Introduction To Conformal Field Theory
Introduction To Conformal Field Theory
PoS(Modave VIII)001
Antonin Rovai∗†
Université Libre de Bruxelles and International Solvay Institues
E-mail: [email protected]
∗ Speaker.
† FNRS Research Follow
c Copyright owned by the author(s) under the terms of the Creative Commons Attribution-NonCommercial-ShareAlike Licence. http://pos.sissa.it/
Introduction to Conformal Field Theory Antonin Rovai
Contents
Foreword 2
Introduction 3
PoS(Modave VIII)001
1.1 Definitions 4
1.2 Noether’s theorem 6
1.3 The energy-momentum tensor 8
1.4 Consequences for the quantum theory 10
References 32
Foreword
These lectures notes were written for the eighth edition of the Modave Summer School in
Mathematical Physics. During this edition, there were several lectures [1, 2] relying on Conformal
Field Theory basics, and thus self-consistency of the school required an elementary introduction to
this vast subject. The present lectures were scheduled at the very beginning of the school, right after
the gong show on the first day and in the morning of the second day. People were then smoothly
taken from very basics to advanced research topics throughout the week, overall in the relaxed
atmosphere of the charming village of Modave, Belgium. This eighth edition was a real success,
and I hope this school will continue to exist for many more years to come.
2
Introduction to Conformal Field Theory Antonin Rovai
Introduction
These lecture notes consist of an elementary introduction to Conformal Field Theory (CFT).
This class of theories have the property that they are invariant under so called conformal transfor-
mations, that may be seen as local scale tranformations. Conformal invariance is a vast topic, and
in these notes we will be able to cover only the very basic elements. Historically, conformal invari-
ance was first used to study statistical systems at a critical point, where scale invariance follows
from the divergence of the correlation length. Since then, the use of CFTs has spread to the field
of fundamental interactions: it is at the heart of string theory, the world-sheet theory being itself
PoS(Modave VIII)001
a two-dimensional CFT; in two dimensions, when quantum gravity is coupled to a matter CFT,
the resulting effective gravitational action is the Liouville action which is a CFT; finally, the best
understood example of a gauge/gravity duality that we know of involves a CFT on the field theory
side, namely N = 4 super-Yang-Mills in four dimensions.
The plan of these notes is as follows. We start very slowly in section 1 with basic results
concerning symmetries in field theory, both at the classical and at the quantum level. The discussion
will be quite general and several examples will be presented, with special emphasis on the energy-
momentum tensor, which play a crucial rôle in the following. In section 2, we introduce and study
conformal transformations in d dimensions. We discuss finite transformations as well as the algebra
of the conformal group, and we explore some simple consequences of conformal invariance on two-
and three-point functions. Finally in section 3 we focus on the special case of d = 2. We introduce
primary fields, study Ward identities and the associated consequences for the energy-momentum
tensor, operator formalism and operator product expansion. We end these lectures with a discussion
of the trace anomaly.
For (much) more details on Conformal Field Theory, we refer the interested reader to the
abundant literature. The standard reference for a modern treatment is the very nice book by Di
Francesco, Mathieu and Sénéchal [3], on which most of the present lectures is based. See also the
lecture notes by Ginsparg [4]. For a more old-fashioned presentation, see the lectures by Coleman
[5]. Note also that Polchinski’s book [6] and Tong’s lecture notes [7] on String Theory both include
a quick introduction to conformal invariance in two dimensions. See also the chapter 15 in [8] for
a nice introduction to more advanced topics.
Our conventions are the following: unless explicitly stated, we work in the euclidean d-
dimensional flat space Rd . Space-time points are denoted by x = (xµ ) with µ = 1, . . . , d, the last
component xd being Euclidean time (except at the end of chapter 3, from section 3.3.1). Vector and
form indices are lowered and raised using the flat metric, and we sum over repeated indices. We
use natural units in which c = h̄ = 1 and we use the symbol ≡ for an equality valid by definition.
Symmetries lie at the heart of our modern conception of physics. It is therefore very important
to understand how we formulate the symmetry properties of a given theory and their consequences
on observables. In this section we quickly review some elementary notions along these lines,
preparing the ground for the analysis of conformal invariance.
3
Introduction to Conformal Field Theory Antonin Rovai
1.1 Definitions
We consider a classical theory for some fields, collectively denoted as Φ, which are functions
on a space-time manifold that we shall take to be flat Rd . The dynamics of the fields Φ is fixed by
a Lagrangian density L (Φ, ∂µ Φ) or by the action S[Φ] defined by
Z
S[Φ] = dd x L (Φ, ∂µ Φ) . (1.1)
Consider a map x 7→ x0 (x), where x0 ∈ Rd is some invertible function of x ∈ Rd , together with some
transformation of the fields Φ 7→ Φ0 defined by
PoS(Modave VIII)001
Φ0 (x0 ) = F(Φ(x)) , (1.2)
for some function F. Under such a transformation, the action will generally be modified: S 7→ S0 ,
with S0 defined by the equation S0 [Φ0 ] = S[Φ], and the transformation is a symmetry if S = S0 .
Let us consider some examples.
x0 = x + a , (1.3)
where a ∈ Rd . Most of the fields Φ that we consider are scalars under translation, that is, F
reduces to the identity:
ϕ 0 (R · x) = ϕ(x) (1.7)
where we use the common notation (R·x)µ = Rµ ν xν . For a vector field V µ the transformation
is
V 0µ (R · x) = Rµ ν V ν (x) , (1.8)
and so on for tensors of various ranks. In the quantum theory, we are also interested in
spinor representations of the rotation group, that is, representations up to a phase. We shall
however continue to write “representation of rotations” instead of “representation of spinor
representation of the rotation group.”
For a field Φ transforming in any representation L of the rotation group, we write the corre-
sponding transformation function F as
Φ0 (R · x) = LR [Φ(x)] , (1.9)
4
Introduction to Conformal Field Theory Antonin Rovai
x0 = λ x (1.10)
PoS(Modave VIII)001
where λ 6= 0. We define Φ0 by
Φ0 (λ x) = λ −∆ Φ(x) , (1.11)
where the number ∆ is called the scaling dimension of Φ. In this case the jacobian of the
transformation is |∂ x0 /∂ x| = λ d and hence
Z
S0 [Φ0 ] = dd x0 L (Φ0 (x0 ), ∂µ0 Φ0 (x0 )) (1.12)
Z
= λd dd x L (λ −∆ Φ(x), λ −1−∆ ∂µ Φ(x)) . (1.13)
1
L (ϕ, ∂µ ϕ) = ∂µ ϕ∂ µ ϕ , (1.14)
2
and the action S is invariant if and only if the scaling dimension ∆ is given by
d −2
∆= · (1.15)
2
m2
1
Z
0 0
S [ϕ ] = d xd
∂µ ϕ∂ µ ϕ + λ 2 ϕ 2 6= S[ϕ] . (1.16)
2 2
Mass terms are thus prohibited in dilatation invariant theories, as well as dimensionful coupling
constants.
5
Introduction to Conformal Field Theory Antonin Rovai
PoS(Modave VIII)001
where summation on the index a is understood. We define the generators Ga by
and hence
δ xµ δF
iGa Φ = ∂µ Φ − · (1.20)
δ ωa δ ωa
The factor of i in the definition (1.19) is introduced for later convenience.
We now identify the parameters and the generators in our examples.
1. Translations: the parameters ωa for an infinitesimal translation are the components aµ of the
infinitesimal vector a defining the infinitesimal transformation, and thus the index a is in this
case a space-time index µ: ωµ = aµ . Using (1.4) we thus find that
δ xµ δF
= δ µν , = 0. (1.21)
δ ων δ ωµ
The generator, that we write Pµ and define by equation (1.20), reads
Pµ = −i∂µ . (1.22)
6
Introduction to Conformal Field Theory Antonin Rovai
iD = x · ∂ + ∆ , (1.28)
PoS(Modave VIII)001
where we used the convenient notation x · ∂ = xµ ∂µ .
µ
For any continuous transformation, we can define the associated canonical current densities ja by
∂L ∂L δF
ν
δx
∂ν Φ − δν L
µ µ
ja = − · (1.29)
∂ (∂µ Φ) δ ωa ∂ (∂µ Φ) δ ωa
When the transformation under consideration is a symmetry and when the equations of motion are
satisfied, the currents (1.29) are conserved,
∂µ jaµ = 0 . (1.30)
To show this, it is useful to assume for a moment that the parameters ωa are arbitrary (but small)
space-time functions ωa (x). Using the formula det(1 + J) = 1 + tr J + O(J)2 where J = ∂ x0 /∂ x is
the jacobian matrix for the map x 7→ x0 , we find
∂ x0
δ xµ
= 1 + ∂ µ ωa + O(ω)2 . (1.31)
∂x δ ωa
Working to first order in ωa , we find after some algebra the following formula for the variation
δ S(Φ) ≡ S0 (Φ) − S(Φ),
!
δ xµ ∂L δ xρ ∂L δF
Z
d
δ S(Φ) = d x ∂µ ωa L− ∂ µ ωa ∂ρ Φ + ∂µ ωa . (1.32)
δ ωa ∂ ∂µ φ δ ωa ∂ ∂µ Φ δ ωa
µ
where the currents ja are given by (1.29). When the equations of motion are satisfied, the action is
by definition stationary under arbitrary variations and thus δ S = 0. Since the functions ωa in (1.33)
µ
are arbitrary, this is possible only if ∂µ ja = 0, completing the proof.
µ
To any current ja we associate a conserved charge Qa by the formula
Z
Qa = dd−1 x jad , (1.34)
Σ
µ
where Σ is the xd = constant hyper surface. When the current ja is conserved, the charge Qa is
also conserved in the sense that dQa /dxd = 0. Note that several currents may give rise to the same
charge. For example, if we define
j˜aµ = jaµ + ∂ν Bνa µ (1.35)
7
Introduction to Conformal Field Theory Antonin Rovai
for Ba = −Ba , then conservation of ja implies that j˜a is conserved as well. Moreover, the charge
µν νµ µ µ
PoS(Modave VIII)001
reduce to the identity and thus
∂L
∂ν Φ − δν L ,
µ µ
Tc ν = (1.36)
∂ (∂µ Φ)
where the subscript c indicates that this is the canonical energy-momentum tensor. Notice that the
µ
index a in the general definition of the current ja in (1.29) is now replaced by a space-time index
µ
µ, since translations are parameterised by a space-time vector. The associated current Tc ν , defined
in (1.36), thus carries two space-time indices.
The associated conserved charges also carry a space-time index, and are the components of
the energy-momentum vector Pµ : Z
Pµ = dd−1 x Tc dµ . (1.37)
Σ
For future reference, let us quickly write the currents associated to Lorentz and dilatation
invariance. From the definition (1.29) and the formulas derived in the examples of sections 1.1 and
1.2, we find the following expressions for the currents:
∂L
j µνρ = Tcµν xρ − Tcµρ xν + i Sνρ Φ (rotations) , (1.38)
∂ (∂µ Φ)
µ µ ν ∂L
jD = Tc νx + ∆Φ (dilatations) . (1.39)
∂ (∂µ Φ)
µν
In general, the canonical energy-momentum tensor Tc is not symmetric under the exchange
of µ with ν. In a Lorentz invariant theory however, it is however always possible to define a
µν
symmetric energy-momentum tensor by adding to Tc a suitable improvement term, following
µν
the discussion below (1.35). The resulting symmetric energy-momentum tensor TB is called the
Belinfante stress tensor, and is given by
µν
TB = Tcµν + ∂ρ Bρ µν , (1.40)
8
Introduction to Conformal Field Theory Antonin Rovai
µν µν
As a consequence, the constraints ∂µ j µνρ = 0 and ∂µ TB = 0 imply the symmetry of TB :
µν νµ
TB = TB . (1.43)
Let us remark that the Belinfante stress tensor TB µν is symmetric only when the equations of motion
µν
are satisfied, because we used the conservation laws ∂µ j µνρ = 0 and ∂µ TB = 0 to show (1.43).
PoS(Modave VIII)001
metric and generalises to curved space-times. Assume the energy-momentum tensor T µν has been
made symmetric. Under an arbitrary infinitesimal change of coordinates
x 7→ x + ε(x) , (1.44)
1
Z Z
d
δS = − d xT µν
∂µ εν = − dd x T µν (∂µ εν + ∂ν εµ ) . (1.45)
2
Now consider the generalization of this theory on a curved space-time of metric g, and let S[Φ; g]
be a diffeomorphism invariant action such that it reduces to our original action S when restricted
on flat space1 where the metric g is the identity matrix that we write δ ,
The action S[Φ; g] is by definition invariant under the infinitesimal diffeomorphism (1.44). Evalu-
ating its variation δ S on flat space, the requirement δ S[Φ, g] = 0 implies
δ S[Φ; g] 1 µν
Z
d
δ S[Φ; g] g=δ = d x δ gµν − T (∂µ εν + ∂ν εµ ) = 0 . (1.47)
δ gµν g=δ 2
Plugging the variation of the metric around flat space δ gµν = −∂µ εν − ∂ν εµ into (1.47) yields
δ S[Φ; g]
T µν = −2 · (1.48)
δ gµν g=δ
Formula (1.48) suggest the following definition for the energy-momentum tensor in curved space-
times:
2 δ S[Φ; g]
T µν (g) ≡ − √ (curved space-time) , (1.49)
g δ gµν
the factor involving the determinant of the metric being introduced in order to have a tensor (rather
than a density). Loosely speaking, the energy-momentum tensor captures the variation of the theory
under an arbitrary variation of the metric. Notice that it is formula (1.49) that is used to define the
energy-momentum tensor in general relativity, where Tµν appears as a source term in Einstein’s
equations.
1 The action S[Φ; g] is not uniquely defined by the condition (1.46), as any term that vanishes on flat space may be
added to it.
9
Introduction to Conformal Field Theory Antonin Rovai
PoS(Modave VIII)001
where Φ1 (x1 ), . . . , Φn (xn ) are fields (or operators) of the theory evaluated at some space-time points
x1 , . . . , xn ∈ Rd .
It should be kept in mind that if the classical theory has some symmetry, then in general this
does not mean that the quantized theory also has this symmetry, in the sense of (1.50). When the
symmetry is broken quantum mechanically, we say that the classical symmetry is anomalous. We
shall encounter an example of anomaly in section 3.4. In this section (and in most of these lectures),
we assume that the symmetries that we consider are not anomalous.
Let us write the infinitesimal version of the condition (1.50) for a continuous symmetry. Using
the path-integral formalism for example, it is a simple matter to show that if ωa is lifted to an
arbitrary function ωa (x), then
Z
δ hΦ(x1 ) · · · Φ(xn )i = dd x ∂µ h jaµ (x)Φ(x1 ) · · · Φ(xn )iωa (x)
Z n
= −i dd x ωa (x) ∑ hΦ(x1 ) · · · Ga Φ(xi ) · · · Φ(xn )iδ d (x − xi ) , (1.51)
D i=1
where the domain D ⊂ Rd contain all the points xi , i = 1, . . . , n at which the fields are evaluated.
Since this is valid for any value of the parameters ωa , we find the so-called Ward identities
n
∂
h
∂ xµ a
j µ
(x)Φ(x1 ) · · · Φ(xn )i = −i ∑ δ d (x − xi )hΦ(x1 ) · · · Ga Φ(xi ) · · · Φ(xn )i . (1.52)
i=1
Equations (1.52) are the quantum versions of the classical conservation law (1.30). Loosely speak-
µ
ing, one can say that in the quantum theory, the currents ja fail to be conserved when inserted in
correlation functions because of the terms on the right hand side of (1.52). Such terms, that are
µ
non-zero (and singular) only when the current ja is evaluated at one of the points xi , are called
contact terms.
µν µν νµ
Similarly to the classical case, the addition of an improvement term ∂ν Ba with Ba = −Ba
µ
to the current ja as in (1.35), has no consequence on the Ward identities (1.52). Indeed, thanks
to the antisymmetry property of Ba under the exchange on µ and ν, we can replace ja by j˜a =
µν µ µ
µ µν
ja + ∂ν Ba into equation (1.52) without any further changes.
10
Introduction to Conformal Field Theory Antonin Rovai
S doesn’t contain all the quantum effects of the theory. To state this more precisely, let us consider
the path-integral expression for the partition function Z[g],
Z
Z[g] = DΦg e−S[Φ;g] . (1.53)
PoS(Modave VIII)001
to keep the idea that the energy-momentum tensor captures the response of the theory under a
perturbation of the metric, it is very natural to define it through the equation
1 d √
Z Z Z
−S[Φ;g+δ g]
Z[g + δ g] = DΦg+δ g e = DΦg 1 + d x g T δ gµν e−S[Φ;g] ,
µν
(1.54)
2
where δ g is an arbitrary, small perturbation of the metric. If we define as usual the generating
functional W by
W [g] = − ln Z[g] , (1.55)
then the definition of T µν is equivalent to
2 δW [g]
hT µν i = − √ · (1.56)
g δ gµν
We will use equation (1.56) when we consider the trace anomaly in section 3.4, and it will also be
very useful in other lectures of this school [1]. Formula (1.56) and (1.49) are quite analogous, but it
should now be clear that in (1.56), the subtle quantum effects contained in the path-integral (1.53)
have been taken into account. Of course, if we consider the classical limit of our theory, then the
generating functional W goes to the classical action S of the theory, and we recover (1.49) from
(1.56).
In this section we define conformal transformations in any dimension d, and explore their basic
properties.
∂ x0ρ ∂ x0λ
gρλ (x0 ) = Λ(x)gµν (x) , (2.1)
∂ xµ x ∂ x ν
x
together with some field transformation Φ 7→ Φ0 to be specified later and where Λ is some strictly
positive function.
Let us insist on the fact that under a conformal transformation, the metric is not changed,
g 7→ g. As a side remark, let us mention that conformal transformations are sometimes defined as
11
Introduction to Conformal Field Theory Antonin Rovai
diffeomorphisms that leave the metric unchanged up to a factor. This formulation is misleading
since the metric is not changed under a conformal transformation, while for diffeomorphisms it
is. Since we measure distances before and after a conformal transformation with the same metric,
distances between points change, because the points have moved. On the other hand, distances are
invariant under a diffeomorphism. Moreover, a diffeomorphism-invariant theory is clearly not au-
tomatically conformally invariant: conformal transformations are not a subset of diffeomorphisms.
We will now study the consequences of (2.1) on the map x 7→ x0 for flat space, gµν = δµν .
The set of conformal transformations obviously form a group called the conformal group. Let us
already note that this group should have the Euclidean group as a subgroup,2 as it corresponds to
PoS(Modave VIII)001
the special case Λ(x) = 1. Otherwise, the Jacobian |∂ x0 /∂ x| of the conformal transformation is
Λ(x)−d/2 , as can be seen by evaluating the determinant of both sides of equation (2.1).
Consider an infinitesimal transformation x 7→ x + ε(x). Setting f = 1 − Λ, we find the con-
straint
∂µ εν + ∂ν εµ = f δµν , (2.2)
∂ρ2µ εν + ∂ρν
2
εµ = ∂ρ f δµν , (2.4)
which implies
2
2 ∂µν ερ = δµρ ∂ν f + δνρ ∂µ f − δµν ∂ρ f . (2.5)
(d − 1) ∂ 2 f = 0 . (2.7)
The case d = 1 is trivial: any coordinate transformation obviously satisfies (2.1). The case d = 2
is particular and will be studied in detail in the next section (although, as we will see then, some
results derived below will also apply to d = 2), and we focus on the case d ≥ 3 from now on.
2 f = 0 is solved by
The equation ∂µν
f (x) = A + Bµ xµ , (2.8)
for some constants aµ , bµν , cµνρ with cµνρ = cµρν . Let us analyse these transformations case by
case.
2 In the Minkowskian case, the conformal group contain the Poincaré group.
12
Introduction to Conformal Field Theory Antonin Rovai
It is always possible to decompose bµν as its symmetric part b(µν) plus its antisymmetric part
b[µν] ≡ mµν . Hence
ρ
b ρ
bµν = δµν + mµν . (2.11)
d
PoS(Modave VIII)001
It is not difficult to see that the first term consists of an infinitesimal dilatation x 7→ (1 + α)x,
µ
where α = b µ /d, while the last term corresponds to a rotation.
We have now explored all possibilities for conformal transformations at the infinitesimal level. To
find the finite transformations, we must exponentiate the different infinitesimal transformation that
we just found. Although this is straightforward in principle, it can be tedious (in particular for the
SCT). The result is
x0 = x µ + a µ
µ
(Translations) (2.13a)
0µ
x = λ xµ (Dilatations) (2.13b)
0µ µ ν
x = R νx (Rotations) (2.13c)
x µ − b µ x2 ,
x0 =
µ
(SCT) (2.13d)
1 − 2b · x + b2 x2
where λ , aµ and bµ are numbers and Rµ ν is a matrix that satisfies the constraint (1.6).
To get some intuition about an SCT, remark that under such a transformation x 7→ x0 we have
x0 µ xµ
0 2
= 2
− bµ , (2.14)
x x
so a SCT is nothing but the composition of an inversion, a translation characterised by bµ and again
an inversion.
Note that the inversion x → x/x2 satisfies (2.1) and hence is a conformal transformation. But
for this transformation to be defined at the origin xµ = 0, one should add to space-time an extra point
“at infinity” such that the origin is mapped onto it. Moreover, by composition with a translation,
any point may be mapped to this point at infinity, and hence it has nothing special: it is an extra
space-time point. In two dimensions d = 2, the resulting space-time C ∼ = R2 with this new point
“at infinity” has the topology of a sphere and is called the Riemann Sphere.
13
Introduction to Conformal Field Theory Antonin Rovai
Algebra
Following the general discussion on generators in section 1.2, we easily find the generators of
conformal transformations:
PoS(Modave VIII)001
The generators given in (2.15) are valid for fields invariant under conformal transformations, i.e.
for F = 0 in the notations of section 1.1. We will later on consider the action of these generators
on fields transforming non-trivially under conformal transformations, see section 2.2.
The algebra satisfied by these generators is
It is now manifest that the Euclidean algebra is a subalgebra of the above conformal algebra, since
the commutation relations of Pµ and Lµν do not involve any of the generators D or Kµ . We now
consider redefinitions of the generators that will allow us to identify the conformal group explicitly.
Consider the new generators {Jab } with a, b, . . . = −1, 0, 1, . . . , d defined as
η̃−1,−1 = −1 , (2.18a)
η̃00 = 1 , (2.18b)
η̃µν = δµν . (2.18c)
Remark that the matrix η̃ is of Minkowski signature (−1, +1, . . . , +1) when we consider Euclidean
space-time, while for Minkowskian space-time η̃ is of signature (−1, −1, +1, . . . , +1).
In term of η̃ and Jab , the conformal algebra (2.16) reads
[Jab , Jcd ] = i (η̃ad Jbc + η̃bc Jad − η̃ac Jbd − η̃bd Jac ) . (2.19)
14
Introduction to Conformal Field Theory Antonin Rovai
We thus conclude that the conformal group in Euclidean space is SO(d + 1, 1), while in Lorentzian
space-time it is SO(d, 2). In either case, a generic conformal transformation has (d + 1)(d + 2)/2
parameters.
PoS(Modave VIII)001
Pµ Φ(x) = −i∂µ Φ(x) , (2.20)
Lµν Φ(x) = i xµ ∂ν − xν ∂µ Φ(x) + Sµν Φ(x) , (2.21)
where Sµν are the spin-s representation matrices of the rotation group. Consider first the subgroup
of conformal transformations that leave the point x = 0 invariant, that is, the subgroup generated
by dilatations, rotations and SCTs. The matrices representing the action of these transformations
on Φ are denoted respectively by D̃, Sµν and K̃. These generators must satisfy
The action of the generators at any space-time point x is obtained by conjugation with the transla-
tion operator eix·P . The result may be easily derived using the Hausdorff formula:
1 1
e−A BeA = B + [B, A] + [[B, A], A] + [[[B, A], A], A] + · · · (2.23)
2 3!
Explicitly,
K̃Φ(x) = 0 , (2.24)
DΦ(x) = −ix · ∂ + D̃ Φ(x) , (2.25)
2
ν
Kµ Φ(x) = K̃µ + 2xµ D̃ − x Sµν − 2ixµ x · ∂ + ix ∂µ Φ(x) . (2.26)
We now restrict our attention to a field Φ(x) that belongs to an irreducible representation of the
rotation group. Since D̃ commutes with all the rotation generators Sµν , D̃ must be proportional to
the identity matrix by Schur’s lemma. By comparing the action of D̃ on Φ(x) with the definition of
scaling dimension given in section 1, we deduce that D̃ = −i∆.
To conclude, let us write the finite transformation for a spinless field φ (x) under a general
conformal transformation: φ → φ 0 with
∂ x0 −∆/d
φ 0 (x0 ) = φ (x) . (2.27)
∂x
Fields transforming this way are called quasi-primary fields.
15
Introduction to Conformal Field Theory Antonin Rovai
PoS(Modave VIII)001
will see, this may not be enough.
Let us define the virial V µ of a field Φ as the combination
∂L
Vµ = (δ µρ ∆ + iSµρ )Φ . (2.29)
∂ (∂ ρ Φ)
Then if V µ can be written as ∂λ σ λ µ for some tensor σ , we may construct an improvement term
such that the resulting modified energy-momentum tensor is traceless, and hence the theory is
conformal. Explicitly, if we define X λ ρ µν by
2 h λ ρ (µν)
X λ ρ µν = δ σ − δ λ µ σ (ρν) − δ λ µ σ (νρ) + δ µν σ (λ ρ)
d −2
1 λ ρ µν i
+ δ δ − δ λ µ δ ρν σ (κσ ) δκσ , (2.30)
d −1
then the modified stress tensor Tm µν is given by
1
Tm µν = Tc µν + ∂ρ Bρ µν + ∂λ ∂ρ X λ ρ µν , (2.31)
2
where B is defined in (1.41). Note that since X (µν)[ρλ ] = 0, Tm µν is also symmetric and satisfies
µ µ
∂µ jD = Tm µ. (2.32)
We then see explicitly that in this case, scale invariance implies conformal invariance. Moreover,
we get a new expression for the dilatation current, valid when the equations of motion are satisfied:
µ µ ν
jD = Tm νx . (2.33)
Let us mention that for two-dimensional quantum field theory and under some broad hypothe-
ses (including unitarity), rotation and scale invariance imply conformal invariance [9].
16
Introduction to Conformal Field Theory Antonin Rovai
Consider the canonical current associated to rotational invariance j µνρ . Assuming the stress
tensor T µν has been symmetrized, we have
j µνρ = T µν xρ − T µρ xν . (2.34)
PoS(Modave VIII)001
i=1
µ µ µν
where ∂i denote the derivative with respect to xi and Si are the generators of the Lorentz group
in the representation of the field Φi . For translations, the Ward identity reads
n
∂µ hT µν (x)Φ1 (x1 ) · · · Φn (xn )i = − ∑ δ d (x − xi )∂iν hΦ1 (x1 ) · · · Φn (xn )i . (2.36)
i=1
which is the quantum analog of the statement that the stress tensor is symmetric: as usual when
there are no anomalies for the considered classical invariances, the classical property of symmetry
T µν = T ν µ translates into a similar statement in terms of correlation functions up to the addition of
contact terms, as can be seen on the right hand side of (2.37). If we further assume that the stress
µ
tensor is traceless, then the dilatation current jD is
µ µ
jD = T ν xν , (2.38)
where ∆i is the scaling dimension of the field Φi . Equation (2.39) then implies
n
hT µ (x)Φ(x1 ) · · · Φ(xn )i = −i ∑ δ d (x − xi )∆i hΦ(x1 ) · · · Φ(xn )i ,
µ
(2.40)
i=1
which is the quantum analog of the classical statement that the stress tensor is traceless.
17
Introduction to Conformal Field Theory Antonin Rovai
where φ1 and φ2 are scalar operators that we assume to be quasi-primary and not necessarily ap-
pearing in the collection of fields {Φi }. Under a conformal transformation x 7→ x0 , the quantum
invariance condition (1.50) together with (2.27) imply
∂ x0 ∆1 /d ∂ x0 ∆2 /d
hφ1 (x1 )φ2 (x2 )i = hφ1 (x01 )φ2 (x02 )i . (2.42)
∂x x1 ∂x x2
PoS(Modave VIII)001
for some function f of one real variable. Focusing on a dilatation x 7→ λ x in formula (2.42), we
find that the function f defined in (2.43) must be such that
We now consider constraint (2.44) at the special point s = 1/λ . This yields the following formula
for f ,
C12
f (s) = ∆ +∆ , (2.45)
s 1 2
where C12 ≡ f (1) is an unknown constant. Plugging (2.45) into formula (2.43) yields
C12
hφ1 (x1 )φ2 (x2 )i = · (2.46)
|x1 − x2 |∆1 +∆2
If we consider now an SCT defined in (2.13d), it is easy to check that the constraint (2.42) implies
that the two-point function (2.41) is automatically zero for ∆1 6= ∆2 . Our final result is therefore
C12
hφ1 (x1 )φ2 (x2 )i = if ∆1 = ∆2 and 0 otherwise. (2.47)
|x1 − x2 |2∆1
In particular, a one-point function (obtained from (2.41) by taking one of the two operators to be
the identity) is non-zero only if the scaling dimension of the operator is zero,
Let us mention that it is possible to show that in a unitary CFT, identity is the only operator with
zero scaling dimension.
Let us now turn to the three-point function,
Using the constraints obtained from (1.50) applied to translations, rotations and dilatations, we find
that the three-point function (2.49) must take the form
(abc)
C123
hφ1 (x1 )φ2 (x2 )φ3 (x3 )i = a b xc
+ all permutations in (a, b, c) , (2.50)
x12 x23 13
18
Introduction to Conformal Field Theory Antonin Rovai
(abc)
where C123 are unknown numbers, the three numbers a, b, c satisfy the constraint a + b + c = ∆1 +
∆2 + ∆3 and xi j ≡ |xi − x j |. Applying now to (2.49) the constraint (1.50) for an SCT characterised
by a vector bµ as in (2.13d) yields
(abc)
C123 (γ1 γ2 )a/2 (γ2 γ3 )b/2 (γ1 γ3 )c/2
hφ1 (x1 )φ2 (x2 )φ3 (x3 )i = a xb xc
γ1∆1 γ2∆2 γ3∆3 x12 23 13
+ all permutations in (a, b, c) , (2.51)
PoS(Modave VIII)001
γi = 1 − 2bµ xi + b2 xi2 .
µ
(2.52)
We now turn to the study of conformal field theories in two dimensions. Actually, much of
what we derived in section 2 remains valid when d = 2, because the set of conformal transforma-
tions in two dimensions contains those satisfying (2.9). As we shall see, there are however much
more solutions to the constraints (2.1) in two dimensions.
∂ wρ ∂ wλ µν
δ ρλ = δ . (3.1)
∂ z µ ∂ zν
19
Introduction to Conformal Field Theory Antonin Rovai
There are two ways one can solve these equations for wµ (z): either we set
∂ w0 ∂ w1 , ∂ w0 ∂ w1 ,
PoS(Modave VIII)001
= − = 0 (3.4)
∂ z0 ∂ z1 ∂ z1 ∂z
or we can choose
∂ w0 ∂ w1 , ∂ w0 ∂ w1
= 1 = − · (3.5)
∂ z0 ∂z ∂ z1 ∂ z0
These a precisely the Cauchy-Riemann relations for (respectively) an holomorphic or an anti-
holomorphic function! It is then very convenient to define the complex variables z = z0 + iz1 and
z̄ = z0 − iz1 and extend space-time to a complex space-time, where z̄ is considered as independent
of z. We recover our original space-time by imposing the reality condition z∗ = z̄. We have the
usual complex analysis relations
1 1
z0 = (z + z̄) , z1 = (z − z̄) , (3.6a)
2 2i
1 1
∂z = (∂0 − i∂1 ) , ∂z̄ = (∂0 + i∂1 ) , (3.6b)
2 2
∂0 = ∂z + ∂z̄ , ∂1 = i(∂z − ∂z̄ ) . (3.6c)
Note that we did not yet imposed that the coordinate transformation is everywhere well defined,
and the Cauchy-Riemann relations are of course valid only when evaluated at points where the
functions (w(z), w̄(z̄)) exist. In the rest of these notes, when we say that “ f is holomorphic” (or
“anti-holomorphic”) we actually mean meromorphic (or anti-meromorphic), that is, holomorphic
(or anti-holomorphic) on the whole complexified space-time except maybe at some isolated points
where the function may have poles, and we write f (z) (or f (z̄) respectively). The set of conformal
transformations in two dimensions is hence infinite-dimensional, as it consists of all holomorphic
and anti-holomorphic maps of the Riemann sphere.
20
Introduction to Conformal Field Theory Antonin Rovai
Generators
We consider an infinitesimal transformation z 7→ z+ε(z), z̄ 7→ z̄+ ε̄(z̄), where ε is holomorphic
and ε̄ is antiholomoprhic. We assume that ε admits, at least in a non-empty open set, the following
Laurent series around zero:
∞
ε(z) = ∑ cn zn+1 , (3.9)
n=−∞
where cn are its Laurent coefficients, and similarly for ε̄ with some coefficient c̄n (that are not
a priori related to the coefficients cn ). Consider a scalar field φ = φ (z, z̄). Since we are in two
dimensions, the scaling dimension of φ is ∆ = 0, and hence under our infinitesimal conformal
PoS(Modave VIII)001
transformation we have
δ φ (z, z̄) = φ 0 (z, z̄) − φ (z, z̄) = −∂z φ (z, z̄)ε(z) − ∂z̄ φ (z, z̄)ε̄(z̄) . (3.10)
and hence
`n = −zn+1 ∂z , `¯n = −z̄n+1 ∂z̄ . (3.12)
Note that the two sets of generators `n and `¯n commute. The algebra defined by the relation (3.13a)
is a Witt algebra for the generators `n (and similarly for the `¯n ) and thus the algebra (3.13) of the
generators of conformal transformation in two dimensions at the classical level consists of two
copies of the Witt algebra.
21
Introduction to Conformal Field Theory Antonin Rovai
and similarly for `¯n , with n = −1, 0, 1. Algebra (3.15) is a subalgebra of the Witt algebra (3.13a).
Here are the infinitesimal and finite transformations corresponding to these generators (since the
algebra for the `¯n is similar to the one for the `n , we focus on the holomorphic part and similar
PoS(Modave VIII)001
results will apply for the anti-holomorphic part):
`−1 : z 7→ z + ε , z 7→ z + α , (3.16a)
`0 : z 7→ z + εz , z 7→ λ z , (3.16b)
z ,
`1 : z 7→ z + εz2 , z 7→ (3.16c)
1−βz
where ε, α, λ and β are complex parameters, ε being in addition infinitesimal. Combining the three
above transformations, we find that a general global holomorphic conformal transformation is of
the form
az + b ,
z→ (3.17)
cz + d
where the complex numbers a, b, c, d are such that ad − bc = 1. It is easy to check that the result of
the composition of two such transformations (parametrized, say, by a1 , b1 , c1 , d1 and a2 , b2 , c2 , d2 )
corresponds to a transformation parametrized by a, b, c, d with ad − cb = 1 and such that
! ! !
ab a2 b2 a1 b1
= . (3.18)
cd c2 d2 c1 d1
Global holomorphic conformal transformations hence form a group isomorphic to SL(2, C)/Z2
with matrix multiplication, the Z2 factor coming from the fact that flipping the sign of a, b, c, d
returns the same conformal transformation (3.17) and preserves the relation ad − bc = 1.
Similar results apply for anti-holomorphic global conformal transformations. If we reduce to
real space-time by imposing z̄ = z∗ , it can be shown that the resulting group for global conformal
transformations is SO(3, 1). Recall that for d ≥ 3, the conformal group is SO(d + 1, 1), and there-
fore we have shown that this result is actually also valid for d = 2. This means that all results
derived in section 2 from invariance under SO(d + 1, 1), as for instance the transformation law for
scalar fields, Ward identities and constraints on two- and three-point functions apply for d = 2 as
well.
Transformation laws for fields with arbitrary spin and primary fields
Consider a field Φ of spin s and scaling dimension ∆. We define for this field its holomorphic
weight h and anti-holomorphic weight h̄ by
∆+s , ∆−s ,
h= h̄ = (3.19)
2 2
22
Introduction to Conformal Field Theory Antonin Rovai
and consider a global conformal transformation z → w(z), z̄ → w̄(z̄). Then it can be shown3 that
the formula (2.27), valid for scalar fields only, generalizes to
−h −h̄
0 dw dw̄
Φ (w, w̄) = Φ(z, z̄) . (3.20)
dz dz̄
Again, fields transforming this way are termed quasi-primary fields. If, in addition, the field Φ
transforms as (3.20) for any conformal transformation (not necessarily global), then Φ is called a
PoS(Modave VIII)001
primary field. Obviously, a primary field is always quasi-primary, but the converse is not true. As
we will shortly see, the energy-momentum tensor of a generic CFT is an example of quasi-primary
that is not primary.
For future reference, we note that under a purely holomorphic conformal transformation z 7→
z + ε(z), we have for a primary field Φ that Φ 7→ Φ + δε Φ with
δε Φ = −ε∂ Φ − h∂ ε , (3.21)
We will now write the Ward identities for translation, rotation and scale invariance in two
dimensions. Note that the generators of rotations Sµν , being two by two antisymmetric matrices,
are automatically proportional to εµν . The proportionality factor is nothing but the spin s of the
representation: Sµν = s εµν . With this observation, the Ward identities (2.36), (2.37) and (2.40) are
(in any coordinates) equivalent to
n
∂µ hT µν (x)Φ1 (x1 ) · · · Φn (xn )i = − ∑ δ 2 (x − xi )∂iν hΦ1 (x1 ) · · · Φn (xn )i , (3.22a)
i=1
n
εµν hT µν (x)Φ1 (x1 ) · · · Φn (xn )i = −i ∑ δ 2 (x − xi )si hΦ1 (x1 ) · · · Φn (xn )i , (3.22b)
i=1
n
hT µ (x)Φ1 (x1 ) · · · Φn (xn )i = − ∑ δ 2 (x − xi )∆i hΦ1 (x1 ) · · · Φn (xn )i .
µ
(3.22c)
i=1
We now specialize to our complex coordinates (z, z̄) defined in (3.6). Using the standard formula
1¯1 1 1
δ 2 (x) = ∂ = ∂ , (3.23)
π z π z̄
3 To convince yourself, you might want to start by showing this formula for an infinitesimal transformation.
23
Introduction to Conformal Field Theory Antonin Rovai
2π∂z hTz̄z Φ1 (w1 , w̄1 ) · · · Φn (wn , w̄n )i + 2π∂z̄ hTzz Φ1 (w1 , w̄1 ) · · · Φn (wn , w̄n )i
n
1
= − ∑ ∂z̄ ∂w hΦ1 (w1 , w̄1 ) · · · Φn (wn , w̄n )i , (3.24a)
i=1 z − wi i
2π∂z hTz̄z̄ Φ1 (w1 , w̄1 ) · · · Φn (wn , w̄n )i + 2π∂z̄ hTzz̄ Φ1 (w1 , w̄1 ) · · · Φn (wn , w̄n )i
n
1
= − ∑ ∂z ∂w̄ hΦ1 (w1 , w̄1 ) · · · Φn (wn , w̄n )i , (3.24b)
i=1 z̄ − w̄i i
PoS(Modave VIII)001
2hTzz̄ Φ1 (w1 , w̄1 ) · · · Φn (wn , w̄n )i + 2hTz̄z Φ1 (z1 , z̄1 ) · · · Φn (zn , z̄n )i
n
= − ∑ δ 2 (z − wi )∆i hΦ1 (w1 , w̄1 ) · · · Φn (wn , w̄n )i , (3.24c)
i=1
−2hTzz̄ Φ1 (w1 , w̄1 ) · · · Φn (wn , w̄n )i + 2hTz̄z Φ1 (z1 , z̄1 ) · · · Φn (zn , z̄n )i
n
= − ∑ δ 2 (z − wi )si hΦ1 (w1 , w̄1 ) · · · Φn (wn , w̄n )i , (3.24d)
i=1
with a similar expression for T̄ . The quantity between brackets is hence holomorphic everywhere,
and in our notations we may replace T (z, z̄) by T (z). One should not forget, however, that T (z) fails
to be holomorphic when inserted at coincident point in a correlation function. Similarly, T̄ (z, z̄) is
replaced by T̄ (z̄).
24
Introduction to Conformal Field Theory Antonin Rovai
where the domain D contains all points xi at which the operators are inserted. This integral is not
zero only because of the singularities of the correlations function at coincident points. Performing
the derivative in (3.30) and decomposing the term with ∂µ εν in its symmetric and antisymmetric
part,4 we find for the integrand of (3.30)
PoS(Modave VIII)001
n
ε µν ∂µ εν
2 ∂ ·ε
= − ∑ δ (x − xi ) ε · ∂i + ∆i + si hΦ1 (x1 ) · · · Φn (xn )i . (3.31)
i=1 2 2
Thanks to the δ -functions in the right hand side of (3.31), the integral (3.30) is trivial to perform,
and (3.29) we end up with
Z
δε hΦ(x1 ) · · · Φ(xn )i = d2 x ∂µ hT µν (x)εν (x)Φ(x1 ) · · · Φ(xn )i . (3.32)
D
We now apply Gauss’s theorem and specialize to the complex coordinates defined in (3.6). Using
the definitions (3.27) of T and T̄ we finally find
1
I
δε,ε̄ hΦ(w1 , w̄1 ) · · · Φ(wn , w̄n )i = − dz ε(z)hT (z)Φ(w1 , w̄1 ) · · · Φ(wn , w̄n )i+
2πi C
1
I
dz̄ ε̄(z̄)hT̄ (z̄)Φ(w1 , w̄1 ) · · · Φ(wn , w̄n )i , (3.33)
2πi
where C = ∂ D is the boundary of D and we explicitly wrote that the variation depends on two
(independent) functions ε and ε̄. Using the general formula (1.51) relating the variation of a cor-
relator under a symmetry transformation and the corresponding conserved current, the currents
j(z) and j¯(z̄) associated with holomorphic and anti-holomorphic transformations respectively are
straightforwardly read-off from the Ward identity (3.33):
While we assumed that the fields Φ(xi ) appearing in the equation (3.33) are primaries, it may be
shown that the same formula holds for non-primaries. This will be clearer in the operator formalism
described in section 3.3.
For primaries, we may actually push further formula (3.33), thanks to equation (3.28) and
performing the contour integration using Cauchy’s theorem. In the case of a purely holomorphic
transformation z 7→ z + ε(z), we find
n
δε hΦ1 (x1 ) · · · Φn (xn )i = − ∑ ε(wi )∂wi + ∂ ε(wi ) hΦ1 (x1 ) · · · Φn (xn )i , (3.35)
i=1
25
Introduction to Conformal Field Theory Antonin Rovai
PoS(Modave VIII)001
ticular, a coordinate (“time”) is singled out. In Euclidean space-time however, there is no preferred
direction for a time-like coordinate. We shall hence choose it arbitrarily, although in other contexts
(like string theory for example) this is actually the natural choice: time will be the radial direction
in the complex plane. The resulting quantization is called radial quantization.
where the plus sign is selected if one of the operators Φ1 or Φ2 is a commuting-number valued
operator, while we chose the minus sign if both Φ1 and Φ2 are anticommuting-number valued.
Although we will not write explicitly the radial ordering operator R , it should be kept in mind that
it is present in any expression we consider.
Let us consider two arbitrary operators A and B, and assume they may be represented as
I I
A= dz a(z) , B= dz b(z) , (3.37)
o o
where a(z) and b(z) are some operators depending on z and the integrals are taken over a small
H
circle around the origin of the complex plane. If we denote by Cw the integral taken on a small
circle around a point w of complex Riemann sphere, we have
I I I
dz a(z)b(w) = dz a(z)b(w) − dz b(w)a(z) = [A, b(w)] , (3.38)
Cw Co+ Co−
where Co+ is a circle centred at the origin with radius slightly bigger that |w|, Co− is a circle also
centred at the origin but with radius slightly smaller that |w| and we have used the implicit presence
of the radial operator R on the left hand side. We therefore conclude that the commutator [A, B] of
the two operators A and B can be expressed, using a(z) and b(z), as
I I
[A, B] = dw dz a(z)b(w) . (3.39)
Co Cw
26
Introduction to Conformal Field Theory Antonin Rovai
We thus see that [A, B] essentially depends on the singular values of the product of the operators
a(z) with b(w) as z → w. This observation will be our motivation to define the operator product
expansion in subsection 3.3.2.
We introduce the following mode expansion5 by decomposing the energy-momentum opera-
tors T and T̄ as
∞ ∞
Ln L̄n
T (z) = ∑ n+2 , T̄ (z̄) = ∑ n+2 , (3.40)
n=−∞ z n=−∞ z̄
for some operators Ln , L̄m whose interpretation will be given in a moment. These relations may be
easily inverted using Cauchy’s theorem:
PoS(Modave VIII)001
1 1
I I
Ln = dz zn+1 T (z) , L̄n = dz̄ z̄n+1 T̄ (z̄) . (3.41)
2πi Co 2πi Co
We will now show that the operators Ln , L̄m are the conserved charges associated to conformal
transformations. Let us consider an infinitesimal holomorphic conformal transformation z 7→ z +
ε(z). In the radial quantization scheme, the associated conserved charge Qε , defined by (1.34),
reads
1 1
I I
Qε = dz jε (z) = dz ε(z)T (z) , (3.42)
2πi Co 2πi Co
where we used the explicit expression (3.34) for the current jε (z). Therefore, the associated varia-
tion δε Φ reads
1
I
δε Φ(w) = −[Qε , Φ(w)] = − dz ε(z)T (z)Φ(w) . (3.43)
2πi Co
We now decompose ε as
∞
ε(z) = ∑ zn+2 εn (3.44)
n=−∞
We deduce from (3.45) that the operators Ln are the conserved charges corresponding to conformal
invariance (for holomorphic transformations). A similar result holds for anti-holomorphic trans-
formations, and so the Ln , L̄m are the quantum analogues of `n , `¯m introduced in (3.11). In order
to determine the algebra satisfied by the operators Lm , L̄n , we first need to introduce the concept of
operator product expansion.
27
Introduction to Conformal Field Theory Antonin Rovai
of the integrand. To get some control on this, we assume that the product a(z)b(w) is given by (we
restrict to holomorphic operators for convenience)
N
{ab}n (w) ,
a(z)b(w) = ∑ n
(3.46)
n=−∞ (z − w)
for some operators {ab}n . The Operator Product Expansion is defined as the sum of the singular
terms in the series of the right hand side of (3.46), and we write
N
{ab}n (w) ,
a(z)b(w) ∼ ∑ (3.47)
PoS(Modave VIII)001
n
n=1 (z − w)
where by definition the right hand side is the OPE of the left hand side. As an equality between
operators, (3.47) should always be thought of as inserted in some correlator, possibly with other
fields inserted.
The first example of OPE is the OPE of T with a primary operator Φ, that we can infer from
equation (3.28):
h 1
T (z)Φ(w, w̄) ∼ 2
Φ(w, w̄) + ∂w Φ(w, w̄) . (3.48)
(z − w) z−w
What can we say about the OPE of T with a quasi-primary operator Φ? Under an infinitesimal
global conformal transformation z 7→ z + αz, we have
δ Φ(z, z̄) = − αz∂ Φ(z, z̄) + ᾱ z̄∂¯ Φ(z, z̄) + (hα + h̄ᾱ)Φ(z, z̄) .
(3.49)
On the other hand the Ward identity (3.33) with ε(z) = αz yields
1 1
I I
δ hΦ(z, z̄)i = − dw αwhT (w)Φ(z, z̄)i + dw̄ ᾱ w̄hT̄ (w̄)Φ(z, z̄)i (3.50)
2πi 2πi
= −αh{T Φ}2 (z, z̄) + z{T Φ}1 (z, z̄)i + ᾱh{T̄ Φ}2 (z, z̄) + z̄{T̄ Φ}1 (z, z̄)i . (3.51)
In conclusion, the most general form of the OPE of T with a quasi-primary field Φ is of the form
h ∂ Φ(w, w̄) ,
T (z)Φ(w, w̄) ∼ · · · + 2
Φ(w, w̄) + (3.54)
(z − w) z−w
and similarly for T̄ , where the dots denote extra terms that diverge like (z − w)−n with n ≥ 3. When
all these extra terms vanish, (3.48) implies that Φ is primary.
Since T is itself a quasi-primary (of weights (h, h̄) = (2, 0)), we can look at the case where Φ
is taken to be T in formula (3.54). It can be shown that in this case, thanks to unitarity and using
a dimensional argument, the only possible terms in addition to those already shown in (3.54) is
proportional to (z − w)−4 . As a consequence, the T T OPE is
c/2 2T (w) ∂ T (w) ,
T (z)T (w) ∼ 4
+ 2
+ (3.55)
(z − w) (z − w) z−w
28
Introduction to Conformal Field Theory Antonin Rovai
for some number c, called the (holomorphic) central charge of the CFT. The origin of this name
will be explained in the next subsection.
We similarly have, considering the anti-holomorphic operator T̄ , that
PoS(Modave VIII)001
c 3
δε T = − ∂ ε − 2T ∂ ε − ε∂ T . (3.57)
12
Comparing with (3.21), we see that except when c = 0, T is not a primary field. It is a nice exercise
to show that the finite transformation T 7→ T 0 coming from the infinitesimal transformation (3.57)
is, for z 7→ w(z), given by
−2
0 dw c
T (w) = T (z) − {w; z} , (3.58)
dz 12
where {w; z} is called the Schwarzian derivative of w with respect to z and is defined by
2
d3 w/dz3 3 d2 w/dz2
{w; z} ≡ − . (3.59)
dw/dz 2 dw/dz
29
Introduction to Conformal Field Theory Antonin Rovai
PoS(Modave VIII)001
must have7
hT µ µ i = aR , (3.61)
for some dimensionless constant a. Our goal is now to find the precise value of a. To this end,
we will compare to first order around flat space the transformation properties of both sides of (the
derivative of) (3.61) under an infinitesimal conformal transformation composed with a suitably
chosen diffeomorphism.
We use the usual complex coordinates (z, z̄) and start with the operator relation
a
Tzz̄ = gzz̄ R . (3.62)
2
Applying the operator ∇z to equation (3.62) yields
a
∇z̄ Tzz̄ = ∂z R , (3.63)
2
where we used the fact that ∇z = gzz̄ ∇z̄ . Using the conservation law (valid upon taking the expec-
tation value) ∇µ T µν = 0, we deduce that
a
∇z Tzz = − ∂z R . (3.64)
2
We now consider the variation of both sides of equation (3.64) under the composition of the fol-
lowing two transformations:
∂ x0
= 1+∂ε , (3.65)
∂x
1 c 3 1 ε
δε Tzz = − δε T = ∂ ε + T∂ε + ∂T , (3.66)
2π 24π π 2π
where we used equation (3.57).
6 The trace anomaly is also know as the conformal anomaly.
7 A word of caution: the arguments given here does not constitute a complete proof of equation (3.61). For a detailed
proof, we refer the reader to the literature, see e.g. [3].
30
Introduction to Conformal Field Theory Antonin Rovai
Under the composition of the two above transformations, the net effect on Tzz is simply
c 3
δtot Tzz = ∂ ε. (3.68)
24π
PoS(Modave VIII)001
Going to the conformal gauge gµν = e2ω δµν and working to first order in ω, we find that the
variation of the left hand side of (3.64) is given by
c ¯ 3
δtot ∇z Tzz = ∂∂ ε . (3.69)
12π
Let us now determine the effect on R of the same transformation. The net effect on the coordinates
is the identity, while the net effect on the metric is an infinitesimal Weyl rescaling,
g 7→ (1 + ∂ ε)g . (3.70)
The effect of an arbitrary Weyl rescaling g 7→ g̃ = e2σ g on the Ricci scalar is straightforward to
work out (see e.g. in [11] for a proof of this formula generalised to any dimension) and reads
√
g̃R(g̃) = g R(g) − 2∇2g σ ,
p
(3.71)
where ∇g is the covariant derivative compatible with the metric g. In the case of the infinitesimal
Weyl rescaling (3.70), we find using formula (3.71) the following expression for the variation δ R
of the Ricci scalar:
δ R = −R∂ ε − ∇2 ∂ ε . (3.72)
Moreover, in the conformal gauge gµν = eω δµν and in complex coordinates (z, z̄), the curved lapla-
cian ∇2 reads
∇2 = 4e−2ω ∂z ∂z̄ . (3.73)
Plugging (3.73) into (3.72) and working at first order in ω we find, for the net variation of the right
hand side of (3.64),
a
− δtot ∂z R = 2a∂ 3 ∂¯ ε . (3.74)
2
Finally, comparing equation (3.69) and (3.74), we find a = c/(24π). Our final result is thus
c
hT µ µ i = R. (3.75)
24π
Acknowledgements
Many thanks to Blagoje Oblak for his careful reading of the first version of these notes. The
author is a Research Fellow of the Belgian Fonds de la Recherche Scientifique-FNRS.
31
Introduction to Conformal Field Theory Antonin Rovai
References
[3] P. Di Francesco, P. Mathieu and D. Sénéchal, Conformal field theory, Springer, New York
1997.
PoS(Modave VIII)001
[4] P. H. Ginsparg, Applied Conformal Field Theory, hep-th/9108028.
[6] J. Polchinski, String theory. Vol. 1: An introduction to the bosonic string, Univ. Pr., Cambridge
UK, 1998.
[8] J. Polchinski, String theory. Vol. 2: Superstring theory and beyond, Univ. Pr., Cambridge UK
1998.
[9] J. Polchinski, Scale And Conformal Invariance In Quantum Field Theory, Nucl. Phys. B 303
(1988) 226.
[10] J. D. Brown and M. Henneaux, Central Charges in the Canonical Realization of Asymptotic
Symmetries: An Example from Three-Dimensional Gravity, Commun. Math. Phys. 104 (1986)
207.
[11] M. Nakahara, Geometry, topology and physics, Taylor & Francis, Boca Raton 2003.
32