent). We introduce the Slavnov operator
δ 1 A B C δ δ
S = cA δA ΦB B − FBC c c A
+ iB A A (3.63)
δΦ 2 δc δc̄
A simple calculation show that S 2 = 0 if and only if the
generalized Jacobi identity hold
E B E
F[AB FCD] + δ[A FCD] =0 (3.64)
where the square brackets denote total anti-symmetrization
with respect to the free indices, or taking the cyclic sum.
Finally let us introduce the Hodge operator
δ
H = ic̄A A (3.65)
δB
Then we get a dilatation operator by taking the anti-
commutator with S
δ δ
{S, H} = −c̄A A − B A A (3.66)
δc̄ δB
The Slavnov operator does not change the total number of B
and c̄ fields, so if we expand a BRST invariant functional
as a sum of term each containing a definite number N of B
and c̄ fields

I= IN , SI = 0 (3.67)
N ≥0

then each term must be separately invariant: SIN = 0. But


for what we have seen

{S, H}IN = SHIN = −N IN

therefore
IN = −N −1 SHIN (3.68)
and it follows that
∑ 1
I = I0 + SK, K=− HIN (3.69)
N ≥1
N
and we see that indeed I is co-homologous to a gauge in-
variant functional I0 of ghost number zero, containing no B
and c̄ fields, and therefore no c fields either.

89
3.4 Physical states and ghost decoupling
The quantum version of the previous important results can
be obtained by the following argument. A variation of the
action due to a variation of the gauge fixing functional
should not change any physical result because these must be
gauge invariant if the theory has to make any sense. From
(3.57) this change doesn’t affect the cohomology because it
is of the form δI = δB δK. Now what is the variation of a
physical matrix element induced by δI? We have first of all
∫ [ ∫ ]
⟨ϕ, out|χ, in⟩ ∝ DADψm DcDc̄DB exp iI[A, ψm ] + i δB Kd x
4

where the proportionality factor is the product of the wave


functionals of the in- and out-states9 and the ψm are the
matter fermion and scalar fields (everything else but the
gauge fields). Hence
δ ⟨ϕ, out|χ, in⟩ = i ⟨ϕ, out|δI|χ, in⟩ = i ⟨ϕ, out|δB δK|χ, in⟩ = 0 (3.70)
where the last equality is what we must require. By the
way, the first equality is a version of the famous Schwinger
action principle, in which δB δK is of course the operator
version of the corresponding functional appearing in the
functional integral. Then using the BRST charge we have
δ ⟨ϕ, out|χ, in⟩ = − ⟨ϕ, out|{QB , δK}|χ, in⟩ (3.71)
and this can vanish for arbitrary states if and only if

QB |χ, in⟩ = ⟨ϕ, out| QB = 0.

This is then the requested condition: physical states must


be BRST-closed:

QB |phys⟩ = 0

Trivial solutions of this equation are states which are


BRST-exact, namely of the form |ψ⟩ = QB |ξ⟩, because QB is
nilpotent. They have also zero norm by the same reason,
9
Recall the transition from matrix elements to vacuum-vacuum ampli-
tudes in the path integral in Sec .[2.7].

90
showing by the way that the Hilbert space must have an
indefinite scalar product. However, zero norm states are
orthogonal to every state vector, so they have zero tran-
sition probability. Therefore physical states are actually
elements of the BRST cohomology, just like the gauge in-
variant functional we considered before. It is reasonable
that in the quotient Hilbert space describing the cohomology
the scalar product is positive definite, since after all we
confined the zero norm states in the null equivalent class.
But unlikely as it seems, there could still be negative
norm states like those associated to the time component of
the gauge field. For a full formal proof that this is not
the case, and that the state cohomology is a true Hilbert
space with no ghosts or anti-ghosts, see.25 A proof for free
fields will be given soon.
One should also prove that the S matrix for states in the
physical Hilbert space is unitary, but this follows if the
S matrix is (pseudo) unitary in the larger space containing
also the unphysical degrees of freedom, which will be the
case if the Lagrangian is Hermitian.
Let us sketch briefly how this question of unitarity can
be resolved in gauge theories using BRST symmetry. First
notice that since QB is conserved and leaves invariant the
Lagrangian, it must commute with the Hamiltonian and the
scattering operator as well:
[QB , H] = 0, [QB , S] = 0 (3.72)
The physical state space is therefore left invariant by the
time evolution and by the scattering operator. Also, both
L and H are Hermitian in the indefinite state space of the
theory, so in this space the scattering operator computed
with this Lagrangian is (pseudo) unitary, S ∗ S = SS ∗ = 1,
implying for each initial state, |i⟩, and final state ⟨f |,
the condition holds10

σn ⟨f |S ∗ |n⟩ ⟨n|S|i⟩ = ⟨f |i⟩ (3.73)
all states

where σn = 0, ±1 determines the scalar product of the countable


basis via ⟨n|m⟩ = σn δnm . In particular, if |i⟩ is physical and
10
∑ Because the completeness relation in a indefinite metric reads
n |n⟩ σn ⟨n| = 1.

91
normalized,

σn | ⟨n|S|i⟩ |2 = 1 (3.74)
n

and we have “negative probabilities” for certain final


states, an example being a ghost-anti-ghost pair |c̄k , cp ⟩. In
general there will be also unphysical positive norm states,
for example a state with two gauge bosons with unphysical
polarization, and these too should be eliminated from the
unitarity sum. That is why in gauge theories one looks for
a true Hilbert subspace of “physical states”, i.e. one pos-
sessing a positive definite scalar product, which is stable
under the time evolution and the scattering operator. In
the Yang-Mills case this is the BRST cohomology. Under
these conditions, and using SS ∗ = S ∗ S = 1, one derives as a
consequence of Eq. (3.72), that

S ∗ S = SS ∗ = 1 (3.75)

where S is the restriction of S to the cohomology, such


that given a cohomology class |ψ⟩ = {|ψ⟩ + QB |α⟩}, S |ψ⟩ is
the cohomology class containing S |ψ⟩. One can easily see
that this definition is consistent. Thus for states in the
cohomology it is true that

⟨ψ|S ∗ S|ϕ⟩ = ⟨ψ|SS ∗ |ϕ⟩ = ⟨ψ|ϕ⟩ (3.76)

Inserting a basis |n⟩ for the cohomology, all of which have


positive norm, we get also

⟨ψ|S ∗ |n⟩ ⟨n| S|ϕ⟩ = ⟨ψ|S |n⟩ ⟨n| S ∗ |ϕ⟩ = ⟨ψ|ϕ⟩ (3.77)
n

and for |ψ⟩ = |ϕ⟩ we have that probabilities really add up to


unity

| ⟨n|S|ϕ⟩ |2 = 1 (3.78)
n

since there do not appear any negative sign. The reason


behind Eq. (3.76) is that since [QB , S] = 0, for |ψ⟩ physical
also S |ψ⟩ is physical, that is QB S |ϕ⟩ = 0. This then implies
that S |ϕ⟩ differs from a physical state with positive norm
(an element of the cohomology) by at most a BRST-exact, zero

92
norm physical state, which is orthogonal to everything else.
Therefore the equation ⟨ψ|S ∗ S|ϕ⟩ = ⟨ψ|ϕ⟩, for example, really
depends only on the cohomology classes of the vectors |ϕ⟩,
|ψ⟩, as Eq. (3.76) indicates11 , showing that S is unitary in
the ordinary sense. By the same token, this means that in
the unitarity equation we can restrict the summation over
the physical, positive norm states, that is put σn = 1 in
(3.73) so that
∑ ∑
⟨ψ|ϕ⟩ = σn ⟨ψ|S ∗ |n⟩ ⟨n| S|ϕ⟩ = ⟨ψ|S ∗ |n⟩ ⟨n| S|ϕ⟩ (3.79)
all states physical
states

Clearly for this to be possible there must be cancellations


among the intermediate unphysical states. The important
point is that for every negative norm state there is an
unphysical positive norm state which eliminates the nega-
tive contribution from the unitarity sum (3.74) and (3.79).
The pairing of negative norm states with positive norm but
unphysical vectors is a consequence of gauge invariance. We
will see an explicit example of these cancellations in the
next section.

3.5 BRST Cohomology in the Fock repre-


sentation
We shall give few examples of this ghost decoupling theorem
using the theory at zero coupling, which should naively
describe the kinematical asymptotic states. We are using
this terminology because the true asymptotic states is a
dynamical affair. For example, it is widely believe that
the SU (3) gauge theory known as QCD permanently confines
all color states in colorless hadrons and mesons, quite
at odds with the free particle spectrum one can read from
the Lagrangian. One can easily quantize the free fields
of the gauge theory. We have in effect N copies of an
abelian gauge field (the gauge index a ∈ {1, . . . , N }) and a
ghost-antighost pair plus an auxiliary field B a all obeying
11
More precisely, this means that S induces a well defined operator S
on the quotient space.

93
the wave equation □c̄a = □ca = □B a = 0, and also
1
B a = ∂ µ Aaµ (3.80)
ξ
The auxiliary field is a zero norm field, since its momentum
πB = −A0 is unrelated to Ḃ and the equal time commutation
relations are [B a , B b ] = [Ḃ a , B b ] = 0, leading to
[B a (x), B b (y)] = 0, [B a , cb ] = [B a , c̄b ] = 0 (3.81)
The only non trivial CAR are in the ghost sector and are
seen to be
[ca (x), c̄b (y)]+ = [ca (x), (c̄b )∗ (y)]+ = iδ ab ∆c (x − y) (3.82)
The remaining anti-commutators all vanish, for example
[ca (x), (cb )∗ (y)]+ = [c̄a (x), (c̄b )∗ (y)]+ = 0 (3.83)

[ca (x), cb (y)]+ = [c̄a (x), c̄b (y)]+ = 0 (3.84)


The causal function here is massless

1
∆c (x) = ϵ(x0 )δ(k 2 )eik·x d4 k (3.85)
(2π)3
The gauge field will be taken in the Feynman gauge ξ = 1, so
that
[Aaµ (x), Abν (y)] = ηµν δ ab ∆c (x − y) (3.86)
and
i ∂∆c (x − y)
[B a (x), Abν (y)] = δ ab (3.87)
ξ ∂xν
All fields are therefore local fields12 . Making Fourier
transforms (see below for notations) we write
∑ ∑
c= [ck eikx + c∗k e−ikx ], c̄ = [c̄k eikx + c̄∗k e−ikx ] (3.88)
k k


B= [Bk eikx + Bk∗ e−ikx ] (3.89)
k

12
This is not a violation of the spin-statistics theorem because the
Hilbert space carries an indefinite metric and spectral conditions are
also violated by the ghosts.

94

A= [Akα ϵαk eikx + A∗kα ϵα∗
k e
−ikx
] (3.90)
k,α

Notational conventions: the sum over k for a continuous


spectrum in all formulas is a symbolic way to denote the
integral with the measure dm(k) = (2|k|)−1/2 (2π)−3/2 d3 k, and δkl =
δ 3 (⃗k − ⃗l).
It will be convenient to choose the polarization vectors ϵα ,
α = (+, −, 1, 2) (these are to be considered as internal Lorentz
indices, not as space-time indices), for given k µ (we omit
k labels for simplicity), in such a way that η µν ϵαµ ϵ± ν = 0 for
α = 1, 2, while for α = ±
1
ϵ± · ϵ± = 0, ϵ+ · ϵ− = −1, ϵ± = √ (|k|, ±⃗k) (3.91)
2|k|
We call ϵ± the forward/backward polarizations, respectively.
The completeness relations read

η µν ϵαµ ϵβν = η αβ , ηαβ ϵαµ ϵβν = ηµν

Here ηµν is the usual metric tensor in cartesian coordi-


nates but ηαβ is the metric tensor in the “coordinates”
y α = (+, −, 1, 2), namely,
 
0 −1 0 0
 −1 0 0 0 
ηαβ =  0
 (3.92)
0 1 0 
0 0 0 1
This can be used to define “covariant” polarizations, via
ϵα = ηαβ ϵβ . The non vanishing CCR now translate to (with 1c
the identity in color space)

[c̄l , c∗k ]+ = −[ck , c̄∗l ]+ = iδkl 1c (3.93)

[Ak+ , (Al− )∗ ] = [Ak− , (Al+ )∗ ] = −δkl 1c (3.94)

i√ ⃗
[Bk∗ , Al+ ] = [A∗l+ , Bk ] = − 2|k|δkl 1c (3.95)
ξ
while physical polarizations have the usual commutation re-
lations corresponding to positive norm states. Note that

95
|±, k, a⟩ = Aa∗
k± |0⟩ are zero-norm states, so the linear combina-
tions |+, k, a⟩ + |−, k, a⟩ are negative norm improper states. From
Eq. (3.80) we have also
i∑ µ α a i√
Bka = k ϵµ Akα = − 2|k| Aak− (3.96)
ξ α ξ

which is consistent with the CCR and will be used later. With
these operators we can build as usual the Fock representation
over a vacuum |0⟩ defined by

cak |0⟩ = c̄ak |0⟩ = Aakα |0⟩ = Bk |0⟩ = 0 ∀k, α, a (3.97)

The ghost quantum number is


∑∫
Qg = −i d3 k (ca∗ a a∗ a
k c̄k + c̄k ck ) (3.98)
a

and has eigenvalues N = n − n̄, n (n̄) being the number of


ghosts (anti-ghosts). The BRST charge at zero coupling
takes the form (see Eq. (3.52))
∑∫
k Bk − Bk ck )
d3 k (ca∗ a a∗ a
QB = i (3.99)
a

Note that [Qg , QB ] = QB , meaning that QB has ghost quantum


number +1. One has the non vanishing commutation relations

[QB , cak ]+ = [QB , Bka ] = 0 (3.100)

i∑ µ α a i√ ⃗ a
[QB , c̄ak ]+ = − k ϵµ Akα = 2|k|Ak− (3.101)
ξ α ξ


[QB , Aak− ] = 0, [QB , Aak+ ] = − 2|⃗k|cak (3.102)

where (3.96) has been used in the second line. We are


ready to analyze the BRST cohomology. Consider the number
operator counting the gauge and ghost modes
∑∫
k ck − ck c̄k ) − Ak+ Ak− − Ak− Ak+ ]
d3 k [i(c̄a∗ a a∗ a a∗ a a∗ a
N= (3.103)
a

The spectrum of N is given by non negative integers and


the only states with eigenvalue zero are those containing

96
unconstrained fields, namely the matter degrees of freedom
and the transverse polarizations. The operator N is also
BRST-invariant

[QB , N ] = 0 (3.104)

so we can investigate the cohomology in a fixed eigenspace.


Actually N is BRST exact: N = [K, QB ]+ , where
√ ∑ ∫ d3 k a∗ a
K = −i 2 (Ak+ c̄k + c̄a∗ a
k Ak+ ) (3.105)

2|k|
a

as can be easily checked. This implies that if |ψ⟩ is a BRST


invariant state which contains a non zero number of gauge
and ghost modes,

N |ψ⟩ = n |ψ⟩ , n ̸= 0

then it is BRST exact because


N 1 ( )
|ψ⟩ = |ψ⟩ = (KQB + QB K) |ψ⟩ = QB Kn |ψ⟩ (3.106)
n n
The only non trivial cohomological class is thus given by
states at ghost number n = 0, with no forward/backward po-
larized gluons, and the only physical degrees of freedom
surviving in the cohomology are therefore the matter fields
and the transversely polarized gluons. This decoupling of
the ghosts and gauge modes (which form “quartets”) from
the cohomology and the scattering operator is known as the
quartet mechanism, after Kugo-Ojima first described it.
The ghost charge operator vanishes on the cohomology, so
the frequently stated condition that physical states should
also obey Qg |phys⟩ = 0 is actually superfluous: it is a
consequence of QB |phys⟩ = 0 and the equivalence relation
|phys⟩ ≃ |phys⟩ + QB |χ⟩.
Digression I: at the level of one-particle states we can see
explicitly how the BRST symmetry works. Adding a transverse
gauge boson to a physical state |ψ⟩ we have

i
QB |ψ, ϵkα ⟩ = k µ ϵαµ |ψ, ck ⟩
ξ

97
so this is physical because k µ ϵαµ = 0 is the usual transversal-
ity condition characterizing physical polarization states.
On the other hand we have


|ψ, cak ⟩ = QB √ |ψ, ϵk− ⟩
2|⃗k|

so adding a ghost we get a physical but exact state, co-


homologous to zero. At the same time, adding a forward
polarized gauge boson gives an unphysical state (containing
a ghost), not in the cohomology. Adding an anti-ghost gives

QB |ψ, c̄k ⟩ = ik · ϵk− |ψ, ϵk− ⟩ ̸= 0

which is not physical too, and does not belong to the


cohomology. Adding a backward gauge boson we obtain an
exact state

|ψ, ϵk− ⟩ = QB (i 2|⃗k|)−1 |ψ, c̄ak ⟩

Continuing in this way we can see that the cohomology is free


from ghosts, anti-ghosts and unphysical gauge bosons, but of
course this is only a rephrasing of the general treatment
we gave above. We only note the following fact, which
illustrates how the ghosts can cancel the backward/forward
amplitude (1.57) thus decoupling from the scattering matrix.
Consider the state |c̄k , ϵp+ ⟩ of an anti-ghost with momentum k
plus a forward polarized gluon with momentum p: then using
the BRST algebra given above we have
( )
QB i√12|⃗k| |c̄k , ϵp+ ⟩ = |ϵk− , ϵp+ ⟩ − i ||⃗⃗k|
p|
|c̄k , cp ⟩ ≡ |+, −; c, c̄⟩ (3.107)

It follows then that for physical state the S-matrix ele-


ments with |+, −; c, c̄⟩ all vanish,

⟨out, β|S|+, −; c, c, out⟩ = 0, ⟨in, +, −; c, c|S|α, in⟩ = 0 (3.108)

Now consider the expansion of the initial q q̄ state over the


final states, putting in evidence the unphysical ones

|α, in⟩ = · · · + ⟨ϵk− , ϵp+ |S|α, in⟩ |ϵk− , ϵp+ , out⟩ + ⟨c̄k , cp |S|α, in⟩ |c̄k , cp , out⟩
+ ······

98

Note that there are no intermediate states k |c̄k , ϵk± ⟩ ⟨c̄k , ϵk± |,
since these have non zero ghost number. From (3.108) we now
obtain
[ ]
|p|
|α, in⟩ = · · · + ⟨ϵk− , ϵp+ |S|α, in⟩ |ϵk− , ϵp+ , out⟩ − i |c̄k , cp , out⟩ + · · ·
|k|
We see that the amplitude for production of unphysical glu-
ons, which worried us in §1.5, appears as the coefficient
of a BRST-exact state whose amplitude vanishes when S is
restricted on physical states (it is a zero-norm state). So
what happens here is that the positive norm, but unphysical,
two-gluon state combines with the negative norm two-ghost
state to form a zero-norm state, which is the zero vector
in the cohomology. That this cancellation occurs for any
intermediate state, not only in the two-particle subspace
just seen, is simply the statement that the cohomology has
positive norm and the scattering operator is well defined
on it. We note that if any unphysical state, like our
|ϵk− , ϵp+ , out⟩, were orthogonal to the physical subspace there
would be no need of this discussion, the unwanted term in
the last equation would simply be absent.
Digression II: the operators (Qg , QB ) generate an algebra A

[Qg , Qg ] = {QB , QB } = 0
(3.109)
[Qg , QB ] = QB
whose representation theory is worth analyzing. If we assume
they do not suffer from spontaneous symmetry breaking, one-
particle states should be classified by representations of
(3.109). These are at most four-dimensional due to the
nilpotence of QB , hence singlets, doublets or quartets.
Kugo and Ojima found three types of representations: (I)
BRST singlets are the physical particles, having Qg = 0,
(II) singlet pairs, which are ghost number conjugate pairs
of two BRST singlets, and (III) the quartet, a ghost number
conjugate pair of two BRST doublets. Actually there are
four types of pseudo-unitary representations of (3.109) in
a pseudo-Hilbert space equipped with a non degenerate scalar
product ⟨·|·⟩. The theorem is:13
13
See M. Henneaux & C. Teitelboim, Quantization of gauge systems,
Ch. [14], 1991.

99
The most general pseudo-unitary representation of the al-
gebra [A] is the direct orthogonal sum of the the following
four blocks:

(i) Singlet:

QB |ϕ⟩ = 0
Qg |ϕ⟩ = 0 (3.110)
⟨ϕ|ϕ⟩ = ±1

(ii) Non null doublet:


QB |ϕ⟩ = 0, QB |ϕ ⟩ = 0
′ ′
Qg |ϕ⟩ = k |ϕ⟩ , Qg |ϕ ⟩ = −k |ϕ ⟩ , k ̸= 0 (3.111)

⟨ϕ|ϕ ⟩ = 1

(iii) Non null doublet at ghost number ±1/2:

QB |f ⟩ = |g⟩ , QB |g⟩ = 0
Qg |f ⟩ = − 12 |f ⟩ , Qg |g⟩ = 12 |g⟩ (3.112)
⟨f |g⟩ = ±1

(iv) Quartet:

QB |f ⟩ = |g⟩ , QB |g⟩ = 0
′ ′ ′
QB |f ⟩ = |g ⟩ , QB |g ⟩ = 0
Qg |f ⟩ = (k − 1) |f ⟩ , Qg |g⟩ = k |g⟩ (3.113)
′ ′ ′ ′
Qg |f ⟩ = −k |f ⟩ , Qg |g ⟩ = (−k + 1) |g ⟩
′ ′
⟨f |g ⟩ = − ⟨f |g⟩ = 1

(The scalar product that are not written, in particular


the norms, are zero).

100
Representation (iii) is only relevant if there are fermionic
constraints, the case (iv) for k = 1 is relevant to Yang-
Mills theories as we have seen. The quartet of states can
′ ′
be taken as |f ⟩ = |ϵ+ ⟩, |f ⟩ = i |c̄⟩, |g ⟩ = |ϵ− ⟩, |g⟩ ∝ |c⟩. All
four states have zero norm, nevertheless there can be some
overlap between them because |ϵ+ ⟩ and |c̄⟩ are not physical and
Hphys has no orthogonal complement in H.

3.6 Slavnov-Taylor identities


BRST symmetry is a powerful tool to prove renormalizability
of non abelian gauge theories. As with any symmetry of
the action and the functional measure, there correspond
important identities among the Green functions which cannot
immediately extrapolated from the analogous identities in
QED. Perhaps the simplest way to obtain the sought for
consequences of the symmetry, is to start from the equation

i[QB , T (Φj1 (1)Φj2 (2) · · · Φjn (n))]± = (3.114)



= (−)ϵk T (Φj1 (1)Φj2 (2) · · · δB Φjk (k) · · · Φjn (n))
k

where Φj (j) indicates any of the fields in the theory eval-


uated at ∑ point xj , each with a certain Grassmann parity σj ,
and ϵk = a<k σja , ϵ0 = 0; moreover δB Φj (x) = i[QB , Φj (x)] is the
BRST variation14 . The equation holds because (i), QB , be-
ing time independent, can be moved inside the time ordering
symbol, (ii), the (anti) commutator acts as a derivation
of the field algebra. Taking the matrix element of (3.114)
between physical states, annihilated by QB , we get

(−)ϵk ⟨Ψ|T (Φj1 (1)Φj2 (2) · · · δB Φjk (k) · · · Φjn (n))|Φ⟩ = 0 (3.115)
k

For the vacuum, this is the statement that the functional


integral

Πj DΦj Φj1 (1)Φj2 (2) · · · Φjn (n) exp iI[Φ] (3.116)

14
It may be easier to consider the commutators with the even generator
λQB , λ being a Grassmann parameter.

101
is unaffected by the the change of variables Φj → Φj + δB Φj .
Adding sources to the action we can say the same of the
generating functional
∫ ( ∫ )
Z[J] = Πj DΦj exp iI[Φ] + i Jk Φk (3.117)

leading to the compact identity


∫ ∫ ∑ ( ∫ )
4
d x Πj DΦj Jk (x)δB Φk (x) exp iI[Φ] + i Jk Φk = 0 (3.118)
k

As an example, let us prove the transversality condition of


the self-energy function of the gauge field. To this aim we
consider the following ST identity for the unrenormalized
fields
δB ⟨0|T ∂ µ Aaµ (x)c̄b (y)|0⟩ = 0 (3.119)
Using the BRST transformations we get
⟨0|T ∂ µ ∇µ ca (x)c̄b (y)|0⟩ + ξ −1 ⟨0|T ∂ µ Aaµ (x)∂ ν Abν (y)|0⟩ = 0 (3.120)
By the ghost field equations only the second term survives,
and this can be written in the equivalent form

⟨0|T Aaµ (x)∂ ν Abν (y)|0⟩ = δ(x0 − y 0 ) ⟨0|T [Aa0 (x), ∂ ν Abν (y)]|0⟩ (3.121)
∂xµ
The commutator is

δ(x0 − y 0 )[Aa0 (x), ∂ ν Abν (y)] = −ξδ(x0 − y 0 )[Aa0 (x), π0b (y)] = iξδ 4 (x − y)δ ab

since ∂ µ Aaµ = −ξπ0a is the momentum conjugate to Aa0 . On the


other hand ⟨0|T Aaµ (x)∂ ν Abν (y)|0⟩ = ∂ ν ⟨0|T Aaµ (x)Abν (y)|0⟩, so finally
∂ ∂
µ ν
⟨0|T Aaµ (x)Abν (y)|0⟩ = iξδ 4 (x − y)δ ab (3.122)
∂x ∂y
Passing to Fourier space, this takes the form
˜ µa;νb (k) = −iξδab
kµkν ∆ (3.123)
For example, the free propagator is

˜ 0µa;νb (k) = −iδab (ηµν − (1 − ξ)kµ kν /k 2 )



k 2 − iϵ

102
and the identity is indeed satisfied. Introducing the self-
energy15 iΠ = ∆−1 − ∆−1
0 , the exact propagator is given by

˜ µa;νb (k) = ∆
∆ ˜ 0µa;νb (k) + ∆
˜ 0µa;ρc (k)iΠρc;σd (k 2 )∆
˜ σd;νb (k) (3.124)

Therefore the identity is true provided the self-energy is


transversal

kµ Πµc;νd (k 2 ) = 0 (3.125)

or

Πµc;νb (k 2 ) = δ ab (k 2 η µν − k µ k ν )π(k 2 ) (3.126)

This will also be true for the renormalized fields as long as


the necessary regularization preserves the BRST symmetry.
The solution of (3.124) is seen to be
( )
−iδ ab
kµ kν kµ kν
∆˜ µa;νb (k) = ηµν − 2 + ξ(1 + Π(k )) 2 (3.127)
2
k 2 (1 + Π(k 2 )) − iϵ k k
This says that gauge vector bosons stay massless to all
order, provided the self-energy does not develop a pole at
k 2 = 0.
Another example is given by equations (3.108), which should
correspond to Slavnov-Taylor identities involving matrix
µ (y)|0, in⟩, with ⟨out, α| any
elements of the kind ⟨out, α|ca (x)Ab+
asymptotic physical state, and Ab+ µ (y) is the part of the
field operator which destroys or creates forward polarized
gluons. We leave it as an exercise to the reader16 .

15
The sum of all 1PI Feynman diagrams with two external lines carrying
momentum k into the diagrams.
16
Use δB Aa+ a
µ (x) = ∂µ c (x), then pass to momentum space.

103
Chapter 4

The effective action


formalism

The EAF is a powerful method enabling one to include all


quantum correction to a given classical theory into one
single non local functional known as the effective action.
This perform many services and is to be counted as one of
the important quantities belonging to QFT.

4.1 Preliminaries
We proceed formally for now. Consider a scalar field theory
with generating functional
∫ ∫ ( ∫ )
Z[J] = ⟨0|T exp i Jϕdx|0⟩ = Dφ exp iA[ϕ] + i Jφdx (4.1)

where A[φ] is the action. We calculate Z[J] by summing all


Feynman diagrams with any number of external lines con-
tracted with the external current J. It is well known that
the functional

W [J] = −i log Z[J]

is the sum of all connected diagrams. The functional deriva-


tive
δW
= φ(x) (4.2)
δJ(x)

104
is clearly the vacuum expectation value of the field operator
in presence of the current J (we may call it the mean field)

φ(x) = ⟨0|T ϕ(x) exp i Jϕdx|0⟩ = ⟨ϕ(x)⟩J (4.3)

and for J = 0 it is simply ⟨0|ϕ(x)|0⟩. We shall assume pro-


visionally that this vanishes, so ⟨ϕ(x)⟩J → 0 as J → 0. Let
J[φ] be the solution of Eq. (4.2), assuming it exists. The
Legendre transformation

Γ[φ] = − φ(x)J(x)dx + W [J] (4.4)

where J = J[φ], is known as the effective action. From this


we may easily the dual equation
δΓ[φ]
= −J(x) (4.5)
δφ(x)
This equation is similar to, and to some extend generalizes,
the classical equations of motion

δA[φc ]
= −J(x)
δφc (x)

In fact Γ[φ] always has the form

Γ[φ] = A[φ] + O(ℏ) quantum corrections (4.6)

A one-loop approximation to Γ will be presented soon using


a path integral description. Given the current J(x), the
possible mean fields are the solutions of the effective field
equation (4.5), which summarizes the quantum corrections to
the classical equations of motion. However, in absence
of symmetry breaking and for J → 0 also φ → 0 since this
corresponds to the vacuum expectation value of the field
operator. But the homogeneous equation
δΓ[φ]
=0 (4.7)
δφ(x)
is actually richer, because it has many more non trivial so-
lutions (as the classical field equations after all). These

105
correspond to matrix elements (rather than to expectation
values)

φ(x) = ⟨out, φ(−) |ϕ(x)|φ(+) , in⟩ (4.8)

between coherent states of the asymptotic fields1 (one can


use smeared fields) defined by the eigenvalue problem
(−)
ϕf (x) |in, φ(−) ⟩ = φ(−) (x) |in, φ(−) ⟩ (4.9)
(+)
⟨out, φ(+) | ϕf (x) = ⟨out, φ(+) | φ(+) (x) (4.10)
(±)
where ϕf (x) are the associated creation/annihilation oper-
ators. Note that this φ(x) has only the negative (positive)
frequency part fixed in the future (past), but it has al-
together both components in the asymptotic regions. For a
proof that (4.8) solves (4.7), see De Witt.22
Each solution of (4.7) with Γ constructed out of Feynman
Green functions corresponds to some set of coherent-state
boundary conditions, and is fixed by giving the positive
frequency part in the remote past and the negative frequency
part in the remote future, as indicated in (4.8). Thus Γ[φ] is
the generator of the quantum dynamics through the effective
field equations (4.7), englobing all quantum corrections to
the classical dynamics.

4.2 Proper vertices


Besides generating the quantum dynamics there is another
property of the effective action that is very important
for field theory. Recall that W [J] generates the connected
n-point functions via the formulae2
δ n W [J]
≡ in−1 ∆n (x1 , . . . , xn ) (4.11)
δJ(x1 ) · · · δJ(xn ) |J=0
= in−1 ⟨0|T ϕ(x1 ) · · · ϕ(xn )|0⟩conn
1
The fields entering the LSZ asymptotic conditions of scattering
theory.
2
One virtue of the Euclidean formalism is that this proliferation of
factors i is totally absent.

106
can be computed by the sum of all connected diagrams with
n external legs. Then Γ[φ] is the generator of the one-
particle irreducible (abbr. 1PI ) amputated diagrams with
n external lines, known as the proper vertices Γn (x1 , . . . , xn ),
δ n Γ[φ]
Γn (x1 , . . . , xn ) = (4.12)
δφ(x1 ) · · · δφ(xn ) |φ=0

We indicate them with a grayed blog, and the functions ∆n


with an n-legs white blog. To show this we differentiate
(4.2) w.r.t. φ, and use (4.5), obtaining readily

δ2Γ δ2W
d4 y = −δ 4 (x − z) (4.13)
δφ(x)δφ(y) δJ(y)δJ(z)

Figure 4.1: The proper vertex

′′
Therefore the second functional derivative Γ (x, y) is (minus)
the inverse of the full propagator. We shorten the equation
by writing symbolically
′′ ′′
Γ ⋆ W = −1 (4.14)
a ⋆ denoting the kernel’s convolution and a prime the func-
tional derivative. The Fourier transform of a convolution
is the product of the Fourier transform of the factors,
therefore if we set

e2 (p) = −p2 − m2 + Π(p2 ), f ′′ (p) = 1


Γ W
p2 + m2 − Π(p2 )

it follows then that iΠ(p2 ) = (1PI diagrams) with two legs
carrying momentum p. Taking a further derivative in φ we

107
get
′′′ ′′ ′′ ′′′ ′′
Γ ⋆W −Γ ⋆W ⋆Γ =0 (4.15)
which shows that the three-point function is
′′′ ′′ ′′′ ′′ ′′
W =W ⋆Γ ⋆W ⋆W (4.16)
or more explicitly
⟨0|T ϕ(x1 )ϕ(x2 )ϕ(x3 )|0⟩ = ∆3 (x1 , x2 , x3 )

′′′
= iΓ (y1 , y2 , y2 )∆2 (x1 , y1 )∆2 (x2 , y2 )∆2 (x3 , y3 )dy1 dy2 dy3 (4.17)
′′′
thus confirming the interpretation of iΓ (x1 , x2 , x3 ) as the
proper three-point vertex, the sum of all connected am-
putated 1PI diagrams with three external lines, because
′′
W = i∆2 is the full propagator. It is represented in
Fig. 4.2, with full propagators on the external lines.

Figure 4.2: The full three-point function

Additional φ-derivatives of the master identity (4.14)


determine the higher order correlation functions in terms
of proper vertices. As a last example, taking the derivative
in J of (4.16) we get
′′′′ ′′′ ′′′ ′′ ′′ ′′ ′′′ ′′′ ′′ ′′ ′′′ ′′ ′′′
W = W ⋆Γ ⋆W ⋆W +W ⋆Γ ⋆W ⋆W +W ⋆Γ ⋆W ⋆W
′′ ′′ ′′′′ ′′ ′′
+ W ⋆W ⋆Γ ⋆W ⋆W (4.18)
Then using (4.16) again we obtain a formula whose graphical
representation3 can be seen in Fig. 4.3, confirming that
iΓ4 (x1 , x2 , x3 , x4 ) is the proper four-point vertex.
3
The sum is over the three independent ways to arrange the position
of the external legs.

108
+
=

Figure 4.3: The full four-point function

One may take any number of J-derivatives of equations like


(4.16) or (4.18): the general structure is always
′′ ′′

| ⋆ ·{z
W (n) = W · · ⋆ W } ⋆Γ(n) + 1P reducible terms (4.19)
n times

so Γ(n) is 1PI. Note that we can easily write Eq. (4.1) in


terms of the effective action alone, by using Γ = W − J ⋆ φ

with J = −Γ to eliminate W . Then
∫ ( )

exp iΓ[φ] = Dχ exp iA[φ + χ] exp −iΓ [φ] ⋆ χ (4.20)


If φ were the true vacuum expectation value of ϕ then Γ [φ]
would drop out from (4.20), and the integral would be taken
over the shifted action, keeping only 1PI diagrams of course.
In this way

exp iΓ[φ] = Dχ exp iA[φ + χ] (4.21)
1P I

But this must be true for general φ, because the integral


doesn’t know nothing about the true value of φ. This means
that the role of the linear term in (4.20) is precisely
to cancel all 1P reducible diagrams from the perturbative
evaluation of Γ[φ], so the effective action is determined by
1PI diagrams alone.
There is one more interesting property of the effective
action worth to mention. Consider the functional integral
and define a new WΓ via

iWΓ [J]/ℏ
e = Dφ exp(iΓ[φ] + iJ ⋆ φ)/ℏ (4.22)

109
where ℏ has been reinserted as an expansion parameter in the
semi-classical limit ℏ → 0. Each internal line (propagator)
carries then a factor ℏ and each vertex a factor ℏ−1 , so
each connected diagram with I propagator and V vertices has
a global factor ℏI−V = ℏL−1 , where L = I − V + 1 is the number
of loops. Therefore the left hand side has a loop expansion

ℏ−1 WΓ [J] = ℏL−1 WL [J]
L≥0

In the semi-classical limit the functional integral is dom-


inated by the stationary phase of the exponent, which occurs
at solutions of the effective field equations

δΓ[φ]
= −J(x)
δφ(x)

Let the solution be φJ , a functional of J. To leading order


in ℏ the functional integral takes the value

i i
exp (Γ[φJ ] + J ⋆ φJ ) (1 + O(ℏ)) = exp W [J](1 + O(ℏ))
ℏ ℏ

by the definition of W [J] as the Legendre transform of the


effective action. We see that
W [J] = W0 [J] (4.23)
which means that W [J] can be calculated by summing the
tree level diagrams generated by the full effective action,
that is using exact propagators for lines and exact proper
functions for vertices.

4.3 The effective potential


The expansion of the effective action in powers of the
external field is
∑ 1 ∫
Γ[φ] = Γn (x1 , . . . , xn )φ(x1 ) · · · φ(xn )dx1 · · · dxn (4.24)
n
n!

The (divergent) constant Γ0 = Γ[0] is just the sum of all


1PI vacuum diagrams, and should be cancelled by the bare

110
cosmological constant to enormous precision. We simply put
it to zero. The function Γ1 (x) is the sum of all tadpole
diagrams, the 1PI diagrams with only one external line.
Clearly

δΓ
= Γ1 (x)
δφ(x) |φ=0

If the sum of all tadpole graphs happens to be non vanishing,


this is the sign that we are expanding around the wrong
vacuum and φ = 0 is not the true ground state expectation
value.
Of particular interest is the case of constant φ: in that
case we put the field in a large space-time box with volume
Ω = V3 T and define the effective potential as the ordinary
function Veff (φ)

−Ω−1 Γ[φ] = Veff (φ) (4.25)

This is just the classical potential plus quantum correc-


tions, up to renormalization.
The effective potential has many interesting properties, of
which a very important one is the energy interpretation:
Veff (φ) is the minimum value of the energy density of the
quantum field taken over the class of unit norm states,
(Ψ, Ψ) = 1, for which

(Ψ, ϕ(x)Ψ) = φ(x)

In other words, for the given unit norm states and the
Hamiltonian H
1
min(Ψ, HΨ) = − Γ[φ] = V3 Veff (φ) (4.26)
T
The proof rests on the fact that for an adiabatic switching
of the current from zero to some constant value J (x) which
last some long time T and whose Fourier transform do not
intersect the spectrum of the mass operator −Pµ P µ (so that
there is not particle creation), and finally adiabatically
switches off to zero again, the vacuum is converted to a
state |ΨJ ⟩ with a certain energy EJ , a functional of J .

111
The vacuum-vacuum amplitude Z[J] takes then the value

Z[J] = exp(−iEJ T )

because this is the phase accumulated by the state |ΨJ ⟩


during the time T . Therefore

W [J] = −EJ T

The quantum mechanical adiabatic theorem4 guarantees that


EJ is the ground state of the Hamiltonian H − J ⋆ ϕ, that
is there is a state |ΨJ ⟩ such that

H |ΨJ ⟩ − J ⋆ ϕ |ΨJ ⟩ = EJ |ΨJ ⟩

But this is the solution of the variational problem posed


above (via the Lagrange multipliers method5 ) for that current
Jφ such that ⟨ΨJφ |ϕ|ΨJφ ⟩ = φ. Therefore the minimum is attained
at |ΨJφ ⟩, and taking the scalar product with ⟨ΨJφ | we find

min (Ψ, HΨ) = ⟨ΨJφ |H|ΨJφ ⟩


( ∫ )
1 1
= EJφ + Jφ ⋆ φ = − W [Jφ ] − 4
Jφ φd x = − Γ[φ] (4.27)
T T

as we wanted to show, if we recall that Γ[φ]/T = −V3 Veff (φ).


There is another way to get the same result. It is a
marvelous property of the Lagrange multipliers method that
the multiplier at a solution, in our case the current Jφ ,
is the rate at which the minimum (or the maximum for that
matter) changes as φ(⃗x) changes, that is

δ ⟨ΨJφ |H|ΨJφ ⟩ δΓ[φ]


= Jφ = −
δφ(⃗x) δφ(⃗x)

the last equality being just one of the basic properties of


the effective action. Therefore up to constant Eq. (4.27)
4
See the well known textbook “Quantum Mechanics”, Vol. 2, by A. Mes-
siah.
∫ Which consist3 in minimizing the expression L = (Ψ, HΨ) − α[(Ψ, Ψ) − 1] −
5

Jn [(Ψ, ϕn Ψ) − φn ]d x over Ψ, α and J for fixed φ.

112
holds (question: why the factor 1/T ?).

4.4 Symmetry and symmetry breaking


Symmetry - If ϕ is a multiplet, with components ϕn , belonging
to some orthogonal representation of an internal symmetry
group G of the action, and if the path integral measure
is invariant
∑ as well, then Z[J] is invariant under the map
Jn → m Rmn Jm , for R ∈ G

Z[J] = Z[R · J], W [J] = W [R · J]

This becomes obvious by changing variable φ → R−1 · φ in the


path integral. Then the effective action and the effective
potential have exactly the same symmetry
Γ[R · φ] = Γ[φ] (4.28)
as can be also seen from (4.21). For a continuous symmetry
with infinitesimal variation

δφn = i ϵα (T α )nm φm
m

the invariance condition takes the form


∑ ∫ δΓ[φ]
(T α )nm φm (x)d4 x = 0 (4.29)
nm
δφ n (x)

valid for any symmetry generator T α . In particular, the


effective potential must satisfy the invariance conditions
∑ ∂Veff (φ)
(T α )nm φm = 0 (4.30)
nm
∂φ n

a useful formula to study the occurrence of symmetry break-


ing. The symmetry of the effective action can be extended
to transformations much more general than those of linear
symmetry groups, as soon as the measure and the action or
their product are both invariant.
Symmetry breaking - As anticipation to the following chap-
ter, we briefly introduce the case of symmetry breaking.

113
Above we assumed that ⟨0|ϕ(x)|0⟩ vanishes for vanishing source
J. But it may not. When ⟨0|ϕn (x)|0⟩ = vn (a non zero constant
vector), namely when
δW [J]
= vn (4.31)
δJn (x) |J=0
we say that the symmetry is broken by the expectation value
of the field, because obviously a non zero v would not be in-
variant under all symmetry transformations in G. This is not
in contradiction with the invariance condition W [RJ] = W [J],
which together with (4.31) seems to imply the invariance of
v, because Eq. (4.31) has multiple solutions in the case
of symmetry breaking. For in fact, if the limit of δW /δJn
as Jn → 0 were unique, then differentiating the invariance
condition W [RJ] = W [J] and setting J = 0 we would obtain the
invariance of v under all G, hence v = 0. On the other hand,
there could be set of elements h such that h·v = v, which form
in this case the residual symmetry group H, and we say that
the symmetry is spontaneously broken from G down to H. The
space of vacua is then isomorphic to the coset space G/H.
Whether or not there is a residual symmetry group, if we
pick a vacuum vector vn and J is the current such that the
eq. (4.31) holds, then W [J] is the generating functional
of the connected n-point functions of the shifted fields
ϕen = ϕn − vn
δ n W [J] e 1 ) · · · ϕ(x
e n )|v⟩
W (n) ≡ = in−1 ⟨v|T ϕ(x (4.32)
δJ(x1 ) · · · δJ(xn ) |J=0 conn

where |v⟩ is the vacuum in which ⟨v|ϕn |v⟩ = vn . For example,


since W [J] = −i log Z[J],
′′ δ 2 log Z[J] ′′ ′ ′
W = −i = −i(Z [0] − Z [0]Z [0]) (4.33)
δJ(x1 )δJ(x2 ) |J=0
e 1 )ϕ(x
= i(⟨ϕ(x1 )ϕ(x2 )⟩ − ⟨ϕ(x1 )⟩ ⟨ϕ(x2 )⟩) = i ⟨v|T ϕ(x e 2 )|v⟩
conn

The general case can be proved by induction (see, e.g. Abers


and Lee’s renowned review26 ).
The effective field equations now become
δΓ[φ]
= −J(x) (4.34)
δφ(x)

114
But φn = vn when Jn → 0, therefore

δΓ[φ] ∂Veff
= 0 or =0 (4.35)
δφ(x) φ=v ∂φ φ=v
According to this important equation, the vacuum expecta-
tion value v = ⟨v|ϕ|v⟩ is a stationary point of the effective
potential, one which is actually a minimum as we know from
the energy interpretation (as we will see, Veff is actually
a convex function). Note that this is not incompatible
with the symmetry of the effective action, which states
the invariance Γ[Rφ] = Γ[φ]. In fact this is the landmark
of spontaneous symmetry breaking: a symmetry of the action
which is not a symmetry of its solutions.
Higher order functional derivatives of the effective action
still give the proper vertices of the theory, except that
they are evaluated at φn = vn rather than φn = 0. Therefore
since Γ(1) (x) = 0 (equation (4.35)) the generating functional
Γ[φ] has the expansion
∑∞
1 (n)
Γ[φ] = Γ ⋆ (φ − v)i1 ⋆ · · · ⋆ (φ − v)in (4.36)
n=2
n! i1 ...in

For constant φ(x) = ϕ this is the expansion of the effective


potential
∑∞
1 e(n)
Veff (φ) = − Γi1 ...in (0)(ϕ − v)i1 · · · (ϕ − v)in (4.37)
n=2
n!

e(n) (p1 , . . . , pn )(2π)4 δ 4 (∑ pi ) is the Fourier transform of


where Γi1 ...in
the proper vertices. In particular, the total tadpole at
zero momentum vanishes: Γ e(1) (0) = 0, so that
i

∂V
+ Ti = 0 (4.38)
∂ϕi |ϕ=v

where Ti is the contribution of radiative corrections to


the tadpole and V is the classical potential. Thus ϕ = v
does not minimize the classical potential when the tadpole
is non vanishing. A formula for the shift in the vacuum
expectation value in one-loop approximation can be simply

115
derived. Writing v = vc + δT v, where vc is the tree level
minimum, we get, since δT v is very small

∂ 2V (1)
δT vj + Ti = 0
∂ϕi ∂ϕj |vc

(1)
where Ti is the one-loop approximation to the tadpole. An
immediate constraint arises: since the mass matrix Mij (the
second derivative term) has zero eigenvalues with eigenvec-
(1)
tors ui , then obviously ui Ti = 0. In the orthogonal subspace
M is strictly positive, and solving for δT v we get
δT vi = −Mij−2 Ti
(1) (1)
= −i∆ij (0)Tj (4.39)
where the last term is the scalar propagator, ⟨ϕi (p)ϕj (−p)⟩,
evaluated at zero momentum.

4.5 The one-loop approximation


Finally we provide a formula for the one-loop evaluation of
the effective action. To this purpose the useful equation
is (4.21). Expanding the action to second order gives

′ 1 ′′
A[φ + χ] = A[φ] + A [φ] ⋆ χ + χ ⋆ A [φ] ⋆ χ + · · ·
2
The first term is sorted out of the path integral. The linear
term does not contribute to 1PI graphs and can be omitted.
It also vanishes for solutions of the classical equations of
motion. Neglecting the higher order terms is just the one-
loop approximation. The integral of the quadratic term is
automatically 1PI, because a Gaussian path integral has no
other diagrams. The path integral reduces to a functional
determinant, therefore remembering the normalization Γ[0] = 0
we get6
i ( ′′ ′′
)
Γ[φ] = A[φ] + log det(A [φ])/ det A [0] (4.40)
2
For most operator appearing in quantum field theory the
determinant can be defined in Euclidean space, where the
6
The classical action A[φ] vanishes for zero field because we set the
bare cosmological constant to zero.

116
energy becomes imaginary according to Wick’s rule, p0 → ip4 .
This then eliminates the i in (4.40).
The second order derivatives of the action is typically a
differential operator; for example in the case of a scalar
field theory with quartic interaction, it is
′′ λ
A [φ] = (□ − m2 − φ2 (x))δ 4 (x − y) (4.41)
2
The computation of the determinant is obviously practical
for constant background field, but generally impossible for
varying backgrounds, and further approximations are needed.
It is interesting that inserting ℏ back in the functional
integral and expanding in powers of ℏ1/2 χ around φ we dis-
cover soon that the quantum one-loop correction is of order
ℏ. That is the loop expansion is a power series in ℏL−1 .
The particular case of a constant φ can be computed making
use of the formula log det M = Tr log M , which gives the ef-
fective potential as the integral in d-dimensional Euclidean
space (whereby p0 = ipd )
∫ ∞ ( )
1 d−1 λφ2 /2
Veff (φ) = Vcl (φ) + d d/2 p dp log 1 + 2 (4.42)
2 π Γ(d/2) 0 p + m2
In one dimension (quantum mechanics) the integral converges,
in d = 4 it diverges quadratically. The divergent terms
are proportional to φ2 and φ4 , respectively, and can be
eliminated by mass and coupling constant renormalizations,
as would be expected on the basis of renormalization theory.
Remarkably, even for zero mass the integral converges in the
infrared. The case

1 λ
Vcl (φ) = m2 φ2 + φ4
2 4!
is well studied. With the renormalization conditions
′′ ′′′′
Veff (0) = m, Veff (0) = λ

and for m2 > 0, the result is the convex function


( )
3m4 1 2 λ λ
Veff (φ) = 2
+ m 1+ 2
φ2 + φ4
128π 2 32π 4!
( ( ) )
λ 2 λ
( 2 φ + m2 )2 φ 2
+ m2
3
+ log 2 − (4.43)
64π 2 m2 2

117
For negative m2 this potential is complex precisely in
the range of φ for which the classical potential is non
convex, and the one-loop approximation breaks down. In fact,
the full effective potential of a theory which possesses a
Euclidean formulation with positive two-point correlation
function must be a convex function.
The massless case is more interesting. In this case we
cannot define the renormalized coupling at zero field, but
employ instead the safer condition at some arbitrary mass
scale M
′′′′
V (M ) = λM (4.44)

The integral (4.42) for m2 = 0 exists, since it can be


reduced to the form
∫ ∞ ( )
M2 M 2n+2 π d−2 ϵ
n
y log 1 + =− , n= =1−
0 y n + 1 sin πn 2 2
′′
with M 2 = λϕ2 /2. One can verify that V (0) = 0. In terms of
the coupling λM the potential takes the Coleman-Weinberg
form
[ ( 2) ]
λM 4 λ2M 4 ϕ 25
Veff (ϕ) = ϕ + ϕ log − (4.45)
4! 256π 2 M2 6
This exhibits now two symmetric minima away from the ori-
gin, indulging to the conclusion that radiative corrections
generate symmetry breaking. But as we will see, in this
particular case the conclusion is wrong.
However an interesting idea now creeps in: the coupling
λM must exhibit a mass dependence such that the effec-
tive potential in invariant under changes of the arbitrary
mass scale M . To one loop order this condition gives the
differential equation
dλM 3λ2
M = M2 + O(λ3M ) (4.46)
dM 16π
showing that λM increases with the mass scale. The phys-
ical meaning is that the ϕ4 interaction becomes stronger
at shorter distances, invalidating perturbation theory for
sufficiently high energy.

118
4.6 Convexity
One can show that if the Euclidean path integral is well
defined, then the effective potential must be a convex
function. In fact in Euclidean space we have

exp W [J] = (dϕ)e−A[ϕ]+J⋆ϕ (4.47)

where A[ϕ] is now the positive Euclidean action. We use


Hölder inequality for functions in Lp spaces, which states
that
∫ (∫ )1/p (∫ )1/q
dµ|f · g| ≤ |f | dµ
p
|g| dµ
q
(4.48)

for f ∈ Lp , g ∈ Lq and 1
p
+ 1q = 1 and dµ a positive measure. Then
for x ∈ [0, 1] we get

exp W [xJ1 + (1 − x)J2 ] = (dϕ)e−A[ϕ] exJ1 ⋆ϕ e(1−x)J2 ⋆ϕ (4.49)
(∫ )x(∫ )1−x
−A[ϕ]+J1 ⋆ϕ −A[ϕ]+J2 ⋆ϕ
≤ (dϕ)e (dϕ)e

= exp(xW [J1 ] + (1 − x)W [J2 ]) (4.50)

from which we get the convexity property

W [xJ1 + (1 − x)J2 ] ≤ xW [J1 ] + (1 − x)W [J2 ] (4.51)

We define the Euclidean effective action using the more


general form of the Legendre transform

Γ[ϕ] = sup(J ⋆ ϕ − W [J]) (4.52)


J

which is also valid for non differentiable functionals. The


sup exists precisely when W [J] is convex. Then using the
standard inequality supI (f + g) ≤ supI + supI g, for any interval
I, we get the convexity of the effective action

Γ[xϕ1 + (1 − x)ϕ2 ] ≤ xΓ[ϕ1 ] + (1 − x)Γ[ϕ2 ] (4.53)

The effective potential is just the Euclidean effective


action per unit euclidean volume, whence its convexity fol-
lows.

119
Chapter 5

Spontaneous symmetry
breaking

In field theory it may happens that a physical system has a


symmetry in the Lagrangian and the equations of motion which
is not a symmetry of their actual solutions, in particular
of the vacuum which then appears to be degenerate1 . In that
case the symmetry of the equations of motion (the symmetry
of the world according to Coleman) cannot be easily inferred
from the properties of their solutions, and one says that
the symmetry is spontaneously broken (hidden symmetry would
a more appropriate designation). This phenomenon has had
many fruitful ramifications, and in the case of the gauge
symmetry it led ultimately to the modern unified theory of
weak and electromagnetic interaction known as the Weinberg-
Salam model.

5.1 Classical theory


Consider as a simple example the Lagrangian for a doublet
of classical (not quantized) scalar fields
1 1 1 1
L = − (∂µ ϕ1 )2 − (∂µ ϕ2 )2 − m21 ϕ21 − m22 ϕ22 (5.1)
2 2 2 2
λ1 4 λ2 4 λ 2 2
− ϕ − ϕ2 − ϕ1 ϕ2
4 1 4 2
1
If the vacuum is unique then it is also invariant, since otherwise
Qα |0⟩ would be a different state with the same energy as |0⟩, where Qα
is any symmetry generator.

120
This theory has a discrete Z2 × Z2 manifest symmetry under
which ϕ1 → −ϕ1 , ϕ2 → −ϕ2 . For certain values of the pa-
rameters it may have enhanced symmetries. For example if
m1 = m2 = m and λ1 = λ2 there is a further exchange symmetry
under which ϕ1 → ϕ2 , ϕ2 → ϕ1 and the group becomes Z2 × Z2 × Z2 .
If also λ1 = λ2 = λ there is an additional continuous SO(2)
symmetry acting in the guise of rotations, but in field
space
ϕ1 → ϕ1 cos θ − ϕ2 sin θ
(5.2)
ϕ2 → ϕ1 sin θ + ϕ2 cos θ
which is rendered manifest by writing the Lagrangian as
1 1 1 λ
L = − (∂µ ϕ1 )2 − (∂µ ϕ2 )2 − m2 (ϕ21 + ϕ22 ) − (ϕ21 + ϕ22 )2 (5.3)
2 2 2 4
It combines with the discrete symmetries to form the group
O(2). Now consider the configurations with minimum energy
density, at first in the classical theory. We have two
important cases
(i) the symmetric phase with m2 > 0: the minimum is at
ϕ1 = ϕ2 = 0, it is rotational invariant with zero energy
density. In the quantum version there is a unique in-
variant vacuum, the symmetry is unitarily implementable
and the particle states fall into unitary representa-
tions of the group. This is all the way familiar.

(ii) the broken phase with m2 < 0: the minimum is degenerate


at points on the circle

m2
ϕ21 + ϕ22 = − = v2 (5.4)
λ
The symmetry maps vacuum configurations to different
vacuums. We (or the system itself via relaxation mech-
anisms) may choose one, for example ϕ0 = (u, v), and con-
sider fluctuations around it: ϕ = (u + φ1 , v + φ2 ). It
24, 38, 39
is a mathematical fact that these solutions form
“vacuum sectors”, stable by the time evolution, but
not invariant under the O(2)-symmetry transformations.
There is a vacuum sector for each choice of the poten-
tial’s minimum, but there are also other, non vacuum,

121
sectors in the space of solutions. The vacuum sec-
tors have the form Sϕ0 = {ϕ0 + φ(t), φ(t) ∈ H 1 (R3 )} (the first
Sobolev space). In particular, there are no finite
energy solutions of the field equations interpolating
between different sectors2 . In a sense, each vacuum acts
as a superselection rule. On the other hand, pertur-
bations not decaying at spatial infinity sufficiently
fast have infinite energy, and are not realizable phys-
ically. Therefore it is the infinite spatial volume and
the non linearity of the field system that makes all
the difference. A symmetry transformation will rotate
the vacuum vector, therefore it cannot act as a sym-
metry within a fixed vacuum sector. For example, the
natural action ϕ0 + φ → ϕ0 + g · φ of the group G furnish
a representation of G, but it doesn’t define a symme-
try transformation because it fails to commute with the
time evolution. This means that the symmetry is broken:
the group transformations do not act on the space of
solutions with a fixed vacuum, and the group is not im-
plementable. But of course for any solution the rotated
one is again a solution, since the equations of motion
are still symmetric, only it belongs to a different
sector and cannot be reached from the former by local
operations only (or finite energy perturbations). In
this sense the symmetry is spontaneously broken. Note
that there is no energetic barrier separating the minima
of the potential related by symmetry in the continuous
case, still there is no physically realizable operation
(a finite energy perturbation, see footnote 2) leading
from one sector to another, even an “adjacent” one.

We can always choose the vacuum sector with ϕ0 = (0, v) and


ϕ = (χ, v + φ). In this ϕ0 -world, we obtain the potential
1 λ λ
V (ϕ) = V0 + m2φ φ2 + (φ4 + χ4 ) + χ2 φ2 + λvφ3 + λvχ2 φ (5.5)
2 4 2
It has no tree level mass term for χ, which is thus a massless
field, but for φ the (zeroth order) mass is m2φ = −2m2 . Note
the presence of trilinear couplings, missing in the original
2
No perturbations with initial data φ(⃗x, 0) ∈ H 1 (R3 ) (the Sobolev space)
and φ̇(⃗x, 0) ∈ L2 (R3 ).

122
theory.
We may understand better the breaking of the symmetry by
studying its relation with the classical Noether theorem,
which implies the existence of a conserved current whether
or not the symmetry is broken. For our simple model this is

J µ = ϕ2 ∂ µ ϕ1 − ϕ1 ∂ µ ϕ2 = v∂ µ χ + φ ∂ µ χ (5.6)

The would be conserved charge (which we recall is the gen-


erator of the symmetry transformation)

Q = lim J 0 (t, ⃗x)d3 x (5.7)
R→∞ |x|<R

actually requires sufficiently fast decay at spatial infin-


ity for the integral to converge. But due to the presence of
the vacuum term and the massless nature of χ, J 0 = −v χ̇ + · · · ,
which makes the integral diverging for solutions with wave-
like character at infinity, unless v = 0, which is not.
Thus the symmetry is not implementable in the vacuum sector
Sϕ0 . It can be shown that each group element g actually
defines a linear map T (g) from sectors to sectors, in the
way T (g) : Sϕ0 → Sg·ϕ0 as one may have guessed.
An analogous statement in quantum theory is that there
are local or quasi-local operators3 Oqloc , one example being
the quantum field itself, such that in a degenerate vacuum

[ ]
lim d3 x ⟨0| J 0 (t, ⃗x), Oqloc |0⟩ ̸= 0 (5.8)
R→∞ |x|<R

which of course could not be true if Q |0⟩ = 0. The symmetry is


said to be spontaneously broken in the vacuum, or realized
in the Nambu-Goldstone mode. The zero mass particle of the
field χ in the quantum version is known as the Goldstone
boson, a zero spin particle with the same parity and quantum
numbers as the corresponding Noether current.
We have here a classical version of the celebrated Gold-
stone’s theorem, that a broken continuous symmetry entails
the presence of a massless field in the theory. But the
3
A local∑operator
∫ for a scalar field theory is an expression of the
form O = n≤M dx1 · · · dxn gn (x1 , . . . , xn )ϕ(x1 ) · · · ϕ(xn ), where gn have compact
support in R4n . If gn ∈ S (R4n ) the operator is quasi-local.

123
symmetry still exists in the Lagrangian, albeit non linearly
realized. Indeed the (affine) transformations

δχ = −α(v + φ), δφ = αχ (5.9)

are still a symmetry of the Lagrangian, but not of the


ground state since they change the expectation value of χ.
The massless field does not appears if a discrete symmetry
is broken by the non vanishing vacuum value of the fields.
For example, let m21 < 0 but m22 > 0 in the Lagrangian (5.1):
then one Z2 is broken by the vacuum values ϕ̄1 = v = −m21 /λ1 ,
ϕ̄2 = 0 but the fluctuations around this vacuum are still
massive, with masses respectively given by −2m21 and m22 .
The Lagrangian (5.3) is naturally generalized to a O(N )-
symmetric model: simply take
( )2
1∑ 1 ∑ λ ∑
L=− (∂ϕn )2 − m2 ϕ2n − ϕ2n (5.10)
2 n 2 n
4 n

If m2 < 0 the potential can be written as


( )2
λ ∑
V (ϕ) = ϕn ϕn − v 2 + V0 (5.11)
4 n

where v 2 = −m2 /λ. There is then a degenerate non zero vacuum


value at any ϕ = ϕ̄, where
∑ m2
ϕ̄2n = v 2 = − (5.12)
n
λ

We may choose one vacuum configuration along the N -th di-


rection: ϕ̄n = vδnN , and write a general field as
 
χ1
 χ2 
 
 . 
ϕ=  (5.13)
. 
 
 χN −1 
v+φ

124
The N − 1 fields χk are the massless Goldstone bosons and
m2φ = −2m2 . The vacuum
 
0
 0 
 
 . 
ϕ̄ =  
 .  (5.14)
 
 0 
v
is left invariant under the O(N − 1) subgroup of O(N ) acting
on the first N − 1 coordinates. In infinitesimal terms the
symmetry acts as

δϕn = i ϵα [Tα ]nm ϕm (5.15)
m

where Tα∗ = −Tα , TαT = −Tα , because we are in a real rep-


resentation of the group. The matrices Tα are the group
generators. The number of broken generators is therefore
dim(O(N )) − dim(O(N − 1)) = N − 1 and we say that the symmetry
is broken by the vacuum from O(N ) to O(N − 1). We have
one massless field for each broken generator. The Noether
current is
∑ ∑
Jαµ = i ∂ µ ϕn [Tα ]nm ϕm = i ∂ µ φn [Tα ]nm (ϕ̄n + φm ) (5.16)
the last expression being valid in a vacuum sector. Thus
the generator has a contribution from the minimum
∑∫
Qα = φ̇n [Tα ]nm ϕ̄m d3 x + · · · (5.17)

and this makes the integral divergent unless m [Tα ]nm ϕ̄m = 0,
i.e for the unbroken generators. Thus the charge Qα exists
for the unbroken symmetry group of the vacuum, not for the
other generators, and the symmetry is truly spontaneously
broken from O(N ) to the stability subgroup O(N − 1) of the
vacuum vector. The fields

πα (x) = φn (x)[Tα ]nm ϕ̄m (5.18)

entering the linear divergent term in Qα are the massless


Goldstone fields of the theory, which take the responsibil-
ity for the non existence of Qα .

125
Here is a more complicated model with SU (2) symmetry. We take
a scalar doublet Φ = (Φ1 , Φ2 ) plus a spinor doublet Ψ = (Ψ1 , Ψ2 ),
with Lagrangian (we assume g, h > 0)

g
L = −∂µ Φ† ∂ µ Φ − µ2 Φ† Φ − (Φ† Φ)2 − ΨL ∂/ΨR − ΨR ∂/ΨL −
2
− hΨR Φ† ΨL − hΨL ΦΨR (5.19)

Here Φ† ΨL is the spinor 2−1 Φ∗1 (1 + γ5 )Ψ1 + 2−1 Φ∗2 (1 + γ5 )Ψ2 and ΨR =
(1 − γ5 )Ψ/2 is the right handed part. For negative µ2 the
symmetry is broken by the minimum at Φ† Φ = −µ2 /g ≡ v 2 ; we
take advantage of the symmetry to take the vacuum value
at Φ = (v, 0), and expand around this vacuum Φ = (v + ϕ, χ) for
generally complex fields ϕ = σ + iπ1 , χ = π2 + iπ3 . The three
fields πj are then massless and σ has mass −4µ2 . Perhaps
more importantly, the expansion generates a mass term for
the fermions

hΨR Φ† ΨL + hΨL ΦΨR → hv(ΨR ΨL + ΨL ΨR ) = hvΨΨ

with equal masses m = hv. So far the discussion has been


done for classical field theory. The transition to quantum
field theory must face several problems, of which two seem
the most urgent: (i) should not quantum theory predicts the
possibility that the system makes quantum transitions to
other vacuums, effectively restoring the assumed breaking
of the symmetry? Correspondingly, should not we sum the path
integral over all minima of the classical potential? (ii)
upon quantization should not radiative corrections raise
the mass of the Goldstone bosons? More generally, how do
the broken symmetry constraint the UV divergences?
In fact the first possibility is often realized for finite
(hence non relativistic) quantum mechanical systems. The
ground state of a self-adjoint Hamiltonian for finitely
many spinless degrees of freedom with a discrete spectrum
is always unique. Consider for example the double well
potential
g
V (q) = (q 2 − v 2 )2
4

126
which has two local minima at q = ±v and a local maximum at
q = 0. The states of minimum energy |±⟩ such that ⟨±|q|±⟩ = ±v
are the closest quantum analogue of the classical vacuums
but the overlap ⟨−|+⟩ ̸= 0, because the potential barrier
in between has finite height. Therefore transitions are
possible, and the true vacuum is in fact a unique symmetric
combination of |+⟩ and |−⟩. For infinite systems, involving
infinitely many dynamical degrees of freedom, in particular
for spatially infinite systems, things are different because
the transition probability among different vacua becomes
vanishingly small and vanishes in the limit.

5.2 Decoupling of vacua


The frequently cited example is the Heisenberg model, where
the Hamiltonian is
J ∑
H=− Sa · Sb (5.20)
2 d
a,b∈Z

and the sum is over the nearest neighbors of each spin


in the d dimensional finite lattice. Clearly, the ground
state is reached when all spin operators are aligned along
an arbitrary direction n (this can be proved mathemati-
cally of course), and therefore has degeneracy 2S + 1, where
S = N j is the total spin for a lattice with N points. Let
|n, j⟩ be a basis for the∑ground state manifold, such that
n · S |n, j⟩ = jN |n, j⟩, S = a Sa ; each vector |n, j⟩ is a pos-
sible ground state solution for the eigenvalue problem of
H, but clearly it has less symmetry than the Hamiltonian
itself, which is fully rotational invariant. But this is
not really a breaking of the symmetry, the Hilbert space
of the Hamiltonian still is acted on by unitary rotation
operators giving a representation of SU (2) or O(3). More-
over, the different ground states are not orthogonal, since
⟨j, n|m, j⟩ = (n · m)N = (cos θnm )N , and even a small perturbation
to H will mix them easily. The expectation values of ten-
sor operators are then given by standard quantum mechanical
formula
Tr(RPΩ )
Exp(R) = (5.21)
TrPΩ

127
where PΩ is the projection operator over the ground state
manifold, i.e. one “sums over all minima’ of H. In par-
ticular, Exp(S) = 0.
Where the symmetry really is broken is in the infinite N
limit, namely for a spatially macroscopic spin system. In
this limit all “vacua” became orthogonal

lim ⟨j, n|m, j⟩ = lim (cos θnm )N = 0 (5.22)


N →∞ N →∞

Not only this: all matrix elements of local operators4 also


vanish in the thermodynamic limit, so if the system for
some reason relaxes to one symmetry breaking ground state
|j, n⟩, no perturbation can drive it to explore other vacua,
they are effectively decoupled. In particular, the only
generator of rotations one can abstractly define5 is S · n,
if n specify the actual ground state of interest, and the
symmetry is really broken down to the subgroup U (1) or SO(2)
of rotations about n.
Note: - The discussion of infinite spin systems was here
oversimplified. A proper treatment involves a wealth of non
trivial mathematics, e.g. the existence of infinite sums
of bounded operators, the representation problem, or the
existence of inequivalent representations of the Heisenberg
algebra, the role of super-selection sectors, the defini-
tion of symmetry operations, phase transitions, and so on.
Our purpose was only to expose the radical change of view
the infinite volume limit leads to.
A very important consequence of the existence of a continu-
ous ground state degeneracy, is that while it is impossible
to rotate all spin variables at once, it is possible to
rotate them locally with little energy cost. There must be
then gapless excitations, with energy Ek ∼ k 2 for small k 2 ,
known in the theory of magnetism as magnons. These are the
non relativistic counterpart of the Goldstone excitations
of relativistic systems, with energy Ek ∼ |k|, that we en-
countered previously in the classical theory.
Whether or not the degenerate vacua are connected by con-
tinuous symmetry operations, the decoupling occurs in the
4
Operators acting only on finitely many spins at a time.
∑∞
5
The total spin would be given by the infinite sum S = a=1 Sa , which
requires a more careful treatment.

128
infinite volume limit in any local quantum field theory.
For example, labelling the vacua with continuous parameters
h, g, . . . , and given any two local field operators, we can
write

⟨h|A(x)B(0)|g⟩ = ⟨h|A(0)|f ⟩ ⟨f |B(0)|g⟩
f
∑∫
+ dPn eiPn ·x ⟨h|A(0)|n, Pn ⟩ ⟨n, Pn |B(0)|g⟩ (5.23)
n
and similarly

⟨h|B(0)A(x)|g⟩ = ⟨h|B(0)|f ⟩ ⟨f |A(0)|g⟩
f
∑∫
+ dPn eiPn ·x ⟨h|B(0)|n, Pn ⟩ ⟨n, Pn |A(0)|g⟩ (5.24)
n
In writing these equations we assumed that P |f ⟩ = 0 for a
vacuum state, and that this is a discrete eigenvalue. We
assumed also the existence of the translation group, with
operators T (x) = exp(−iP · x) such that

A(x) = T (x)A(0)T (x)−1

That is we do not consider the possibility that the trans-


lation group is broken spontaneously (the vacuum is not a
crystal so to speak).
But in a local field theory at space-like separation the
operators commute, [A(x), B(0)] = 0, and the momentum integrals
on the left hand sides go to zero as |x| → ∞, by a known
result of integration theory, at least for sufficiently reg-
ular (or smooth) matrix elements. Therefore all the vacuum
matrices ⟨h|A(0)|f ⟩, ⟨h|B(0)|f ⟩, ⟨h|C(0)|f ⟩, . . . , commute and in a
suitable basis are all diagonal

⟨h|A(0)|f ⟩ = aδhf , ⟨h|B(0)|f ⟩ = bδhf , etc.

In particular, no local perturbations can induce transitions


between the vacua in the diagonalizing basis, so these are
the relatively stable states of the system. The symmetry
generators are non local operators so they are not diagonal
in this basis, except for those leaving the chosen vacuum
invariant.

129
5.3 The Goldstone theorem
The existence of a symmetry group always implies a set of
conserved currents Jαµ , one for each symmetry generator, and
global charges

Qα (t) = dxJα0 (t, x), Q̇α = 0 (5.25)

In quantum field theory when a continuous symmetry is


spontaneously broken a massless particle is present in the
spectrum. For a rapid proof, consider the state

|k⟩ = e−ik·x J 0 (x, 0) |0⟩ dx (5.26)

We claim that this state has momentum k; for in fact

P |k⟩ =
∫ ∫
−ik·x
= e [P , J (x, 0)] |0⟩ = −i e−ik·x ∇J 0 (x, 0) |0⟩ dx =
0


= k e−ik·x J 0 (x, 0) |0⟩ dx (5.27)

But for k → 0, |k⟩ → Q |0⟩ which has zero energy because


HQ |0⟩ = QH |0⟩ = 0, and the energy can go to zero with momentum
k only for a massless particle state, which is our Goldstone
boson.
Here is another proof. Consider a smeared charge

QV (t) = dxJ 0 (x, t) (5.28)
V

and for a local operator the expression6


δa(t) = lim ⟨0|[QV (t), A]|0⟩ ̸= 0 (5.29)
V →∞

It is non zero in the case of symmetry breaking. It is also


conserved since by current conservation
I
dδa(t)
= − lim ⟨0|[J (t, x), A]|0⟩ dx = 0 (5.30)
dt V →∞ ∂V

6
An important theorem due to Kastler states that the limit exist for
any local operator A(x), if only the smearing is done also over the time
variable.

130
the last equality being true because the commutators of
local operators vanish for space-like separation. Inserting
a complete set of states each with given momentum, we may
then write
( )
∑ ∑
⟨0|J(0)|n⟩ ⟨n|A|0⟩ e−iEn t − ⟨0|A|n⟩ ⟨n|J(0)|0⟩ eiEn t δ 3 (Pn ) ̸= 0
n n

On the other hand the conservation theorem gives


( )
∑ ∑
⟨0|J(0)|n⟩ ⟨n|A|0⟩ e−iEn t + ⟨0|A|n⟩ ⟨n|J(0)|0⟩ eiEn t δ 3 (Pn )En = 0
n n

If δa(t) is not identically zero, there must be at least


one state |n⟩ for which ⟨0|J(0)|n⟩ ⟨n|A|0⟩ ̸= 0, in which case
δ 3 (Pn )En = 0. But the energy vanishes with vanishing momentum
only for zero mass states, so their presence is established.
Actually there must be single particle states with zero
mass, because we may take as operator A the quantum field
ϕn itself. In that case this is the transformation law

δan = ϵα ⟨0|[Qα , ϕn ]|0⟩ = − ϵα (Tα )nm ⟨ϕm ⟩0 (5.31)
m

where ⟨ϕm ⟩0 ≡ vm is the vacuum expectation value. Obviously


δa(t) could not be different from zero if the vacuum were
invariant.
This suggest a general criterion for the appearance of sym-
metry breaking in the vacuum: the existence of at least a
local or quasi-local operator OL such that

lim ⟨0|[QV T , OL ]|0⟩ =: [Q, OL ] ̸= 0 (5.32)


V →∞

where

QV T = J 0 (x, t)fT (t)dxdt (5.33)
V

and the test function fT (t) has support contained in the


interval [−T, T ] with unit integral. The charge QV (s) is the
formal value for fT (t) = δ(t − s).
Finally, we may use the powerful method of the effective

131
potential, which we showed has the symmetry of the classical
action. Then the identities hold
∑ ∂V
(T α )nm ϕm = 0 (5.34)
n,m
∂ϕn

Taking derivatives
∑ ∂ 2V ∑ ∂V
(T α )nm ϕm + (T α )nl = 0 (5.35)
n,m
∂ϕ n ∂ϕ l n
∂ϕ n

The symmetry is broken if there are non zero solutions, say


ϕn = vn , of the minimum condition
∂V
=0 (5.36)
∂ϕn
Computing (5.35) at ϕn = vn one finds
∑ ( ∂ 2V )
(T α )nm vm = 0 (5.37)
n,m
∂ϕ l ∂ϕ n |v
∑ α
We then discover that the vector m (T )nm vm′′, if non zero,
is a null eigenvector of the mass matrix Vnm , so for each
broken symmetry generator T α there is a zero mass pole in
the propagator, or a zero at zero momentum in the inverse
of the full propagator

(∆−1 )(0)nm (T α )ml vl = 0
m,l

′′
In fact the matrix Vnm (v) is the sum of all 1PI amputated
Feynman diagrams with two external lines carrying zero mo-
mentum into ∑ the diagrams. Not only this. We noted that the
fields πα = ϕn [Tα ]nm vm can be considered as the Goldstone
modes of the theory. Because the n − th order derivative of
V at ϕ = v is the full proper vertex at zero external mo-
mentum in the quantum theory, then what (5.37) says is that
the amplitude for a Goldstone boson to make transitions at
vanishing momentum into any other scalar vanishes. Taking
the derivative of (5.35) then setting ϕ = v and contracting
with (Tβ v) ⊗ (Tγ v) gives
( )
∂ 3V
(Tα v)n (Tβ v)m (Tγ v)l = 0 (5.38)
∂ϕn ∂ϕm ∂ϕl |v

132
which says that there is no amplitude for a transition
between three zero momentum Goldstone bosons. In particu-
lar, very soft Goldstone bosons are not emitted by external
Goldstone bosons lines in Feynman diagrams.

5.4 The case of broken gauge symmetries


It turns out that global symmetries, as opposed to local
symmetries, are the only ones that may be broken by non zero
expectation values of various fields, and as a consequence
do not become manifest in the properties of physical states.
This result is known as the Elitzur theorem.31 The new fact
that happens when a continuous global symmetry is gauged
is that the Goldstone particles, otherwise present in the
spectrum, are actually absent.
In preparation to more general possibilities, consider the
case of a symmetry group G spontaneously broken to a subgroup
H in a real representation of G. Then the manifold of all
vacua is parametrized by the coset space G/H of equivalence
classes of group elements differing by right multiplication
with elements of H: that is g1 and g2 are in the same
equivalence class iff g1 = g2 h, for at least one h ∈ H. G/H
is a group, with the natural multiplication induced by the
quotient operation. We noted in connection with Eq. (5.18)
that the Goldstone fields are “tangent” to the group manifold
G/H, so they too can be considered as “coordinates” on this
manifold. To see this most clearly it is only necessary to
write a general point ϕn in field configuration space (at
a given spacetime point x) as a G transformation to new
fields σn which are orthogonal to the Goldstone directions,
in the sense that

ϕn (x) = gnm (x)σm (x) (5.39)
n

where the new fields satisfy



σn (x)[T α ]nm ⟨ϕm ⟩0 = 0 (5.40)
n,m

for each broken symmetry generator T α . This can always


be achieved for any compact G in a real representation

133
(for a simple proof see Weinberg2 ). The fields σn are the
massive fields because they oscillate around the minimum
of the potential along directions orthogonal to the flat
directions. One can see this most clearly expanding the
potential (classical or effective, the argument is the same)
to quadratic order in the fluctuating fields σn = vn + hn (x),
vn = ⟨ϕm ⟩0 ,

1 ∂ 2V
V (v + h) = V0 + hn hm + · · ·
2 ∂σn ∂σm |v

The mass matrix has null eigenvectors Tα v (the flat direc-


tions mentioned above), therefore it is positive definite
on the subspace orthogonal to them, which is what the condi-
tion (5.40) requires to the fields hn . The Goldstone boson
fields are then hidden into the matrix gnm (x). For example,
in the O(N ) model we could always write the field as a
x-dependent rotation of a new field directed along the N
axis, say σn (x) = δnN σ(x)
ϕn = RnN (x)σ(x) (5.41)
The condition (5.40) is then [T α ]N m ⟨ϕm ⟩0 = 0. Since there are
N − 1 broken generators which are anti-symmetric matrices,
this means that ⟨ϕm ⟩0 = vδnN , so σ(x) = v +h(x) and the Goldstone
fields are in the N − 1 matrix elements RnN (x).
Now if in (5.39) we replace g with gh, h ∈ H, then (5.40) is
replaced with

σn (x)[h−1 T α h]nm ⟨ϕm ⟩0 = 0
n,m

which is still true because if Tα runs over the unbroken


generators then so do h−1 Tα h. Thus there is not a unique
g satisfying (5.39), (5.40), g and gh being equally good.
Right multiplication by elements of a subgroup H is an
equivalence relation, so the Goldstone fields are to be
thought as taking values in the coset group manifold G/H,
as already indicated.
Now it is clear why the Goldstone theorem breaks down when
a global symmetry broken by non zero expectation values of
scalar or other fields, becomes a gauge symmetry. It is

134
that Eq. (5.39) actually is a gauge transformation, under
which the Lagrangian of the gauge and matter fields is
invariant by definition. Therefore if we wish, we could use
the fields σn (x) everywhere and forget about the Goldstone
fields. The second condition (5.40) becomes then a choice
of gauge, imposed on the scalar fields rather than as a
condition on the gauge potentials. The choice is known in
this context as the unitary gauge. To be more precise, the
fields gnm (x) reappear in the gauge transformed potential,
which becomes Aµ → g −1 Aµ g + ig −1 ∂µ g. But then g(x) disappears
from the equations of motion, from the Feynman rules and
can be made to disappear from the path integral by a simple
change of variables. For example, if the O(N ) model is
gauged, then (5.41) is a gauge transformation and the matrix
Rnm (x) decouples from the Lagrangian. The unitary gauge in
this example is ϕ1 = ϕ2 = · · · ϕN −1 = 0 and σ is the only physical
field.
In general there is no standard way to put coordinates on
the coset space G/H as descriptors of the Goldstone fields.
One natural way to do this is to use the Cartan decomposition
of the Lie algebra of G relative to the subgroup H. One
takes a basis made of generators for the Lie subalgebra of
H, such that by definition

[Ti , Tj ] = ifijk Tk , i, j, k = 1, 2, . . . , dimH

plus additional broken generators Xa forming with Ti a basis


for Lie(G); since fija = −fiaj = 0, then [Ti , Xa ] = ifiab Xb and
[Xa , Xb ] = ifabj Tj + ifabc Xc . Each group element can now be written
in the form

g(x) = exp[iπ a (x)Xa ] exp[iξ(x)i Ti ]

The fields πa (x) are then a convenient parametrization of


the Goldstone fields, because the factor to the right is an
element of H.
We conclude by noting that after the condition (5.39), in
a theory without gauge field the derivatives take the form
∑ ( ) ∑
∂µ ϕn = gnm (x) ∂µ σm + (g −1 ∂µ g)ml σl = gnm (x)[Dµ σ]m
m m

135
and since the Lagrangian is invariant under G-transformation,
it will depend on g(x) only through the “covariant deriva-
tives” Dµ σ = (∂µ + g −1 ∂µ g)σ. This determines the way the Gold-
stone fields interact with themselves and the other fields
in the Lagrangian. For a detailed exposition of these re-
sults and applications to low energy hadron physics we defer
to the (vast) existing literature (a fairly complete treat-
ment including references can be found in the book of one
of the founders of the entire matter, namely S. Weinberg2 ).
However in a theory with gauge fields it is the gauge co-
variant derivative that matters, so we have now
∑ ( )
∇µ ϕn = gnm (x) ∂µ σm − i[g −1 Aαµ Tα g]ml σl + (g −1 ∂µ g)ml σl
m

We recognize the presence of the gauge equivalent potential


g −1 Aµ g + ig −1 ∂µ g, Aµ = Aαµ Tα , allowing the decoupling of the
Goldstone fields represented by the matrix gnm (x).
We leave it to the reader to verify that in a complex
representation of G with fields Ψ we can write the unitary
gauge in the matrix form
( )
Im Ψ† Tα ⟨Ψ⟩V AC = 0 (5.42)
As a hint, first write Ψ = ReΨ + iImΨ, then split Tα Ψ into
real and imaginary parts, which gives the real (reducible)
representation
[ ]
−ImTα ReTα
itα = (5.43)
−ReTα −ImTα
Finally apply condition (5.40) using this tα , and show that
it is identical to (5.42).

5.5 The abelian Higgs mechanism


The above discussion represents only half of the full story,
the other half being more interesting and far reaching. Con-
sider again the abelian model, this time minimally coupled
to the electromagnetic field, namely scalar electrodynam-
ics. The Lagrangian defining the model is
1 λ
L = − F 2 − (∇µ ϕ)∗ ∇µ ϕ − m2 ϕ∗ ϕ − (ϕ∗ ϕ)2 (5.44)
4 2
136
where ∇µ ϕ = ∂µ ϕ − ieAµ ϕ. The case m2 < 0 is the symmetry
breaking phase, or Higgs phase, with a set of minima at
m2
ϕ∗ ϕ = −
= v2 (5.45)
λ
We can trivially write

1
ϕ = √ (v + σ) exp(iθ)
2

with new fields σ and θ, in terms of which the Lagrangian


becomes
1 1 1 1
L = − F 2 − e2 (v + σ)2 (Aµ − ∂µ θ)(Aµ − ∂ µ θ) (5.46)
4 2 e e
1 1 λ
− ∂µ σ∂ µ σ − m2 (σ + v)2 − (v + σ)4
2 2 8
It is here that things become more interesting: the field
Aµ − 1e ∂µ θ is gauge equivalent to Aµ in a gauge theory, so we
can simply omit the gradient ∂θ term in the first line. We
then see the menu of the physical particle: the would be
Goldstone field disappears completely from the Lagrangian
and the gauge field at the same time acquires a mass
µ2A = e2 v 2 (5.47)
This phenomenon is known as the Higgs mechanism7 . Finally,
the field σ has mass m2σ = −m2 /2 = λv 2 /2 and has cubic and
quartic self-interactions. The number of local degrees of
freedom does not change, the θ field being compensated by
the extra polarization state of the newly massive vector
field. It is said that the Goldstone field has been “eaten”
by the gauge field into his longitudinal polarization state.
Surprisingly, the theory in unitary gauge looks non renor-
malizable due to the gauge field mass term, despite in the
form given by Eq. (5.44) it appears to be, and the gauge
transformations trivialized. The field redefinition leading
to (5.46)

1
ϕ → √ (v + σ)
2
7
It was actually and independently discovered by Kibble, Englert,
Brout, Guralnik and Hagen in the same years as Higgs described it.

137
is in fact the unitary gauge for the Higgs abelian model.
As we noted, the quantization in this gauge looks awkward.
We return then to anoth er parametrization
1
ϕ = √ (v + h(x) + iπ(x)) (5.48)
2
The Lagrangian becomes
1 1 1
L = − F 2 − (∂µ h + eAµ π)2 − (∂µ π − eAµ (v + h))2 (5.49)
4 2 2
1 2 λ
− m ((v + h)2 + π 2 ) − [(v + h)2 + π 2 ]2
2 8
This is still invariant under the gauge symmetry
1
δh = −ϵ(x)π, δπ = ϵ(x)(v + h),
δAµ = ∂µ ϵ (5.50)
e
so quantization will need a gauge fixing term, for example
1
Lgf = − (∂µ Aµ )2 (5.51)

Then the quadratic part is
1 1 1 1 1
L2 = − F 2 − µ2A A2 − (∂A)2 − (∂µ h)2 − (∂µ π)2 (5.52)
4 2 2ξ 2 2
1 2 2
− m h + µA Aµ ∂µ π (5.53)
2 h
where m2h = −m2 = λv 2 . We see that π is massless but couples
to the gauge field. Moreover, the interactions are all
of renormalizable types. A more clever gauge fixing term
is possible if we choose it to cancel the π − A derivative
coupling in L2 , namely
1
Lgf = − (∂µ Aµ − evξπ)2 (5.54)

With this choice the Lagrangian becomes a sum of squares
1 1 1 1 1
L = − F 2 − µ2A A2 − (∂A)2 − (∂µ h)2 − m2h h2 (5.55)
4 2 2ξ 2 2
1 1
− (∂µ π)2 − ξµ2A π 2 + · · · (5.56)
2 2
In this gauge, the abelian version of the Rξ gauge,
√ the
Goldstone field has the gauge dependent mass mπ = ξ µA and

138
the propagator ∆π = (p2 + ξµ2A )−1 . It vanishes as ξ → ∞ making
the Goldstone field non dynamical and decoupled from the
other fields. In fact, this limit is nothing else than the
unitary gauge.
The gauge field propagator is
( )
1 p µ pν
∆µν = 2 ηµν − (1 − ξ) 2 (5.57)
p + µ2A p
It has the usual UV behavior allowing the theory to pass
the power counting test of renormalizability. This is why
the gauge (5.54) is known as the Rξ -gauge.
Digression - The propagators8 for the theory with quadratic
part (5.52) can be written with some computations. They are
( )
1 pµ pν pµ pν
∆µν = 2 2
ηµν − 2 +ξ 4 (5.58)
p + µA p p

∆µπ = iξµA (5.59)
p4
1 ξµ2A 1
∆ππ = + , ∆hh = (5.60)
p2 p4 p2 + m2h
The propagators are simpler in Landau’s gauge, ξ = 0. There
are the three degrees of freedom of the massive vector field
and the “Higgs field” h(x). However the Goldstone boson is
really present and propagates as a massless mode. The theory
is also potentially infrared divergent.
Just as an illustration, in Landau gauge the Lagrangian has
two new vertices coming from the π − A interaction and the
A mass term; they correspond to graphs

µ pµ µ ν

The left diagrams has vertex factor −evpµ , the right one has
factor −iηµν e2 v 2 . Taken together they give a contribution to
the photon self-energy,

e2 v 2
−e2 v 2 (ηµν − pµ pν /p2 ) = − (ηµν p2 − pµ pν )
p2
8
These are defined as i times the contractions ⟨0|T Φr Φs |0⟩ of the various
fields. Also, iϵ-factors in denominators are omitted for simplicity.

139
µ ν
−i
p2

Figure 5.1: The Goldstone boson pole

The presence a the pole at p2 = 0 is clearly due to the


propagation of the Goldstone particle, as illustrated in
figure (5.1), which is responsible for the mass term of the
gauge field. In fact the propagator in Landau gauge is

1
∆µν (p) = (ηµν − pµ pν /p2 )
p2 (1 − π(p2 ))

so a pole in π(p2 ) = −e2 v 2 p−2 + · · · clearly produces the mass.


This illustrates some intricacy in the role of the Goldstone
particle even in theories where it is actually absent from
the physical spectrum.

5.6 The general Higgs mechanism


We turn now to the general case of a compact simple gauge
group G broken down to a subgroup H by the vacuum expec-
tation values of a multiplet of scalar fields in a real
representation of G. We are interested in (i), how masses
are generated for the gauge fields at the expense of the
Goldstone bosons, (ii), the quantization problem.
The Lagrangian is the sum of terms from the gauge, scalar
and fermion sectors, plus gauge fixing and ghost terms. We
concentrate on the gauge and scalar sector, with Lagrangian
1 a aµν 1
L = − Fµν F − Dµ ϕT Dµ ϕ − V (ϕ) (5.61)
4 2

Dµ ϕ = ∂µ ϕ − igAaµ Ta ϕ (5.62)

We are using a matrix notation, with ϕT denoting transpo-


sition, V (ϕ) is a symmetry breaking (tree level) potential,

140
Ta the scalar representation with imaginary, anti-symmetric
matrices. Let the constant vector v be a minimum of the
potential, the lowest order approximation to the vacuum
expectation ⟨ϕ⟩V AC in perturbation theory. Then

∂V
=0
∂ϕn |v

and the matrix of second order derivatives is positive semi-


definite. There is also a stability group H for v, con-
sisting of elements h ∈ H such that h · v = v. As usual we
shift the field

ϕ=v+φ

and plug it into LS ; the covariant derivatives are

Dµ ϕ = ∂µ φ − igAaµ Ta φ − igAaµ Ta v, Dµ ϕT = ∂µ φT + igAaµ φT Ta + igAaµ v T Ta

so the kinetic term takes the form

1 1
− Dµ ϕT Dµ ϕ = − (∂µ φT + igAaµ φT Ta )(∂µ φ − igAaµ Ta φ)+
2 2
1 1 1
− igD φ Aµ Ta v + igAaµ v T Ta Dµ φ − g 2 v T Ta Tb vAaµ Abµ (5.63)
µ T a
2 2 2
There are two relevant things here: one is the generation
of a mass term for the gauge fields, with mass matrix

m2ab = g 2 v T Ta Tb v = −g 2 a
Tnm b
Tnl vm vl (5.64)
nml

The generators Ta being imaginary, the mass matrix is symmet-


ric and positive semi-definite, with non zero eigenvalues ∑ of
order µ ∼ gv. On the other
∑ hand, if the generator T = ca Ta
is unbroken, so that Tnm vm = 0, then

m2ab cb = 0
a

Thus we still have a massless gauge field for each unbroken


generator. The case of complex unitary representations can

141
be handled as in connection with Eq. (5.43), with the result
m2ab = 2 ⟨ψ⟩†0 Ta† Tb ⟨ψ⟩0 = ⟨ψ⟩†0 {Ta , Tb } ⟨ψ⟩0 (5.65)
provided the kinetic term be taken in the form

K = −(∇µ ψ)† ∇µ ψ

The second relevant thing is the gauge symmetry, still


present in the form

δAµ = ∇µ ϵ, δφn = ig ϵa [Ta ]nm (vm + φm ) (5.66)
which must be fixed before quantization. For example, the
gauge field Lagrangian after the Higgs mechanism, including
a standard Lorentz gauge fixing term, becomes
1 a aµν 1 2 a bµ 1
L = − Fµν F − mab Aµ A − (∂µ Aµa )2 (5.67)
4 2 2ξ
which gives the propagator
[ ]
∆µa;νb = −i (p2 + m2 )−1 [ηµν − (1 − ξ)pµ pν (p2 + ξm2 )−1 ] ab (5.68)
It looks renormalizable in spite of the presence of the
mass, and is interesting that it converges to the standard
propagator of the massive spin one field in the limit ξ → ∞,
showing that this is the theory in the unitary gauge.
Looking at Eq. (5.63) the scalar Lagrangian contains a
quadratic cross term proportional to φT ∂µ Aµa Ta v together with
its transpose. To diagonalize the Lagrangian it is conve-
nient to adopt another gauge fixing, known as the Rξ -gauge,
designed to cancel the cross term, as follows
1 ( ∑ )2
Lgf = ∂µ Aµa − igξ φn (Ta )nm vm (5.69)

The gauge propagator is unaffected by this, however the
scalar mass pick up a new term, becoming
′′

2
Mnm = Vnm (v) − ξg 2 (Ta v)n (Ta v)m (5.70)
a

This has an eigenvalue m2 = ξµ2 for each eigenvalue µ2 of the


gauge bosons mass matrix, that is for each broken generator.
Evidently the Goldstone bosons have become massive, although

142
with gauge dependent masses. They are infinitely massive in
the unitary gauge, giving a different explanation for their
decoupling.
The scalar propagator ∆nm (p), is the inverse of the matrix
′′

∆−1
nm = p δnm + Vnm − ξg
2 2
(tα v)n (tα )m
α

with a simple computation it takes the form9


′′

∆nm = (p2 + V )−1 2 −2
nm + ξg p (tα v)n (tβ v)m (p2 + ξm2 )−1
αβ (5.71)
α,β

Finally we must supply ghosts. With the given gauge fixing


the FP operator is

ca → ∂ µ ∇µ ca + ξg 2 [Ta ]nl [Tb ]nm (v + φ)m vl cb (5.72)
nml
a
where c is the ghost field. The ghost Lagrangian is then

Lgh = c̄a ∂ µ ∇µ ca + ξg 2 c̄a [Ta ]nl [Tb ]nm (v + φ)m vl cb (5.73)
nml
The free quadratic part is

L0gh = c̄a ∂ µ ∇µ ca + ξg 2 c̄a [Ta ]nl [Tb ]nm vl vm cb
nml
= c̄ ∂ ∇µ c +
a µ a 2 a b
ξmab c̄ c (5.74)
and has mass term Mab 2
= −ξg 2 [Ta v] ⊗ [Tb v] = ξm2ab . Therefore it
has an eigenvalue ξµ for each eigenvalue µ2 of the gauge
2

bosons mass matrix.


The scalar-Yang-Mills Lagrangian plus Lgf + Lgh can be used,
supplemented by fermion fields with Yukawa or other renor-
malizable interactions, to get a sensible quantum field the-
ory for spontaneously broken gauge symmetry. It has been
show by ’t Hooft (1971), Zinn-Justin (1972), Lee (1972,1973)
and others that the gauge symmetry, even in the broken phase,
still constraint the infinities in such a way that all UV
divergences can be cancelled by renormalization of masses,
fields and coupling constants, as with any other renormal-
izable field theory.
9
It seems there is a typo in Weinberg’s text2 inasmuch as the full
′′
matrix M 2 takes the place of V .

143
5.7 The Standard Model Higgs
This is an excellent example to illustrate the symmetry
breaking mechanism in a phenomenological very important
physical theory. We are alluding to the Glashow-Weinberg-
Salam electroweak theory, of course. In this theory the
symmetry breaking is entrusted to a doublet of SU (2) scalars,
the Higgs field
[ + ]
ϕ ϕ1 + iϕ2
Φ= 0 , ϕ+ = √ (5.75)
ϕ 2
whose dynamics is governed by the Lagrangian
λ
L = −[Dµ Φ]† Dµ Φ + µ2 Φ† Φ − (Φ† Φ)2 , µ2 > 0 λ > 0 (5.76)
2
The covariant differentials are given by

Dµ Φ = ∂µ Φ − igAaµ ta Φ − ig Bµ Y Φ (5.77)
where ta = σa /2 are the SU (2) generators, Y = 1/2 × identity
the weak hypercharge and, for future records, q = t3 + Y the
unbroken electric charge generator in units of e = 0.302 (the
positron charge)10 . Then ϕ+ is electrically charged in the
way the notation suggests.
With this potential the symmetry is broken already at a
classical level by a vacuum expectation value which we may
choose in the form after a global SU (2) rotation
[ ]
1 0 µ2
Φ0 = √ , v2 = (5.78)
2 v λ
With this the global gauge symmetry is spontaneously broken.
Note that no generator ta leaves the vacuum value invariant,
but qΦ0 = (t3 + Y )Φ0 = 0, so q is unbroken. We then set
1
ϕ0 = √ (v + H(x) + iϕ3 (x)) (5.79)
2
The local gauge transformations under which L is invariant
are still there, with arbitrary parameters θa (x) and θ4 (x),

δΦ = igθa (x)ta Φ + ig θ4 (x)Y Φ (5.80)
10
Part of these definitions is conventional and can be changed, as can
be observed by reading different authors. Weinberg for example uses
q = t3 − Y , but then
√ Y = −1/2 and nothing changes. His L is one-half our,
but there is no 2 factor in the vacuum, etc.

144
or in terms of components
ig 1 i ′
δϕ+ = (θ − iθ2 )ϕ0 + (gθ3 + g θ4 )ϕ+ (5.81)
2 2

ig 1 i ′
δϕ0 = (θ + iθ2 )ϕ+ − (gθ3 − g θ4 )ϕ0 (5.82)
2 2
These must be fixed by hand via the gauge fixing procedure.
In fact it is impossible to spontaneously break a local
gauge symmetry, a result known as the Elitzur’s theorem.
Standard results of the EW theory are the following relations
between the electric and weak charges2

e = g sin θ = g cos θ (5.83)
and the weak angle

g g
sin θ = √ , cos θ = √ , (5.84)
g2 + g′2 g2 + g′2
It will be convenient to rotate the gauge parameters with
the mixing angle to new parameters θZ , θA , as follows
θ3 = θZ cos θ + θA sin θ, θ4 = −θZ sin θ + θA cos θ (5.85)

where tan θ = g /g by definition. In this way
′ ′ ′
gθ3 + g θ4 = (g sin θ + g cos θ)θA + ((g cos θ − g sin θ)θZ

g2 − g 2
= 2eθA − √ θZ (5.86)
g2 + g′2

′ ′

gθ3 − g θ4 = (g cos θ + g sin θ)θZ = − g 2 + g ′ 2 θZ (5.87)
Thus ϕ+ transform as a positively electrically charged field
and ϕ0 as a neutral field.
The unitary gauge condition of Eq. (5.42) is easily imple-
mented for the Higgs field with vacuum value (5.78). It
correspond to
ϕ1 = ϕ2 = ϕ3 = 0 (5.88)

or equivalently ϕ+ = 0 and ϕ0 = (v + H)/ 2 and real. This
is the only physical field. But we have to keep in mind
that there is an unbroken generator, the electric charge,

145
so the corresponding U (1) symmetry remains unbroken, and
must be fixed too. The massless U (1) gauge field is the
electromagnetic field, a mix of A3µ and Bµ ,

Aµ = A3µ sin θ + Bµ cos θ (5.89)


and can be fixed by the usual Lorentz gauge condition
∂µ Aµ = 0, for example. The other mix is the famous Z 0 parti-
cle, responsible for the existence of the neutral current
interactions
Zµ = A3µ cos θ − Bµ sin θ (5.90)
The point of the mixing (5.89), (5.90) is that it makes the
gauge boson mass term diagonal in the fields A1µ , A2µ , Zµ and
Aµ . From the kinetic term of the Higgs Lagrangian we get
the mass generation
[ ]† 2 [ ]
1 0 ′ 0 1
− igAµ ta Φ + ig Bµ Y
a
≡ − m2ab Aaµ Abµ − ma4 Aaµ B µ
2 v v 2
1
− m244 Bµ B µ (5.91)
2
A little computation gives the diagonal expression
g2v2 [ 1 2 ] v2 ′
− (Aµ ) + (A2µ )2 − (g 2 + g 2 )Zµ Z µ (5.92)
8 8
In terms of the charged vector fields
A1µ + iA2µ A1µ − iA2µ
Wµ = √ , Wµ† = √ (5.93)
2 2
we may write
g2v2 ∗ v2 ′
− Wµ Wµ − (g 2 + g 2 )Zµ Z µ (5.94)
4 8
The masses are then
gv v√ 2
MW = , MZ = g + g ′ 2 , MA = 0 (5.95)
2 2
We now exploit fully the unitary gauge, starting with the
Higgs Lagrangian. The covariant derivatives are

g2 − g 2 g
Dµ ϕ = ∂µ ϕ − ieAµ ϕ − i √
+ + +

Zµ ϕ+ − i √ Wµ∗ ϕ0 (5.96)
2 g2 + g 2 2

146
i√ 2 g
Dµ ϕ0 = ∂µ ϕ+ + g + g ′ 2 Zµ ϕ0 − i √ Wµ ϕ+ (5.97)
2 2

Setting then ϕ+ = ϕ3 = 0 and ϕ0 = (v + H)/ 2 in L , we get
1 g2 1 ′
LH = − ∂µ H∂ µ H − Wµ∗ W µ (v + H)2 − (g 2 + g 2 )Zµ Z µ (v + H)2
2 4 8
1 2 λ
+ µ (v + H)2 − (v + H)4 (5.98)
4 8
from which we read the mass of the Higgs MH2 = µ2 = λv 2
and its coupling to the vector bosons. The absence of
electromagnetic coupling means that H is a neutral field,
and there are self-interactions. The coupling to the Z can
be written in different ways, e.g.

′ g2 e2
g2 + g 2 = =
cos2 θ (sin θ cos θ)2

We now have to consider the ghost sector in the unitary


gauge. To disentangle the broken from the unbroken gen-
erator, we use Eq.s (5.81), (5.82) and (5.87). The gauge
variation of the three gauge fixing conditions ϕ1 = ϕ2 = ϕ3 = 0
are
g g
δϕ1 = (v + H)ϵ2 , δϕ2 = (v + H)ϵ1 (5.99)
2
√ 2
g2 + g′2
δϕ3 = (v + H)ϵ3
2
where ϵZ has been renamed ϵ3 . The FP operator kernel is then
Mij (x, y) = Gij (1 + H(x)/v)δ 4 (x − y) (5.100)
where G is the matrix
 
0 gv
2
0
G= 2 0 
gv

v
√ 0 ′2
(5.101)
2
0 0 2 g +g
It is an ultralocal operator (no partial derivatives) so
its determinant interpreted in a functional sense can be
exponentiated, using the identity (see Weinberg1 p.392-393)
( ∫ )
4 4
Det[Mij ] = exp δ (0) log det Mij (x) d x

147
where Det[Mij ] denotes the functional determinant, det Mij the
matrix determinant. In this approach we would get, up to an
irrelevant multiplicative constant (namely exp δ 4 (0) ln det[Gij ]),
[ ∫ ]
det M = exp i −3iδ (0) log(1 + H(x)/v)d x
4 4
(5.102)

The UV divergent factor is to be interpreted as the limit


δ 4 (0) = lim Ω−1 , where Ω is a space-time volume shrinking to
a point in the limit.
The functional within ∫square brackets is the ghostless ac-
tion to be added to LH as a consequence of the gauge
fixing. Thus in the unitary gauge the ghosts can be re-
moved but they do not entirely decouple, there remain an
additional quartically divergent interaction containing all
powers of the Higgs field, i.e.

L = −3iδ 4 (0) log(1 + H(x)/v) (5.103)

The same result can be obtained using the Feynman rules


derived from the ghost Lagrangian appropriate to this case,

Lgh = c̄i Gij cj (1 + H(x)/v) (5.104)

simply by summing all loops with N external Higgs lines for


N = 1, · · · , ∞. The factor 3 is produced because there are three
ghosts at each vertex, and the factor δ 4 (0) because the ghost
propagator “trivially” does not depend on the momentum, and
the loop integral is just

d4 p
−i 4
= −iδ 4 (0)
(2π)

The effect of the interaction L is twofold: it helps
to cancel δ 4 (0)-singularities coming from other sources in
perturbation theory, and it makes the scattering operator
independent on how the scalar field is defined. This has
been shown by Salam et al34 and Gerstein et al.35 This
conclude our treatment of the unitary gauge for the Higgs
field.
Digression - The matrix (5.99) is a continuous version of
the Kronecker product of a 3 × 3 matrix times the identity

148
in function space. Therefore
∏ ∏
Det[Mij ] = det[Gij (1 + H(x)/v)] = det[Gij ](1 + H(x)/v)3
x x

Taking the logarithm and using


∑ ∑ ∫ ∫
−1 −1 4 4
f (x) = lim Ω Ωf (x) = lim Ω f (x)d x = δ (0) f (x)d4 x
x x

where Ω is a small space-time volume, the result (5.102)


follows.

Figure 5.2: Diagrams contributing to the effective La-


grangian (5.104). Internal lines are ghost propagators,
external lines Higgs particles.

149
Chapter 6

The renormalization group

One of the most significant properties of gauge theories


is the excellent ultraviolet behavior. To study this one
needs (a posteriori) to trace out the evolution of the theory
parameters as the renormalization scale changes. The tool
needed is the renormalization group.

6.1 Bare and renormalized functions


We study the paradigmatic example of quartic scalar self-
interaction, with fundamental bare Lagrangian
1 1 λ0
Lϵ = − (∂ϕ0 )2 − m20 ϕ20 − ϕ40 (6.1)
2 2 4!
where the ϵ subscript is to remind us that the regularized
theory is defined in d = 4 − ϵ dimensions. The renormalized
Lagrangian is Lϵ expressed in terms of renormalized fields,
masses and coupling constant, as follows
1 1 Zλ λ 4
L = − Z(∂ϕ)2 − Zm m2 ϕ2 − ϕ (6.2)
2 2 4!
with one Z for each operator in the theory and where, since
we have simply rewritten (6.1),

ϕ0 = Zϕ, m20 = m2 Zm Z −1 ≡ m2 Z̃m , λ0 µ−ϵ = λZλ Z −2 ≡ λZ̃λ (6.3)

A mass parameter µ has been introduced to keep λ dimension-


less, since in 4 − ϵ dimensions the bare coupling has mass
dimension ϵ. On which parameters are the Zi ’s dependent

150
is generally related to the renormalization scheme and how
infinities are actually subtracted. We use a mass inde-
pendent renormalization scheme which can also we applied
to massless fields and avoid infrared divergences as m → 0
in the coupling parameter (see Weinberg,2 Vol. 2, Ch. 18
for explanations on how this comes about). For example, we
could require the inverse propagator to satisfy
Γ2 (p)
Γ2 (p)|p2 =µ2 = µ2 , =1 (6.4)
dp2 |p2 =µ2
and the four-point proper vertex to satisfy
1
Γ4 (pi ) = −λ at pi · pj = µ2 (δij + ) (6.5)
3
where µ is an arbitrary mass scale. We will use instead
the minimal subtraction scheme, denoted MS, wherein the
renormalized proper vertices (see below) are defined by
I
n 1 dϵ n
Γren = Γ (ϵ) (6.6)
2πi ϵ
In both scheme, the Zi are functions Zi (λ, ϵ) with no mass
dependence. This is because they make sense at m = 0, but
could only depend on m through logarithms terms log(m/µ) in
loop effects, which are singular as m → 0.
The renormalizability of λϕ4 is the statement that one
can choose the bare parameters and fields λ0 , m0 , ϕ0 (or
equivalently the Zi ’s) as functions of the renormalized
ones and the cutoff ϵ in such a way that in the limit
ϵ → 0 all correlation functions of the renormalized fields
are finite and satisfy the general requirements of quantum
field theory. Now the (time ordered) connected correlation
functions are related to the bare functions by
Gn (xi ; λ, m, µ) = Z(λ, ϵ)−n/2 Gn0 (xi ; λ0 , m0 , ϵ) (6.7)
which entails that
Z(λ, ϵ)n/2 Gn (xi ; λ, m, µ) (6.8)
is independent on the scale µ. Similar considerations can
be done for the proper vertices, denoted here Γn (pi , λ, m, µ),
which are obtained by Fourier transforming the connected am-
putated functions, those Gn whose external propagators have

151
been stripped away and the momentum conserving δ-function
factored out.
Since Γ2 (p, λ, m, µ) = Z(λ, ϵ)Γ20 (p, λ0 , m0 , ϵ) is the inverse of the
full propagator, and the bare amputated Green functions are
defined by

n
T
G0 (p1 , . . . , pn ) = Γ20 (pi , λ0 , m0 )G0 (p1 , . . . , pn ) (6.9)
i=1

one obtains easily the scaling relations dual to (6.7)

Γn (pi ; λ, m, µ) = Z(λ, ϵ)n/2 Γn0 (pi ; λ0 , m0 , ϵ) (6.10)

which means that

Z(λ, ϵ)−n/2 Γn (pi ; λ, m, µ) (6.11)

are all independent on the scale µ. Assuming Γ1 (p, λ, m, µ) = 01 ,


which is actually independent of p, the functional Γ[ϕ]
∑ 1 ∫ d4 p1 d4 pn n
Γ[ϕ] = 4
· · · 4
Γ (p1 , . . . , pn )(2π)4 δ(p1 + · · · + pn )ϕ(p1 ) · · · ϕ(pn )
n≥2
n! (2π) (2π)

is called the effective action. Its relation with the bare


regularized action is obviously

Γ[ϕ] = Γ0 [ϕ0 ] = Γ0 [ Zϕ] (6.12)

Formally, the canonical mass dimension of the proper


vertices is easily seen to be δ = d − n(d − 2)/2, which becomes
4 − n in four dimensions. This implies the scaling relations

Γn (et p1 , . . . , et pn , m, λe(4−d)t , µ) = e(d−n(d−2)/2)t Γn (p1 , . . . , pn , e−t m, λ, e−t µ)

or in d = 4

Γn (et p1 , . . . , et pn , m, λ, µ) = e(4−n)t Γn (p1 , . . . , pn , e−t m, λ, e−t µ) (6.13)

The much more interesting relations connecting the left


hand side to the right hand side at the same scale µ will
be deduced shortly.
1
This is the Fourier transform of the vacuum expectation value
⟨0|ϕ(x)|0⟩ = 0, graphically given by the sum of tadpole graphs.

152
6.2 The renormalization group equations
The observation following Eq. (6.11) can be translated into
the statement that at fixed bare parameters λ0 , m0 and ϵ


µ Z(λ, ϵ) Γ (pi ; λ, m, µ)
−n/2 n
=0 (6.14)
∂µ λ0 ,m0 ,ϵ

From Eq. (6.3) we can solve for λ = λ(λ0 µ−ϵ , ϵ), so at fixed
bare parameters m and λ become functions of µ. From (6.14)
we then get

∂ n ∂
µ Γ (pi ; λ, m, µ) + β(λ, ϵ) Γn (pi ; λ, m, µ)+
∂µ ∂λ
∂ n
+ γm (λ, ϵ)m Γ (pi ; λ, m, µ) − nγ(λ, ϵ) Γn (pi ; λ, m, µ) = 0 (6.15)
∂m
where
∂λ ∂ log Z̃λ
β(λ, ϵ) = µ = −ϵλ − λµ (6.16)
∂µ |λ0 ,m0 ,ϵ ∂µ |λ0 ,m0 ,ϵ

∂ log m 1 ∂ log Z̃m


γm (λ, ϵ) = µ =− µ (6.17)
∂µ |λ0 ,m0 ,ϵ 2 ∂µ |λ0 ,m0 ,ϵ

and we defined the anomalous dimension


1 ∂ log Z 1 ∂ log Z
γ(λ, ϵ) = µ = β(λ, ϵ) (6.18)
2 ∂µ |λ0 ,m0 ,ϵ 2 ∂λ
The limit ϵ → 0 exists because we are dealing with a renor-
malizable theory. Similar equations hold for the connected
Green functions

∂ n ∂
µ G (pi ; λ, m, µ) + β(λ, ϵ) Gn (pi ; λ, m, µ)+
∂µ ∂λ
∂ n
+ γm (λ, ϵ)m G (pi ; λ, m, µ) + nγ(λ, ϵ)Gn (pi ; λ, m, µ) = 0 (6.19)
∂m
From (6.3) we see that the beta-function can be written as
∂ log Z̃λ ∂ log Z̃λ
β(λ, ϵ) = −ϵλ − λµ = −ϵλ − λ β(λ, ϵ) (6.20)
∂µ |λ0 ,m0 ,ϵ ∂λ

153
Eq. (6.15) and (6.19) are the homogeneous renormalization
group equation for the proper vertices (first proposed by
Weinberg, ’t Hooft and Zinn-Justin around 1973); it differs
from the Callan-Symanzik equation in that it is homogeneous
and admits a zero mass limit without recourse to asymptotic
expansions in the momenta. As we will see, combined with
dimensional analysis they are able to determine the asymp-
totic behavior of the theory at very large (and often very
small) momentum and energy.
The renormalization constants have expansions (in the MS
scheme)


Z̃n (λ)
Z̃(λ, ϵ) = 1 + (6.21)
n=1
ϵn

and we have made explicit that β, γ, γm are independent of


m and µ. Finally, the infinitely many RG equations can be
compactly written as a single equation for the effective
action,
( ) ∫
∂ ∂ ∂ δΓ d4 p
µ + β(λ) + γm (λ)m Γ[φ] = γ(λ) φ(p) (6.22)
∂µ ∂λ ∂m δφ(p) (2π)4
or for the generating functional of connected functions
W [J] = −i log Z[J],
( ) ∫
∂ ∂ ∂ δW d4 p
µ + β(λ) + γm (λ)m W [J] = −γ(λ) J(p) (6.23)
∂µ ∂λ ∂m δJ(p) (2π)4

6.3 Renormalization group trajectories


Eq. (6.15) expresses the fact that µ is an arbitrary pa-
rameter, and that any change of µ can be compensated by a
change of λ, m and the scale of ϕ. Considering a continuous
change µ → µ̄(t) = et µ we have µ∂µ = ∂t ; consider the solution
of the differential equations
dλ̄ dm̄
= β(λ̄(t)), = mγm (λ̄(t)) (6.24)
dt dt
with initial data

λ̄(0) = λ, m̄(0) = m

154
Then the RG equations tell us that
d [ ]
Z(t)−n/2 Γn (pi ; λ̄(t), m̄(t), et µ) = 0 (6.25)
dt
where

d log Z(t)
= γ(λ̄(t)) (6.26)
dt
or

Z(t)− 2 Γ(n) (pi ; λ̄(t), m̄(t), et µ) = Γ(n) (pi ; λ, m, µ)


n
(6.27)

Thus the theory is invariant under a change of scale if a


change in µ is accompanied by a change in λ and m along
the RG trajectories. If the knowledge of the Γn throughout
momentum and parameter space gives the solution of a quantum
field theory, then we see that not just one value of the
parameters specify the theory, but an entire RG trajectory
does so.
We now show that the functions λ̄(t), m̄(t) are the running (or
effective) coupling and mass, respectively, as measured at
a momentum scale p ∼ et µ. Indeed using dimensional analysis
(see Eq. (6.13)) we get first of all

Γ(n) (et pi , λ, m, µ) = e(4−n)t Γ(n) (pi , λ, e−t m, e−t µ) (6.28)

and combining with (6.27)

Γ(n) (et pi , λ, m, µ) =
( ∫ t )
= exp (4 − n)t − n ds γ(λ̄(s)) Γ(n) (pi , λ̄(t), e−t m̄(t), µ) (6.29)
0

We see two main effects: (i) the exponent receives a cor-


rection and changes from (4 − n)t to (4 − n)t − nδ(t), where δ(t)
∫ t ∫ λ(t)
γ(x)
δ(t) = ds γ(λ̄(s)) = dx
0 0 β(x)

(ii) the breaking of scale invariance of the massless the-


ory, because of the dependence of λ̄ on the scale of the
momenta: the behavior of the theory at momenta of order
et p is determined by the running coupling λ̄(t). Only if the

155
beta function vanishes, β(λ) = 0, can we have exact scale in-
variance, albeit with a modified (or anomalous) dimension.
To be more precise, for a massless theory exact dilatation
invariance would imply the Ward identity
( n )
∑ ∂ ∑n
pi · + (n − 4) Γ (p1 , . . . , pn ) = 0,
(n)
pi = 0 (6.30)
i=1
∂pi i=1
while (6.29) actually implies the true Ward identities (take
the t-derivative of (6.29) and set t = 0 afterward)
( n )
∑ ∂ ∂Γ(n)
pi · + (nd − 4) Γ(n) (p1 , . . . , pn ) = β(λ) (6.31)
i=1
∂p i ∂λ
where d = 1+γ(λ) can be interpreted as the anomalous dimension
of the renormalized field. Clearly only at a fixed point
can we have exact scale invariance. By the way, if we put

φt (p) = e(d−4)t φ(e−t p) or φt (x) = edt φ(et x)

then for example

dφt ∂φ
= δφ(p) = (d − 4)φ − p ·
dt |0 ∂p

is the anomalous variation of the field under dilatations,


and (6.31) becomes a formula for the variation of the ef-
fective action

dΓ[φt , λ] δΓ 4 ∂Γ[φ, λ]
= (d + x · ∂)φ(x) d x = −β(λ) (6.32)
dt |0 δφ(x) ∂λ
If the beta function is known, we obtain the running coupling
from the equation
∫ λ̄(t)
dx
=t (6.33)
λ β(x)
Once we know λ̄(t) from the RG equations we get
[∫ t ] [∫ ]
λ̄(t)
γm (x)
m̄(t) = m exp ds γm (λ̄(s)) = m exp dx (6.34)
0 λ β(x)
and
[∫ t ] [∫ ]
√ λ̄(t)
γ(x)
Z(t) = exp γ(λ̄(s))ds = exp dx (6.35)
0 λ β(x)

156
6.4 Asymptotic behavior and fixed points
Look at Eq. (6.33): for β(λ) > 0 the effective coupling
continues to increase as t → ∞ until it meets one of the
situations: (a) if the integral (6.33) converges
∫ ∞
dx
<∞
λ β(x)

then the coupling will be infinite at a finite energy, a


behavior first conjectured by Landau for the electric charge
in QED. From another point of view this is also known as the
triviality problem, since in this situation the only way to
make sense of the theory at any scale is to put λ = 0; (b) the
integral (6.33) diverges so the coupling continues to grow
with energy and becomes infinite at infinite energy; or (c),
it continues to grow until it reaches a fixed point λ∗ of the
beta-function, for which β(λ∗ ) = 0; and it will decreases to
λ∗ if β(λ) < 0. Therefore a fixed point for which the slope

β (λ∗ ) < 0 will attract the couplings in its neighborhood as
t → ∞, and for that reason it is called an ultraviolet (UV)

stable fixed point. Conversely, if β (λ∗ ) > 0, then λ∗ will
attract in the infrared (IR), as t → −∞, so for that reason
it is called an infrared (IR) stable fixed point. These
possibilities are illustrated in Fig. (5.1).

Figure 6.1: UV stable fixed point (left); IR stable fixed


point (right)

The origin λ = 0 is always a fixed point, since without


interaction there is no renormalization. However, whether
the origin is IR or UV stable depends on the sign of the
beta function in a neighborhood of the origin.

157
If β(λ) = b0 λn + · · · near the origin and b0 > 0 the theory is IR
free, i.e. the coupling decreases with decreasing energy;
if b0 < 0 then the theory is UV free and the coupling decreases
at higher energies. For an asymptotically free theory in
the IR perturbative calculations of amplitudes cannot be
trusted beyond a sufficiently high energy.

Figure 6.2: Asymptotically free theory (left); IR free


theory (right)

Thus for a theory with a single coupling the situation


is pretty simple. For theories with multiple couplings
the behavior in the ultraviolet or infrared can be more
complicated.

6.4.1 Infrared free theory


This the case of Fig. 5.2 right: the coupling continues to
increase and diverges at finite or infinite energy, unless
it meets an UV stable fixed point λ∗ . Then close to λ∗ we
can write

′ ′ dβ
β(λ) ≃ |β |(λ∗ − λ), β (λ) =

which implies the rate of approach to the fixed point

λ(t) →t→∞ λ∗ − c exp(−|β |t)

Also, for large t

m(t) → m exp[γm (λ∗ )t], Z(t) → exp[γ(λ∗ )t]

158
From eq. (6.29) it follows that the UV behavior of the proper
vertices is given by
( )
Γn (et pi , λ, m, µ) −→ e(4−n−nγ(λ∗ ))t × Γn pi , λ∗ , e−t(1−γm (λ∗ )) m, µ (6.36)
t→∞

If all masses vanish and the theory sits at a fixed point,


then it is also exactly scale invariant but with a different
scale dependence form the naive scaling, controlled by the
anomalous dimension γ(λ∗ ). If γm (λ∗ ) < 1 the mass also de-
creases to zero and becomes less important at high energy,
as was to be expected. However, the fixed point probably
lies outside the reach of perturbation theory, so the above
asymptotic UV behavior is not very useful. In the IR the
proper functions simply approach those of a free field the-
ory.
A prime example of a IR free theory is QED. To one loop
order one has2
e3
β(e) = 2
+ O(e5 ) (6.37)
12π
therefore Eq. (6.33) gives

t
e−2 − e(t)−2 =
6π 2

or, using t = log µ̄/µ,


e e(µ)
e(t) = or e(µ̄) = (6.38)
1 − 6π
e2 t
2 1− e2 (µ)
6π 2
log(µ̄/µ)

The conventionally renormalized charge is er = e(m) = 4π/137 ≃
0.3, where m is the electron mass; if we were to take
seriously the one loop result, we would encounter a pole
(the Landau pole) at the huge energy

µ = m exp(6π 2 /e2r ) = 1.6 × 10280 Mev

Conversely, in the infrared as t → −∞ the effective charge


simply vanishes, e(t) → 0. This is usually interpreted as a
screening effect: as we go to large distances the vacuum
polarization screens the charge, and as we go to smaller
2
From the pole term of Z3 = 1 − e2 /6π 2 ϵ in dimensional regularization.

159
distances we penetrate the virtual cloud and see the bare
charge. In fact, we can identify the bare coupling as the
effective charge at the scale t = 1/ϵ, for in fact e20 = e2 /Z3
(see the footnote).

6.4.2 UV free theory


If the theory is asymptotically free in the ultraviolet then
as t → ∞ the coupling is attracted towards the trivial fixed
point λ = 0, also known as the Gaussian fixed point. In this
case everything is under perturbative control. Specifi-
cally, if the theory has β(g) = −b0 g 3 /2 + · · · with b0 > 0 (QCD
with a coupling g for example) then solving (6.33) gives
the effective coupling
g2 t→∞ 1
ḡ 2 (t) = −→ (6.39)
1 + b0 g 2 t b0 t
The effective mass can be calculated using γm (g) = a2 g 2 + · · · ,
(∫ t ) ( 2 )a2 /b0
g a2
m̄(t) = m exp γm (ḡ(s)ds ≃ m 2 (t)
∼ mt b0
(6.40)
0 ḡ
Similarly, using γ(g) = c2 g 2 + · · ·
(∫ t )
c2
Z̄(t) = exp γ(ḡ(s)ds ∼ t b0 (6.41)
0

Putting all together we can determine the ultraviolet be-


havior of the Green functions
( )
t→∞ nc2
(4−n)t − b0 1 a2
−t b0
Γ (e pj , g , m, µ) = −→ e
n t 2
t × Γ pi ,
n
, me t , µ (6.42)
b0 t
We see that as we go to higher and higher energy the cou-
pling constant becomes weaker and weaker and perturbation
theory more reliable. The masses in the same limit are
also truly negligible. Moreover there is no Landau pole in
the ultraviolet for a asymptotically free theory. However,
using only the lowest order result (6.39), there is a pole
at t = −1/b0 g 2 , which corresponds to a singularity of the
propagator, say, for small space-like momenta. However,
for asymptotically free theories physicists think there are
phenomena in the infrared that can cure this behavior, such
as mass generation and confinement.

160
6.5 General results of the RG analysis
We collect few general results concerning the renormaliza-
tion group, beyond the study of asymptotic behavior of Green
functions.

6.5.1 Prescription dependence


Up to now we made use of the M S renormalization scheme,
so one may ask what features of the beta function, or
the anomalous dimensions, et cetera, are independent on
the renormalization scheme. Given two mass independent

renormalization schemes, leading to a new coupling λ and
′ ′
new Za (λ ), we have finite renormalizations
′ ′ ′
λ = G(λ) = λ + aλ2 + bλ3 + . . . , Za (λ ) = Fa (λ)Za (λ) (6.43)

where Fa (λ) = 1 + cλ + dλ2 + · · · . It follows that


′ ′ ∂G
β (λ ) = β(λ) (6.44)
∂λ
Then the following properties are prescription independent:

• The existence of a fixed point λ∗ (but not its location):


′ ′ ′
β(λ∗ ) = 0 if and only if β (λ∗ ) = 0, where λ∗ = G(λ∗ )

• the slope at the fixed point, since


′ ′
dβ (λ∗ ) dλ dG(λ∗ ) dG(λ∗ ) dβ(λ∗ ) dβ(λ∗ )
′ = ′ ′ β(λ∗ ) + =
dλ dλ |λ∗ dλ dλ dλ dλ

and G(λ∗ ) = λ∗ .

• the value of the anomalous dimension at the fixed point,


γ(λ∗ ). In fact

′ ′ d log F (λ)
γ (λ∗ ) = γ(λ∗ ) + β(λ∗ ) = γ(λ∗ )

• the first two term in the expansion of the beta function


and the first term in the γ function. In fact, writing

161
β(λ) = b0 λ2 + b1 λ3 + · · · , we have from (6.43), (6.44),
′ ′ ′ ′ ′ ′ ′ ′ ′
β (λ ) = b0 λ 2 + b1 λ 3 + · · · = b0 λ2 + (2ab0 + b1 )λ3 + O(λ4 )
= (1 + 2aλ + 3bλ2 + · · · )(b0 λ2 + b1 λ3 + · · · )
= b0 λ2 + (2ab0 + b1 )λ3 + O(λ4 )
′ ′
from which b0 = b0 , b1 = b1 . Similarly, if γ(λ) = γ1 λ + · · ·
′ ′ ′ ′ ′
and γ (λ ) = γ1 λ + · · · then γ1 = γ1 . This requires another
small computation.

6.5.2 Physical Parameters


The mass m, the coupling and may be other parameters defined
in the M S scheme are not necessarily physical parameters,
in particular m will not be the pole of the propagator as
it is required of the physical mass. Physical observables
are the S-matrix elements and the masses as defined by the
poles of the propagator. We are going to show that any such
quantity P is a renormalization group invariant, it does
not depend on the arbitrary scale µ in the sense that

RP = 0 (6.45)

where R is the operator


∂ ∂ ∂
R=µ + β(g) + mγm (6.46)
∂µ ∂g ∂m
Consider first the physical mass: it is defined by the
position of the pole of the propagator, which near it has
the structure
R2 (m, g, µ)
G2 (p; g, m, µ) = + A(p) (6.47)
p2 + m2p
The two-point function satisfies the RG equation (6.19)

RG2 (p; g, m, µ) = −2γ(g)G2 (p; g, m, µ) (6.48)

Then
RR2 (g, m, µ) R2 (m, g, µ)
− 2 Rm2p + RA = −2γ(g)G2 (p; m, g, µ) (6.49)
p2 + m2p (p + m2p )2

162
From this we derive the result

Rmp = 0 (6.50)

as was to be shown, and also the identities

RR + γ(g)R = 0, RA = −2γA (6.51)

Now we consider the S-matrix. An S-matrix element is given


in terms of proper vertices by the LSZ reduction formulae

S(p1 , . . . , pn ; g, m, µ) = Rn (g, m, µ)Γn (p1 , . . . , pn )|p2i →−m2p (6.52)

Since mp is invariant after Eq. (6.50), using (6.51) and the


RG equation (6.15), we obtain immediately the RG invariance
condition of the scattering amplitudes

RS(p1 , . . . , pn ; g, m, µ) = 0 (6.53)

6.5.3 Dynamically generated masses


It is not necessarily true that in the massless limit m →
0, the physical mass mP → 0, too. If not, then a mass
can be generated dynamically, for example as a result of
a spontaneously broken symmetry that would be present in
the zero mass limit. To explore whether this dynamically
generated mass is possible, we use the RG for a theory with
a single coupling constant g. The physical mass has the
form
m
mP = mf˜(g, ) = mf (g, z), z = log(m/µ) (6.54)
µ
for some function f (g, z) and must satisfy RmP = 0. Therefore
[ ]
∂ ∂ ∂
µ + β(g) + mγm mf (g, z) = 0 (6.55)
∂µ ∂g ∂m
or
[ ( )]
∂ ∂ ∂
− + β(g) + γm 1 + f (g, z) = 0 (6.56)
∂z ∂g ∂z
Defining

β γm
B= , Γ=
1 − γm 1 − γm

163
the solution is
(∫ z )
f (g, z) = f (1, ḡ(z)) exp Γ(ḡ(t))dt (6.57)
0

where
dḡ(t)
= B(ḡ(t)), ḡ(0) = g (6.58)
dt
The limit m → 0 corresponds to z → −∞, so we are probing
the infrared limit of the RG. If there is an IR stable fixed
point g∗ , and if γm (g∗ ) < 1, then
m→0 1
mP −→ mf (1, g∗ ) exp[zΓ(g∗ )] = f (1, g∗ )m 1−γm (g∗ ) → 0 (6.59)

Thus in this case it is impossible that dynamical mass


generation can occur. In the non-asymptotically free case,
where the coupling in the infrared is attracted to the
origin, that is g∗ = 0, we simply have mP → m1+Γ(g∗ ) → 0 as
m → 0, since Γ(0) = 0; so a non asymptotically free theory
cannot generate a dynamical mass.
Consider now the case of an asymptotically free theory, and
set m = 0. If this theory generates dynamically a physical
mass, it is given by mP = µf (g) by dimensional analysis.
Then
df
RmP = µf (g) + µβ(g) =0 (6.60)
dg
or making the trivial integration
( ∫ )
g(µ)
dx
mP = µ exp − (6.61)
β(x)

Thus for example, if β(g) ∼ ±g 3 , as g → 0, we conclude


{
− 1
µe g2 theory is asymptotically free
mP ∼ 1 (6.62)
µe g2 theory is infrared stable
The behavior of an IR stable theory is absurd, since no
masses can be generated by turning off the interaction, and
mP must vanish. For an AF theory, on the other hand, the
behavior is reasonable: the mass turns off rapidly as we
turn off the coupling, and has an essential singularity at

164
g = 0, as was to be expected for a dynamically generated mass
which vanishes at any order in perturbation theory if it is
protected by some symmetry, like gauge invariance or chiral
symmetry. If there is a UV cutoff Λ (perhaps the Planck
mass at ∼ 1019 Gev), and we send µ → Λ, then g 2 decreases in
the limit
1
g 2 (µ) −→
µ→Λ log(Λ/mP )

driving the bare coupling g02 = g 2 (Λ) to even small values (see
the end of the next section).

6.5.4 Dimensional transmutation


Consider a theory with no dimensional parameters, like
masses, and one dimensionless coupling g. We have seen
that if the theory is AF some particle may develop a phys-
ical mass. One example is QCD with massless quarks, which
has chiral symmetry. This is broken dynamically, a physical
mass scale is produced and the proton, for example, will be
massive. Any such mass will have a dependence on µ and g
as in Eq. (6.61). If any other physical parameter in the
theory, call it P(g, µ), has mass dimension M ∆ , then it can
be written as

P(g, µ) = m∆
P f (g)

But then the RG condition RP(g, µ) = 0 gives β(g)∂f /∂g = 0,


and f (g) is independent of the coupling. Thus all physical
observables are given by powers of the physical mass times
calculable, dimensionless numbers.
This fact is known as dimensional transmutation. Initially
we had only one dimensionless parameter g, but we ended with
another parameter, the mass scale mP given by one of the
physical masses of the theory. So apart from this mass,
whose value has no meaning in the absence of another mass
scale, the theory has no adjustable parameters. This means
that there are no small parameters available, and that non
perturbative dynamics is required in order to produce the
physical mass scale.

165
It is instructive to look at the behavior of the physical
mass given by (6.61), as µ is scaled up to the order of
the UV cutoff Λ. Then, given β(g) = −b0 g 3 /2 (as for QCD) and
µ → Λ, the coupling scales as g(µ) → g(ln Λ); but if Λ is the
UV cutoff, this is the bare coupling g0 , so that
( ∫ ) ( ) ( )
g(µ)
dx 1 1
mP = µ exp − = Λ exp − 2 = Λ exp − 2
β(x) b0 g (ln Λ) b0 g 0

Therefore as Λ → ∞, the bare coupling vanishes as g02 ∼


1/b0 ln Λ, so as to insure that mP remains finite.

6.5.5 The effective potential


The effective action given by Eq. (6.12) satisfies the RG
equation (6.22). For constant φ(x) = ϕ, the Fourier trans-
form is φ̃(p) = ϕ(2π)4 δ 4 (p), and the effective action takes the
form Γ[φ] = −Ω4 V (ϕ), where Ω4 is the space-time volume. The
ordinary function V (ϕ) is known as the effective potential,
the classical potential plus quantum corrections. It is the
minimum of the expectation value of the quantum Hamiltonian
taken over the set of unit norm states |Ψ⟩ for which the
expectation value of the quantum field φ is the classical
function ϕ: ⟨Ψ|φ|Ψ⟩ = ϕ.
From Eq. (6.22), V (ϕ) satisfies the RG equation
( )
∂ ∂ ∂
µ + β(λ) + γm (λ)m + γ(λ) V (ϕ, m, λ, µ) =
∂µ ∂λ ∂m
∂(ϕV (ϕ))
= γ(λ) (6.63)
∂ϕ

Just as for the Green functions and proper vertices, this


equation is easy to solve. We obtain
( ∫t )
t 4t − 0 duγ(λ(u))
V (e ϕ, λ, m, µ) = e V e ϕ, λ(t), m(t), µ ) (6.64)

where the running quantities λ(t), m(t), etc., are on some


renormalization group trajectory of the beta functions. We
can use this solution to explore the behavior of V (ϕ) for
small (IR) and large (UV) values of ϕ.

166
The small ϕ behavior is reached as t → −∞, therefore it is
controlled by the infrared stable fixed point of the theory.
A pure λϕ4 /4! theory with λ > 0 is infrared stable, the beta
function being given by β(λ) = 3λ2 /16π 2 + · · · . Hence the flow
is

λ 16π 2
λ(t) = ∼ for large negative t
1 − 16π
3λt
2 3|t|

Note the presence of a Landau pole at t = 16π 2 /3λ. Using


(6.64) and the expression of the classical potential, which
is valid at small coupling, we derive the limiting behavior
ϕ→0 16π 2 ϕ4
V (ϕ, λ, m, µ) −→ (1 + O(1/| ln ϕ|)) (6.65)
3 | ln ϕ|
This definitely rules out the generation of spontaneous
symmetry breaking by radiative correction for this simple
theory (the Coleman-Weinberg mechanism), since (6.65) says
that ϕ = 0 is at a minimum rather than at a local maximum.
The large ϕ behavior is controlled by a UV stable fixed
point, which is the case of this theory provided λ < 0. In
that case

λ 16π 2
λ(t) = 3|λ|t
∼− for large positive t
1+ 3t
16π 2

The large ϕ dependence is now


ϕ→∞ 16π 2 ϕ4
V (ϕ, λ, m, µ) −→ − (1 + O(1/| ln ϕ|)) → −∞ (6.66)
3 | ln ϕ|
The theory is asymptotically free but non existent (unsta-
ble), since it has states with negatively unbounded energy.
In fact, the only known asymptotically free theories with
bounded energy from below are the non abelian gauge theo-
ries, like QCD.

167
6.6 Calculation of the beta function
The starting point will be Eq. (6.20) together with the
expansion (6.21). We recall both, here, in the form
( )
∂ 1
β(g, ϵ) 1 + g Z̃g (g, ϵ) = − ϵg Z̃g (g, ϵ) (6.67)
∂g 2



zn (g)
Z̃(g, ϵ) = 1 + (6.68)
n=1
ϵn

The beta function has the Laurent expansion



β(λ, ϵ) = βj ϵj (6.69)
j≥0

The first equation can be used to compute β(g, ϵ) if Z̃g (g, ϵ) is


known, and conversely, to compute Z̃g (g, ϵ) if β(g, ϵ) is known.
To compute β(g, ϵ), the strategy is to match the coefficients
of ϵn in both sides of the equation. The right hand side of
Eq. (6.67) has no powers of ϵ greater than one, so matching
the terms proportional to ϵi for i ≥ 2, gives immediately
βi = 0 for i ≥ 2. Then we can write β(g, ϵ) = β0 (g) + β1 (g)ϵ; but the
term proportional to ϵ fixes β1 (g) = −g/2, so

β(g, ϵ) = β0 (g) − ϵg/2 (6.70)

while the term in ϵ0 gives β0 ≡ β as


1 ′ ′ dzi
β(g) = lim β(g, ϵ) = g 2 z1 (g), zi (g) = (6.71)
ϵ→0 2 dg
Finally the terms in ϵ−n , n ≥ 1, give the recursion relation
(where z0 = 1)
1 2 ∂zn ∂gzn−1
g = β(g) (6.72)
2 ∂g ∂g
allowing to compute the higher order coefficients, in case
β(g) is known by other means. In conclusion, the beta
function is totally determined by z1 (g), the residue of the
simple pole of Z̃g (g, ϵ).
To compute Z̃g (g, ϵ), instead of proceeding iteratively using
the coefficients zn and Eq. (6.72), one may simply integrate

168
the RG equation (6.67), using for that purpose β(g, ϵ) = β(g)−ϵg.
The solution is (actually the explicit dependence on µ is
apparent, since the ratio is µ-dependent only via the running
of g)
[∫ g ]
g0 µ−ϵ β(x)
Z̃g (g, ϵ) = = exp dx (6.73)
0 x( 2 ϵx − β(x))
1
g

This can be used to express g0 (g, ϵ) as we remove the cutoff,


with interesting consequences. The situation is problematic
for an IR stable theory, assuming β(g) = b0 g 3 + · · · > 0 (this is
the case of QED for g = e and b0 = 1/12π 2 or λϕ4 for λ = g 2 and
b0 = 3/32π 2 ), because the integral can diverge at 21 ϵx − β(x) = 0,

thus for g = g∗ ∼ ϵ/2b0 , see Fig. 5.3.

Figure 6.3: IR stable theory

For small g we can avoid the singularity


√ by expanding g0 as
a function of g, for 0 < g < g∗ = ϵ/2b0 . But the safe region
within with the expansion holds shrinks as ϵ → 0, showing
that we must let g → 0 as we remove the cutoff. Thus, given
a cutoff ϵ, as we vary g from 0 to g∗ (ϵ), g0 varies from 0 to
∞. And conversely, since g∗ (ϵ) → 0 as ϵ → 0, we conclude that
for all bare couplings, the physical coupling must vanish.
If instead of staying in the safe region we try to take
g > g∗ , then g0 is singular because the integral diverges. In
fact from Eq. (6.73) we get, close to g∗ ,
g→g∗ g2 β(g∗ )
g02 ≃ , A(ϵ) = = 1 + O(ϵ) (6.74)
(g∗ − g)A(ϵ) g∗ (β ′ (g∗ ) − 2ϵ )

169
so, for finite ϵ, g0 is complex and the theory fails to be
unitary. Thus it seems that there is no sensible, unitary
definition of a non-asymptotically free theory.
No such problems are encountered for an AF theory, for which
β(g) < 0 close to the origin, since the integral (6.73) has
no singularities as long as g < g1 , where g1 is an infrared
stable fixed point. One finds in this case that g0 → 0 as
we remove the cutoff.
In fact, given β(g) = −b0 g 3 + · · · we have
( ∫ 2 )
g
b 0 x g2ϵ ϵ
g02 ≃ g 2 exp − dx ϵ 2
= 2 + ϵ/2
→ = ḡ 2 (1/ϵ) (6.75)
0 2
+ b 0 x b 0 g b 0

Thus as g is small, so we can trust perturbation theory, g0


is even smaller, so we can also trust perturbative renor-
malization.

6.7 Other renormalization schemes


The ’t Hooft-Weinberg RG equations derived in Sec. 4.2 are
valid if the UV divergences are eliminated by the M S renor-
malization scheme. In other schemes the RG equations are
not only different, but the beta functions could also de-
pend on the renormalized mass. We shall describe few more
schemes in the following, one of which, leading to the
Callan-Symanzik equation, is also very important in QFT.

6.7.1 The RG equation of Gell-mann and Low


In the renormalization scheme employed by the cited authors,
the mass m is the physical mass, defined by the position
of the pole in the propagator, while the coupling g and the
renormalized field are defined by the off-shell subtraction
procedure at the Euclidean point p2 = µ2 , where µ is the
renormalization scale. As m is independent of µ, there
is no mass differentiation and one gets the following RG
equations for the amputated connected functions
[ ( ) ( )]
∂ m ∂ m
µ + β g, − nγ g, Γn (pi , g, m, µ) = 0 (6.76)
∂µ µ ∂g µ

170
and the running coupling satisfies the RG equation
( )
dg m
µ = β g, (6.77)
dµ µ
Unlike the mass independent renormalization scheme, the new
beta function and anomalous dimension depends on m through
the ratio m/µ, by dimensional analysis. Due to explicit
dependence of β and γ on µ, it is in general impossible
to give an explicit solution of the RG equation, like the
one we found previously in the M S scheme. An asymptotic
solution for p, µ ≫ m can be found that is identical to the
previous, since β(g, 0) and γ(g, 0) are essentially identical to
the RG functions computed in the M S scheme.

6.7.2 The Georgi-Politzer RG equation


In the scheme devised by these authors a fully off-shell
renormalization of the mass and other quantities is per-
formed at a space-like momentum scale p2 = µ2 . In particular
the renormalized mass m can be defined by
Γ(2) (p)|p2 =µ2 = m2 + µ2 (6.78)
which is equivalent to the self-energy vanishing at µ2 :
Π(µ2 ) = 0, and the propagator behaving as a free propagator
near that point,

p2 →µ2 1
G(2) (p) −→
p2 + m2

The RG equation takes the form


[ ( ) ( ) ( )]
∂ m ∂ m ∂ m
µ + β g, + γm g, m − nγ g, Γn (pi , g, m, µ) = 0
∂µ µ ∂g µ ∂m µ
where as always
( ) ( )
m dg m ∂ log m
β g, =µ , γm g, =µ (6.79)
µ dµ |g0 ,m0 µ ∂µ |g0 ,m0
( )
m 1 ∂ log Z
γ g, = µ (6.80)
µ 2 ∂µ |g0 ,m0
In the intentions of Georgi-Politzer, this RG equation is
convenient to study the dependence of strong interaction
processes on the quark masses. For the details, see.33

171
6.7.3 The Callan-Symanzik RG equation
This section will be more elaborate. To begin, the renor-
malization scheme employs normalization conditions at zero
momentum to define the renormalized mass3 and coupling, m
and g. Thus
dΓ(2)
Γ(2) (p = 0) = −m2 , = −1, Γ(4) (0, 0, 0, 0) = mϵ g (6.81)
dp2 |p2 =0
We also
∫ need the Green functions with insertions of l oper-
ators ϕ (x)d x = ϕ̃2 (0). This requires a new renormalization
2 d

constant, say Z2 , such that ϕ2 (x) = Z2 Z −1 ϕ20 (x) that has finite
matrix elements with ϕ-correlation functions as one removes
the cutoff ϵ = 4 − d; note that ϕ2 (x) ̸= ϕ(x)ϕ(x) = Z −1 ϕ0 (x)ϕ0 (x).
Because obviously

⟨ϕ2 (1) · · · ϕ2 (l)ϕ(1) · · · ϕ(n)⟩ = Z2l Z −l−n/2 ⟨ϕ20 (1) · · · ϕ20 (l)ϕ0 (1) · · · ϕ0 (n)⟩

the relation with the bare correlation functions is


G(l,n) (q1 , . . . , ql ; p1 , . . . , pn , m, g) = Z − 2 −l Z2l G0
n (l,n)
(q1 , . . . , ql ; p1 , . . . , pn , m0 , g0 , ϵ)
The vertex functions are obtained by stripping away the n
external propagators, the factor (2π)4 δ 4 (q1 + · · · + ql + p1 + · · · + pn ),
and retaining only 1PI diagrams4 with n + l external points,
so that
Γ(l,n) (q1 , . . . , ql ; p1 , . . . , pn , m, g) = Z 2 −l Z2l
n
(6.82)
(l,n)
× Γ0 (q1 , . . . , ql ; p1 , . . . , pn , m0 , g0 , ϵ)
The functions Γ(l,n) computed at zero momenta q1 ∫= q2 = · · · = ql = 0
correspond to l insertions of the operator ϕ2 (x)dd x. The
superficial degree of divergence of the functions Γ(l,n) is

δ = 4 − n − 2l

so the only superficially divergent function with one in-


sertion is Γ(1,2) ; to fix Z2 we impose the additional renor-
malization condition
Γ(1,2) (0; p1 = 0, p2 = 0) = 2 (6.83)
3
This mass is not the physical mass.
Strictly speaking, with the standard Feynman rules it is (−i)l−1 Γ(l,n)
4

that one calculates by summing 1PI diagrams.

172
which is satisfied at zero coupling. From the perturbative
standpoint, the following identity is true

∂ (l,n)
m0 Γ (q1 , . . . , ql ; p1 , . . . , pn , m0 , g0 , ϵ) =
∂m0 0
(l+1,n)
− m20 Γ0 (0, q1 , . . . , ql ; p1 , . . . , pn , m0 , g0 , ϵ) (6.84)

since each internal propagator is doubled under mass dif-


ferentiation, due to the identity
∂ −i −i −i
m0 = −i 2 2m20 2 (6.85)
∂m0 p + m0 − iε
2 2
p + m0 − iε
2
p + m20 − iε
We can also prove it using Feynman’s diagrams: we have

(l,n)

Γ0 = il−1 1P I

where the diagrams in the sum have l + n external points.


Then

∂ (l,n) ∑
m0 Γ0 = −m20 il 1P I
∂m0

where the diagrams in the sum now have l + n + 1 external


(l+1,n)
points. But the right hand side is just Γ0 after the
identity (6.85). Q.E.D.
The Eq. (6.84) is related to the (naïve) scale transformation
property of the theory. This can be seen if we write Γ(n)
(either
∑ bare or renormalized) in one of the two forms (where
s = r p2r )
( s p ·p ) ( )
(n) (4−n)/2 i j 4−n (n) p1 pn
Γ (p1 , . . . , pn , m, g) = s Fn , ,g = m Γ , . . . , , 1, g
m2 s m m
which is possible by dimensional analysis. From this we
easily find (we consider the case l = 1 for simplicity)
( n )
∑ ∂ (n)
(n)
∂Γ0
pi · + (n − 4) Γ0 (p1 , . . . , pn , m0 , g0 , ϵ) = −m0 =
i=1
∂pi ∂m0
(1,n)
= m20 Γ0 (0; p1 , . . . , pn , m0 , g0 , ϵ) (6.86)

173
This is in turn the Fourier transform of the Ward identity
for dilatations, which states that
∑n ∫
⟨ϕ(x1 ) · · · δϕ(xi ) · · · ϕ(xn )⟩ = i d4 x ⟨∆(x)ϕ(x1 ) · · · ϕ(xn ⟩ (6.87)
i=1

where ⟨· · ·⟩ = ⟨0|T (· · · )|0⟩ and

δϕ = −(1 + x · ∂)ϕ

is the infinitesimal variation of the field under scale


transformations: x → λx. The factor ∆(x) is the divergence
of the dilatation current

∆(x) = ∂µ JDµ (6.88)

and is equal to m2 ϕ2 in the present theory (recall we are


discussing massive gϕ4 interaction). Also, the T-product in
these equations is the covariant T-product, commuting with
space-time derivatives.
As we indicated in Sec.4.3, the Eq. (6.84), or the equivalent
(6.86), are not true for renormalized vertex functions, even
for a massless theory, due to the running of the coupling
constant.
We restrict now to the case l = 0. To deduce the correct
equations, we proceed to transform (6.84) into a relation
between renormalized functions, by writing first of all

Γ0 = Z −n/2 Γ(n) (p1 , . . . , pn ), = Z2 Z −n/2−1 Γ(1,n) (0; p1 , . . . , pn )(6.89)


(n) (1,n)
Γ0

We multiply (6.84) by (m/m0 )(∂m0 /∂m)|g0 to convert to m(∂/∂m)|g0 ;


the derivative D = m(∂/∂m)g0 done at fixed bare coupling (and
cutoff ϵ), is calculated using the chain rule5

∂ ∂ ∂
D =m =m + Dg
∂m |g0 ∂m |g ∂g |m

Eq. (6.84) is then, using (6.89)


( ) 1
D Z −n/2 Γ(n) (pi , g, m) = − D(m20 )Z2 Z −n/2−1 Γ(1,n) (0; pi , g, m) (6.90)
2
5
Since g = m−ϵ Zg (g, ϵ)g0 , at fixed g0 it becomes a function of m.

174
or explicitly,
( )
∂ ∂
m + β(g) − nγ(g) Γ(n) (pi , g, m) =
∂m ∂g
= −m2 σ(g)Γ(1,n) (0; pi , g, m) (6.91)

where we defined

β(g) = Dg, 2γ(g) = D log Z, 2m2 σ(g) = Z −1 Z2 D(m20 ) (6.92)

A final touch will determine the coefficient of the right


hand side; for n = 2 and p1 = p2 = 0 the renormalization
conditions (6.81) and (6.83) gives the relation

γ(g) + σ(g) = 1

so the right hand side of (6.91) is finite, and we finally


get the Callan-Symanzik equation
( )
∂ ∂
m + β(g) − nγ(g) Γ(n) (pi , g, m) =
∂m ∂g
= (γ(g) − 1)m2 Γ(1,n) (0; pi , g, m) (6.93)

It can be considered as the correct Ward identity for di-


latation symmetry. In fact, using dimensional analysis as
above in connection with Eq. (6.86), we can write
( n )
∂ ∑ ∂
−m Γ(n) (pi , g, m) = pi · + (n − 4) Γ(n) (pi , g, m)
∂m |g i=1
∂pi

and the CS equation takes the alternative form


( n )
∑ ∂
pi · + (nδ − 4) Γ(n) (pi , g, m) =
i=1
∂pi
∂ (n)
= β(g) Γ (pi , g, m) + (1 − γ(g))m2 Γ(1,n) (0; pi , g, m) (6.94)
∂g

where δ = 1 + γ(g) is the so called “anomalous dimension”. As


in Sec. 4.3, this can be easily turned into a formula for

175
the scale variation of the effective action

δΓ 4 ∂Γ[φ, g, m]
(δ + x · ∂)φ(x) d x = β(g) +
δφ(x) ∂g

+ (2 − δ(g))m2 Γ[φ, g, m] (6.95)
∂m2
so scale invariance is broken even for the massless m = 0
theory. However, if the theory sits at a fixed point g∗
where β(g∗ ) = 0, then for the massless theory scale invariance
is maintained, albeit with a slightly different dimension
for the fields, d∗ = 1 + γ(g∗ ). More interesting is the case
where g∗ is a UV stable fixed point, that is when the running
g(t) → g∗ as t → ∞. In that case scale invariance is recovered
asymptotically in the UV, as was first observed by K. Wilson
and expected on physical grounds.

6.8 The QED renormalization group equa-


tions [incomplete...]
Consider the correlation functions for n pairs of fermion
fields ψ, ψ and l photon fields given by (using tensor
notation to avoid clumping of spinor and vector indices)

∆0nl (x1 , y1 , x2 , y2 , · · · ; z1 , · · · zl , m0 , e0 , ϵ) =
= ⟨ψ0 (x1 ) ⊗ ψ 0 (y1 ) · · · A0 (z1 ) ⊗ · · · ⊗ A0 (zn )⟩ (6.96)

The renormalized correlation functions are related to the


bare functions by the usual moltiplicative renormalization

∆nl (x1 , y1 , x2 , y2 , · · · ; z1 , · · · zl , m, e, ξ, µ) =
−l/2
= Z2−n Z3 ∆0nl (x1 , y1 , x2 , y2 , · · · ; z1 , · · · zl , m0 , e0 , ξ0, ϵ) (6.97)
√ √
where
√ ψ0 = Z2 ψ and A0 = Z3 A are the renormalized fields, e =
e0 Z3 the √renormalized electric charge, m the renormalized
mass, ξ0 = Z3 ξ, µ the renormalization scale and ϵ = 4 − d the
cutoff. The constants Zi being dimensionless are functions
of e and ϵ, and Z2 is also gauge dependent.

176
6.9 The case of QCD [incomplete...]
Quantum chromodynamics is the dynamical theory of the strong
interaction among NF = 6 quarks flavor, based on a gauge
internal symmetry under the group SU (3), known as the color
symmetry.

6.9.1 Renormalization
The bare Lagrangian for an SU (3) gauge theory like QCD with
f massless quark flavors6 including gauge fixing and ghosts,
is (in d = 4 − ϵ dimensions)
∑ ( )
1 a aµν 1 λa a
L = − F0µν F0 − (∂µ A0 ) −
aµ 2
Ψ0f ∂/ − ig0 A/0 + m0 Ψ0f
4 2ξ0 f
2
( )
− i∂ µ θ̄0a ∂µ + g0 f abc Ab0µ θ0c (6.98)

where λa are the standard Gell-mann matrices and a sum


over indices a, b, c, . . . goes from 1 to 8, the dimension of
SU (3). As always, to take into account the ultraviolet
divergences appearing in the calculation of loop integrals,
the Lagrangian is written in terms of renormalized fields
and couplings
√ √ √
A0 = Z3 A, Ψ0f = Zf Ψf , θ0 = Z̃θa , g0 µ−ϵ/2 = Zg g, (6.99)
a

m0 = Z m m

by which it takes the form


1 1 ∑ ( ϵ
)
L = − Z3 Fµν a
F aµν − (∂µ Aaµ )2 − Zf Ψf ∂/ − igµ 2 Z̃g A
/ + mZm Ψf
4 2ξ f
( ϵ
)
− iZ̃∂ µ θ̄a ∂µ + gµ 2 Z̃g f abc Abµ θc (6.100)

where
λa a λa µ a ϵ
A/ = A/ = γ Aµ , F a = dAa + gµ 2 Z̃g f abc Ab ∧ Ac (6.101)
2 2

and we defined Z̃g = Zg Z3 and ξ = ξ0 /Z3 . For perturbative
calculations one may further split the Lagrangian into a
free quadratic part plus interactions terms, and afterward
6
f = (u, d, c, s, t, b)

177
also split the constants as Zi = 1 + δi , putting the terms
depending on δi into a remainder known as the counter-term
Lagrangian Lc , which has the form
1 ( )2 ∑ ∑
Lc = − δ3 ∂µ Aaµ − ∂ν Aaµ − δf Ψf ∂/Ψf − δmf Ψf Ψf
4 f f
ϵ ϵ
− iδ̃ ∂ µ θ̄a ∂µ θa + gµ 2 δ1 f abc ∂ µ θ̄a Abµ θc + igµ 2 δ2 Ψf A/Ψf + · · · (6.102)

where the dots contain the gauge self-interactions coun-


terterms. However we do not do this here, our purpose being
only to identify the independent renormalization constants.
All five renormalization constants differ from one by some
power g κ of g: Zi = 1 + O(g κ ), so the δi are all of order g at
least.
In dimensional regularization, the beta function is deter-
mined from Zg as follows
ϵ ∂ log Zg ϵ ∂ log Zg
β(g, ϵ) = − g − gµ =− g−g β(g, ϵ) (6.103)
2 ∂µ 2 ∂g
where µ derivatives are √ taken at fixed bare parameters and
cutoff ϵ. Now Zg = Z̃g / Z 3 appears in all four interaction
monomials (along with other constants), and therefore it can
be determined in several different ways. For example, Z̃g can
be determined either from the ghost-gluon vertex and Z̃, or
from the quark-gluon vertex and Zf , or from the gluon-gluon-
gluon vertex, and so on, while Z3 can be determined from √
the UV divergences of the gluon propagator; then Zg = Z̃g / Z3
could also be determined. But do all these independent
calculations give the same result? The key is the gauge
invariance of the quantum theory, which implies the Slavnov-
Taylor identities among Green functions which guarantee the
consistency of the different calculations. In fact, the
constants in front of the four interaction monomials
3/2 1/2 1/2
A3g = Zg Z3 , A4g = Zg2 Z32 , Af g = Zg Z3 Z2 , Aθg = Zg Z3 Z̃

namely the three gluon, four gluon, fermion-fermion-gluon


and ghost-ghost-gluon interactions, respectively, are ev-
idently not all independent, there being three relations

178
among them. For example

A3g Af g Aθg A4g


= = =
Z3 Z2 Z̃ A3g

These were once called the Ward (or Slavnov-Taylor) iden-


tities.

6.9.2 Asymptotic freedom


We report on a classical diagrammatic computation of the
beta functions for non abelian gauge theory. For details,
see for example.3, 4, 12 We can use the results of the previous
section to compute the beta function of non abelian gauge
theory. According to the analysis in Sec. 4.6 in dimensional
regularization with modified minimal subtraction, we have
to pick up the coefficient of the simple 1/ϵ pole in the
expansion of Zg . To determine Z3 we need the gluon propagator
to one-loop, the last graph being the contribution of the

Figure 6.4: The gluon self-energy to one-loop

counterterm, which is
( )
−(Z3 − 1)δab k 2 ηµν − kµ kν

179
The self-energy has the same structure (Slavnov-Taylor iden-
tities again)
( )
Πab;µν = π(k 2 , ϵ)δab k 2 ηµν − kµ kν

Picking up the pole this gives (after a certain amount of


calculations!)
( )
g2 5 4 1
Z3 = 1 + 2 T (A) − Nf T (R) + O(g 4 ) (6.104)
8π 3 3 ϵ
where T (A) is the trace of the adjoint representation and
similarly T (R) the trace of the spinor representation; that
is

T (A)δ ab = Tr TAd
a b
TAd = f acd f bcd , T (R)δ ab = Tr T a T b
cd

T (A) = Nc for SU (Nc ), while T (R) = 1/2 for a single quarks


in the SU (3) fundamental representation. To compute Z̃g we
choose the fermion propagator and the vertex. The counter-

Figure 6.5: The quark self-energy to one-loop

term here is ip/(Zf − 1); After some computation one gets


g2 1
Zf = 1 − 2
C(R) (6.105)
8π ϵ
where
∑ N2 − 1
δij C(R) = (T a T b )ij = δij
a
2N

180
the last value being valid for SU (N ), or 4/3 for QCD. To
compute Z̃g we look at the fermion-gluon vertex (fig.4.6),
where the counter-term is

i(Z̃g Zf − 1)g(T a )ij γ µ

Figure 6.6: The quark-gluon vertex to one-loop

Picking up the 1/ϵ-pole gives the result


g2 1
Z̃g Zf = 1 − 2
(T (A) + C(R)) + O(g 4 ) (6.106)
8π ϵ
Note that this is different from Zf , whereas in QED they
would be equal according to a celebrated Ward identity. In
any case this gives
g2 1
Z̃g = 1 − 2
T (A) + O(g 4 ) (6.107)
8π ϵ
−1/2
Since g0 µ−ϵ/2 = Zg g, we actually need Zg = Z̃g Z3 , or
( )
g2 11 2 1 z1 (g)
Zg = 1 − 2 T (A) − T (R) + O(g 4 ) ≡ 1 + + ···
8π 6 3 ϵ ϵ

From Eq. (6.103) we derive the beta function


( )
1 2 dz1 g3 11 4Nf
β(g) = g =− 2
T (A) − T (R) + O(g 5 ) (6.108)
2 dg 16π 3 3
For QCD with three colors this is negative for Nf ≤ 16 quark
flavors, which is the case in the real world. It follows
from the general analysis that the origin is a UV stable

181
fixed point of QCD with less than 16 flavors, or that QCD
is an asymptotically free theory.
The property of asymptotic freedom was a landmark 1973 dis-
covery of D. Gross, F. Wilczek and H. D. Politzer7 . Besides
the totally unexpected features of these properties of non
abelian gauge theories as seen from that times, at the
top of the bootstrap hypothesis and S-matrix theory, the
establishment of AF had many benefits:

• As discussed in previous sections, it made QCD self-


consistent: there is a well defined continuum limit
were the bare charge goes smoothly to zero as the UV
cutoff increases at fixed coupling. No Landau poles are
encountered.

• It explained SLAC deep inelastic experiments. When the


energy is large the quark interactions become weak,
and this is what it was seen. Moreover a quantitative
theory of the deviations from simple scaling could be
developed, which agree well with the data.

• It explained why the strong interaction is so stronger


than the other interactions.

• It gave a cloud, but not a proof, of confinement, since


the interaction becomes strong at larger distances.

6.9.3 The running coupling


We now want to study the behavior of the QCD coupling as
the momentum scale of interest is varied. In terms of the
coupling (known as α-strong)
g2
αs = (6.109)

it is customary in QCD to define the beta function via8
dαs
µ2 = β(αs ) (6.110)
dµ2
7
They were, respectively, 32, 21 and 23 years old at the time of the
discovery!
8
Use of 2µ2 dµd 2 = µ dµ
d
makes trivial the transcription.

182
This beta function up to three loops order is2, 12, 13
β(αs ) = −b0 αs2 − b1 αs3 − b2 αs4 + O(αs5 ) (6.111)
where
33 − 2nF 153 − 19nF
b0 = , b1 = , (6.112)
12π 24π 2
77139 − 15099nF + 325n2F
b2 =
3456π 3
nF is the number of active quarks, those quarks whose masses
are below the scale of interest. The three loop coefficient
b2 is scheme dependent and given here in the M S renormal-
ization scheme, the favorite one in QCD. The graph up to
two loops and nF = 3 is (taken from13 )

We recall that the beta function appears in the scale


dependence of the observables, since they must be invari-
ant under change of the arbitrary scale µ (See Section
5.5.2). Let for example R(Q2 , µ, αs ) a dimensionless observ-
able (a branching ratio, a four-point amplitude, a normal-
ized scattering matrix element, a ratio of physical masses)

183
depending on a momentum scale Q2 (we may ignore the quark
masses for Q2 much bigger than any other mass scale). For
example, the process may be the long studied deep inelastic
scattering of leptons off protons,
2.2 Electromagnetic interaction 9

Px
P

Fig. 2.1 Lowest order graph for deep inelastic scattering.

where Q2 = −q 2 is very large even in comparison with the


proton mass, and 2.2 the Electromagnetic R could be taken as the
observable interaction
ratio of the two structure functions of the process (see for
example15 ).Consider
Since R(Q2 ,(or
an electron αs )muon)
is scattering
dimensionless,
off a nucleon withnaive
an scaling
would suggest electronthat
(or muon) detected at
R takes the top vertex in fig. 2.1.
asymptotically Thus
a constant value
the neutral current cross-section involves , and ZO exchange as
independent on Q. But this is not true experimentally9
well as the interference between the two. For Q2 < 103 Ge V 2
precisely the because the interaction
electromagnetic coupling constant
dominates runs
and, for the momentwith
, Q2 , and
we shall consider
renormalization just the photon
introduces exchange. The
a second massinclusion
scaleof the µ on which
2 weak neutral contribution appears in chapter
R(Q , αs ) actually depends. Using dimensional analysis we can 3.

write then R = R(Q2 /µ2 , αs ), so the µ-independence condition


is 2.2.1 Structure functions
( )
2 d
The2 amplitude
2
corresponding
2 ∂
to fig. 2.1 is ∂
µ R(Q /µ , 1, αs ) = µ + β(αs ) R(Q2 /µ2 , αs ) = 0 (6.113)
dµ2 ∂µ 2 1
T = e u(k', X)rl'u(k, A) 2' <
2 ∂α s
(7 > (2.4)
q
2
Using t = log(Q /µ2 ) we get
Summing over hadronic states, the unpolarised cross-section is
( )
∂ ∂
− + β(αs ) R(et , αs ) = 0 (2.5) (6.114)
∂t ∂αs
which gives
The solution is R(et , αs ) = R(1, αs (Q2 )), where αs (Q2 ) is the so-
lution of the differential equation (2.6)
dαs (Q2 ) dαs
= Q2 2 = β(αs (Q2 )), αs (0) = αs (µ2 ) (6.115)
dt dQ
9
In deep inelastic scattering the deviations are known as Bjorken
scaling.

184
We see that all scale dependence of R is determined by the
running αs (Q2 ), with µ2 hidden in the initial data. The
solution of Eq. (6.115) is implicitly given by
∫ αs (Q2 )
dx Q2
t = exp , t = log 2 (6.116)
αs (µ2 ) β(x) µ

As we go to larger Q2 , αs (Q2 ) becomes smaller so one can


hope to compute the solution in perturbation theory with
increasing accuracy if the beta function is known. We can
now define the active flavors more precisely: these are the
flavors with masses below the scale Q2 at which we want to
compute αs .
From the expression (6.110) of the beta function, we can
write
∂αs
Q2 = −b0 αs (Q2 ) − b1 α22 (Q2 ) + O(αs3 ) (6.117)
∂Q2
Taking for example the one-loop approximation (b1 = b2 = · · · = 0)
we get
αs (µ2 )
αs (Q2 ) = (6.118)
1 + αs (µ2 )b0 log(Q2 /µ2 )
which is valid if both αs (Q2 ) and αs (µ2 ) are in the perturbative
region. We see asympotic freedom at work: the coupling
slowly decreases with the energy scale. An alternative way
to write the solution which eliminates the initial data, is
1
αs (Q2 ) = (6.119)
b0 log(Q2 /Λ2 )
Then a dimensionless parameter, αs , is traded for the mass
scale Λ, a particular instance of dimensional transmutation.
Using the measured value at the Z 0 -boson mass, at MZ = 91.18
Gev,

αs (MZ2 ) = 0.116 ± 0.005

one obtains a value in the neighborhood of Λ ≃ 200 Mev. Adding


the two-loop result (6.111), we get
[ ]
1 b1 log log(Q2 /Λ2 )
2
αs (Q ) = 1− 2 + ··· (6.120)
b0 log(Q2 /Λ2 ) b0 log(Q2 /Λ2 )

185
One says that is working to leading order (LO) and next-
to-leading order (NLO), respectively. Then the solution
R(1, αs (Q2 )) can be determined, in particular its scaling prop-
erties for large Q2 .
We neglected quark masses in the above discussion. If
included, we get of course the more general dependence
R(Q2 /µ2 , αs , m/Q) (we are using only one quark flavor), and
the renormalization group equation is
( )
2 ∂ ∂ ∂
µ + β(αs ) − γm (αs )m R(Q2 /µ2 , αs , m/Q) = 0 (6.121)
∂µ2 ∂αs ∂m
The function γm is the anomalous mass dimension and can be
calculated; the result in M S scheme is10
αs 303 − 10nF 2
γm =+ 2
αs + O(αs3 ) (6.122)
π 72π
The equation for the running is (see (6.24))
∂m
Q2 = −γm (αs )m(Q2 ) (6.123)
∂Q2
with the solution
( ∫ )
Q2
dq 2
m(Q2 ) = m(µ2 ) exp − γm (αs (q 2 )) 2 (6.124)
µ2 q
( ∫ )
αs (Q2 )
γm (x)
= m(µ2 ) exp − dx (6.125)
2
αs (µ ) β(x)
so, again, the non trivial scale dependence is contained in
the running of the coupling and the mass, i.e.

R(Q2 /µ2 , αs , m/Q) → R(1, αs (Q2 ), m(Q2 )/Q)

The running of the mass of the bottom quark is graphically


represented in the next figure (taken from13 ). We have
then a theoretical justification in neglecting quark masses
at large Q2 . The essential point is that the anomalous
dimension in an AF theory leads always to a logarithmic
dependence of m(Q2 ), which can never overcome the canonical
dimension in the factor m(Q2 )/Q appearing in R, but leads
to further suppression. We leave at this point the complete
10
R. Tarrach, Nuc. Phys. B183 (1981),384.

186
(non trivial) discussion to the literature and the books,
e.g.10, 12, 13, 36
A summary pf αs measurements corresponding to the strong
coupling measured at the mass of the Z boson, which has the
value αs (Mz2 ) = 0.116 ± 0.005, is illustrated in the next figure.
For a recent review see.37

187
6.10 Appendix
6.10.1 A - Operator insertions
We used insertions of operators ϕ2 (x) into Green functions to
derive the Callan-Symanzik equation. It is interesting that
they can be managed with the functional integral formalism.
We introduce a source K(x) for ϕ2 (x) and define a new W [J, K]

2
exp iW [J, K] = (dφ)eiA[φ]+iJ⋆φ+iK⋆φ (6.126)

Then with a somewhat implicit notation11

δ l+n W [J, K]
=
δK(1) · · · δK(l)δJ(1) · · · δJ(n) J=K=0
= il+n−1 ⟨ϕ2 (1) · · · ϕ2 (l)ϕ(1) · · · ϕ(n)⟩c (6.127)

Notational simplification: we shall denote the functional


derivatives as W (l,n) , a ⋆ is a convolution, the bracket is
⟨0|T · · · |0⟩c and a suffix c stands for “connected”, that is
11
K(1) = K(x1 ), J(1) = J(y1 ) and so on.

188
the bracket is computed summing connected Feynman diagrams
with n + l external vertices of two types. The figure is the
graphical representation of W (3,5) with dotted lines repre-
senting ϕ2 insertions. Now we perform a Legendre transform

= W (3,5) [J, K]

over J
Γ[ϕ, K] = W [J, K] − J ⋆ ϕ (6.128)
where
δW [J, K]
ϕ= (6.129)
δJ
One assumes that a solution for J exists, a functional of
ϕ and K. We have
δϕ δϕ
= W (0,2) = i ⟨ϕ(1)ϕ(2)⟩J,K , = W (1,1) = i ⟨ϕ2 (1)ϕ(2)⟩J,K (6.130)
δJ δK
We also assume W (1,1) = 0 for J = K = 0, since this is an
odd correlation function, and of course W (0,2) is the full
propagator in this limit. As with the effective action, one
has the dual relations
δΓ[ϕ, K] δΓ[ϕ, K] δW [J, K]
= −J, = (6.131)
δϕ δK δK
where the undifferentiated variables are kept fixed (these
are “partial derivatives” so to speak). From the first we
get the identity
W (0,2) ⋆ Γ(0,2) = −δ (6.132)
At zero sources this says that Γ(0,2) is (minus) the inverse
of the full propagator. Taking further derivatives in the

189
sources gives a chain of identities with the general struc-
ture

W (l,n) = W
|
(0,2)
⋆ ·{z
· · ⋆ W (0,2)} ⋆Γ(l,n) + 1P reducible terms (6.133)
n times

where the convolution is over the variables carried by the


fields, the l variables carried by the insertions being free
(not convoluted). This means that Γ(l,n) (x1 , . . . , xl ; y1 , . . . , yn ) is
1PI, connected and amputated, that is

Γ(l,n) (x1 , . . . , xl ; y1 , . . . , yn ) =
c
= il−1 ⟨ϕ2 (x1 ) · · · ϕ2 (xl )ϕ(y1 ) · · · ϕ(yn )⟩AMP,1PI (6.134)

The superficial degree of divergence of the Feynman ampli-


tudes in this equation is

δ = 4 − n − 2l

since each ϕ2 insertion introduces one more propagator into


the diagram. Hence the only new superficially divergent
correlation functions are ⟨ϕ2 ⟩, ⟨ϕ2 ϕ2 ⟩, ⟨ϕ2 ϕ⟩ and ⟨ϕ2 ϕϕ⟩.
The Fourier transform of the vertex functions has the form

e(l,n) (q1 , . . . , ql ; p1 , . . . , pn ) = (2π)4 δ 4 (q1 + · · · pn )Γl,n (q1 , . . . , ql ; p1 , . . . , pn )


Γ

Integrating over the variables


∫ 2 xi in Eq. (6.134), i.e. in-
sertion of the operators ϕ (xi )dx, we obtain obviously the
e(l,n) (0, · · · , 0; p1 , · · · , pn ).
vertices at zero momentum, Γ

6.10.2 B - The Berezin integral


By a given set of Grassmann’s variables θ1 , . . . , θn we understand
elements of an associative algebra G such that

θi θj + θj θi = 0

Functions over G are polynomials of order n with real co-


efficients

f (θ) = w0 + wj θj + wjk θj θk + · · · wn θ1 · · · θn

190
Under constant translation θi → θi + χi the coefficients are
reshuffled except wn : for example
′ ′ ′ ′
f (θ + χ) = w0 + wj θj + wjk θj θk + · · · wn θ1 · · · θn

with

w0 = w0 + wj χj + · · · + wn χ1 · · · χn

and so on, but in any case wn = wn . Therefore if we want to
define a translation invariant integral
∫ ∫
d θf (θ + χ) = dn θf (θ)
n
(6.135)

a sensible choice is simply to define



dn θf (θ) = wn (6.136)

and extend the operation on sums or even limits of sequences


by linearity. One way to achieve this is imposing the rules
∫ ∫
i
dθ = 0, dθi θj = δ ij

and write the measure as dn θ = dθn · · · dθ1 . Under a linear change


of variables

∑ ′
θj = Ljk θk , θ = Lθ
k

one has (because θj1 · · · θjn is totally anti-symmetric)

f (Lθ) = (det L) wn θ1 · · · θn + · · ·

therefore we deduce
∫ ∫
d θf (Lθ) = (det L) dn θf (θ)
n
(6.137)

equivalently
∫ ∫
n ′ −1 ′
d θ f (θ) = (det L) dn θf (θ), θ = Lθ (6.138)

191
just the converse of the change of variable formula for
bosonic integrals
∫ ∫
n −1
d xf (Lx) = (det L) dn xf (x) (6.139)

or
∫ ∫
n ′ ′
d x f (x) = (det L) dn xf (x), x = Lx (6.140)

Of particular interest is the integral



dn θ exp(θ T M θ/2), M T = −M (6.141)

Consider n = 2: then the exponent is M12 θ1 θ2 and the exponen-


tial

exp(θ T M θ/2) = 1 + M12 θ1 θ2

Therefore in this case the value of the integral is just


M12 , or
∫ ( )
2 1 T √
d θ exp θ M θ = det M (6.142)
2
An anti-symmetric matrix can always be written in block
diagonal form by a linear change of variables, so we may
assume without loss of generality that
 
M1 0 0 ...
 0 M2 0 . . . 
M = 0
 (6.143)
0 ... 0 
0 0 . . . Mp
where 2p = n and
( )
0 xk
Mk = (6.144)
−xk 0
The exponential is now
( ) ∏
p
1 T
exp θ Mθ = exp(xk θ2k−1 θ2k )
2 k=1

192
and since the integral of each factor is xk , the full integral
becomes
∫ ( ) ∏
n 1 T
p

d θ exp θ Mθ = xk = det M (6.145)
2 k=1

Note that for bosonic (namely ordinary) integrals


∫ ( )
1 T
d x exp − x M x = (2π)n (det M )−1/2
n
(6.146)
2
It is possible to introduce formally complex anti-commuting
variables by the trick12
θ1 + iθ2 θ1 − iθ2
θ= √ , θ̄ = √ (6.147)
2 2
so that θθ̄ = −iθ1 θ2 and dθdθ̄ = idθ2 dθ1 , and moreover

θθ̄ + θ̄θ = 0, θ2 = θ̄2 = 0

The integrals
∫ ∫ ∫ ∫ ∫
dθ̄θ̄ = dθθ = 1, dθ̄θ = dθθ̄ = 0, dθdθ̄ θ̄θ = 1

follows then from the previous rules. Functions of such


complex variables are again polynomials

f (θ, θ̄) = w0 + w̄1j θj + θ̄j w1j + · · · + wn θ̄1 θ1 . . . θ̄n θn (6.148)

We choose a convenient order since otherwise the sign of the


coefficients is undetermined. Once more the coefficient wn
is defined to be the integral of the function f

dn θdn θ̄f (θ, θ̄) = wn (6.149)

which suggest the practical definition of the measure

dn θdn θ̄ = dθn dθ̄n · · · dθ1 dθ̄1 (6.150)


12
More abstractly, it means that the Grassmann’s algebra admits an
involution ∗ : θ → θ∗ , with ∗∗ = Id.

193
Eq.s (6.138), (6.145) can be easily extended to complex
anti-commuting variables. For instance, under a change of
variables
′ ′
θ̄ = U θ̄, θ =Vθ
′ ′
the n-th order coefficients in the expansion of f (θ , θ̄ ) is
′ ′
f (θ , θ̄ ) → det U det V wn

Therefore the change of variable formula reads


∫ ∫
n ′ n ′ −1
d θ d θ̄ f (θ, θ̄) = [det U det V ] dn θdn θ̄f (θ, θ̄) (6.151)

The Gaussian integral is



Kn = dn θdn θ̄ exp(θ̄M θ) = det M (6.152)

For n = 1 the proof is trivial since K1 = M , for n = 2 with


a simple computation we have K2 = M11 M22 − M12 M21 = det M and,
more in general, Kn = det M . This follows if we remember
a result of linear algebra, that any complex matrix M
can be transformed to real diagonal form by means of a
transformation M → U M V , with unitary matrices U and V .
The integral then reduces to a product of the n eigenvalues.
Shifting variables is a useful technique to calculate
more general Gaussian integrals, like this one

Sn = dn θdn θ̄ exp(θ̄ T M θ + θ̄ T J + J¯T θ) (6.153)

Changing variables to

θ = χ − M −1 J , θ̄ T = χ̄T − J¯T M −1

and using previous results we get

Sn = det M exp(J¯T M −1 J ) (6.154)

194
Bibliography

Textbooks
[1] S. Weinberg, The Quantum theory of fields. Vol. 1:
Foundations, Cambridge, UK: Univ. Pr. (1995)

[2] * S. Weinberg, The Quantum theory of fields. Vol. 2:


Modern Applications, Cambridge, UK: Univ. Pr. (1996)
[Ch. 15,16,18,19,21,22]

[3] * Peskin, M., Schroeder, D. An Introduction to Quantum


Field Theory, Westview Press Inc. 1995. [Part III:
Ch. 12,15,16,20,21]

[4] * M.Srednicki, Quantum Field Theory, Cambridge Univer-


sity Press, 2007. [Part III: Spin one]

[5] * P. Ramond, Field Theory: a Modern Primer, Second Ed.


Westview Press (1997).

[6] J. Zinn-Justin, Quantum Field Theory and Critical Phe-


nomena, Clarendon Press, Fourth Edition (2002).

[7] M. Henneaux and C. Teitelboim, Quantization of gauge


systems, Princeton Univ. Press (1992)

[8] L. D. Faddeev and A. A. Slavnov, Gauge Fields: intro-


duction to Quantum Theory, Addison-Wesley Second Ed.
(1991).

[9] M. D. Schwartz, Quantum Field Theory and the Standard


Model, Cambridge University Press, (2014).

195
Gauge Theories and QCD
[10] *C. Quigg, Gauge Theories of the Strong, Weak and
Electromagnetic Interactions, Second Ed., Princeton
University Press, (2013).

[11] Ta-Pei Cheng and Ling-Fong Li, Gauge Theory of Elemen-


tary Particle Physics, OUP Oxford (1984).

[12] *T. Muta, Foundations Of Quantum Chromodynamics: An


Introduction To Perturbative Methods In Gauge Theories
(Lecture Notes in Physics), Wspc (1987)

[13] *R. K. Ellis, W. J. Stirling, B. R. Webber, QCD and


Collider Physics, Cambridge monographs on particle
physics, 1996 (first digital printing 2003).

[14] A. Smilga, Lectures on Quantum Chromodynamics, World


Scientific, Singapore (2001).

[15] R. G. Roberts, The Structure of the Proton, Cambridge


University Press, (1990).

Standard Model
[16] *C. Burgess & G. D. Moore, The Standard Model: a
Primer, Cambridge University Press, (2007).

[17] J. F. Donoghue, E. Golowich & B. R. Holstein, Dynamics


of the Standard Model, Second Ed., Cambridge Monographs
on Particle Physics (1994).

[18] *P. Langacker, The Standard Model and Beyond, CRC Press
(2009).

Mathematical
[19] V. Rubakov, Classical Theory of Gauge Fields, Princeton
University Press (2002).

196
[20] M. Chaichian and A. Demichev, Path integrals in
physics. Vol. 1: Stochastic processes and quantum me-
chanics, Bristol, IOP (2001)

[21] J. Dimock, Quantum Mechanics and Quantum Field The-


ory: a Mathematical Primer, Cambridge University Press
(2011).

Advanced
[22] B. De Witt, The Global Approach to Quantum Field Theory,
Oxford University Press, 2003.

[23] M. Shifman, Advanced Topics in Quantum Field Theory,


Cambridge University Press (2012).

[24] F. Strocchi, Symmetry Breaking, Lect. Notes Phys. 643


(Springer, Berlin, Heidelberg 2005).

Papers [vastly incomplete...]


[25] *T. Kugo and I. Ojima, Prog. Theor. Phys. Suppl. 66, 1
(1979) [quantiz. gauge theories]

[26] *E. S. Abers and B. W. Lee, Gauge Theories, Phys. Rep.


9, 1 (1973). [review on gauge th]

[27] *S. Coleman and E. Weinberg, Phys. Rev. D7, 1888 (1973)
[radiative corrections ssb]

[28] *Y. Fujimoto, L. O’Raifeartaigh and G. Parravicini,


Nucl. Phys. B212, 268 (1983) [convexity EFP]

[29] *R. Fukuda and T. Kugo, Phys. Rev. D13, 3469 (1976)
[gauge dep. EFP]

[30] *L. F. Abbott, Intro to the Background Field Method,


Acta Phys. Polonica, B13, 33 (1982)

[31] S. Elitzur, Phys. Rev. D 12, 3978 (1975).

[32] Y. Frishman and R. Roth, Nucl. Phys. B 146 (1978), 20.

197
[33] H. Georgi and H. D. Politzer, Phys. Rev. D 14 (1976)
1829.

[34] A. Salam and J. A. Strathdee, Phys. Rev. D 2, 2869


(1970).

[35] I. S. Gerstein, R. Jackiw, S. Weinberg and B. W. Lee,


Phys. Rev. D 3 (1971) 2486.

[36] *P. Z. Skands, Introduction to QCD, arXiv:1207.2389


[hep-th] (2012) [for students]

[37] A. Deur, S. J. Brodsky and G. F. de Teramond, Prog.


Part. Nucl. Phys. 90 (2016) 1.

[38] C. Parenti, F. Strocchi and G. Velo, Phys. Lett. 62B,


83 (1976).

[39] C. Parenti, F. Strocchi and G. Velo, Phys. Lett. 59B,


157 (1975).

[40] R. Forman, Invent.math. 88, 447-493 (1987).

[41] K. Kirsten and A. J. McKane, J. Phys. A 37, 4649 (2004)


[math-ph/0403050].

198

You might also like