You are on page 1of 9

Conformal field theory in the sense of Segal,

modified for a supersymmetric context


Paul S Green
January 27, 2014

1 Introduction
In these notes, we will review and propose some revisions to the definition
of a conformal field theory first articulated by Graeme Segal in a set of
notes which have had considerable informal circulation and influence since
1989, although they were published only in 2004 in a festschrift volume for
Professor Segal. The proposed revisions are in several directions.
• The explicit introduction of a supersymmetric version of Segal’s defi-
nition.
• The observation that a Riemann surface with exactly one incoming and
one outgoing boundary component with a marked point on each admits
a canonical parametrization of the boundary which gives the set of all
such surfaces the natural structure of a holomorphic semigroup. This
semigroup and its supersymmetric generalization will be seen to play
an important role in the theory.
• An important role for the Krichever-Novikov algebras and their super-
symmetric generalizations.

2 Review and reformulation of Segal’s defini-


tion of a CFT
The version of Segal’s definition that we give here is closer to his reformu-
lation in the preface to the published version of his notes than it is to the

1
original notes; however it is not identical with either.
A conformal field theory in the sense of Segal begins with a functor H from
the category of compact closed connected one-manifolds and orientation-
preserving diffeomorphisms to the category of Hilbert spaces and unitary
isomorphisms, satisfying:

(a) H(−C) = H(C), where −C denotes C with the opposite orientation.

(b) If C and C 0 are disjoint then H(C ∪ C 0 ) = H(C) ⊗ H(C 0 ).

(c) It is useful to adopt the convention that H(φ) = C, where φ denotes


the empty one-manifold.

The other part of the structure assigns to each Riemann surface X an


element A(X) ∈ H(∂(X)). Here, X is given the orientation determined by
its complex structure and ∂(X) denotes the oriented boundary of X. In
particular, if X is closed, then A(X) is a complex number. A satisfies:

(d) If f : X → X 0 is a holomorphic equivalence, then A(X 0 ) = H(f |∂X)(A(X)).

Before stating the next condition, we recall that if V and W are Hilbert
spaces, then the Hilbert tensor product V ⊗H W is naturally isomorphic to
the space linear functions from W to V of Hilbert-Schmidt type. We will
say a Hilbert-Schmidt operator A from W to V is of trace class if A ◦ U is of
trace class for some (and therefore for any) unitary equivalence U : V → W .
Note that the trace can be computed in the tensor product formulation from
the formula tr(v ⊗ w) =< v, U (w) >.1 This allows us to extend the definition
of trace class to a Hilbert tensor product of any number of factors.

(e) If C and C 0 are components of ∂X, and X̂ is the result of gluing C


to C 0 by the orientation reversing diffeomorphism f , so that ∂X =
∂ X̂ ∪ C ∪ C 0 , then A(X) is in the domain of the map from H(∂X)
to H(∂ X̂) induced by the pairing from H(C) ⊗ H(C 0 ) to C that takes
α ⊗ β to < β, H(f )(α) >, and its image is A(X̂). Here, <, > is the
Hermitian inner product in H(C 0 ).
1
It follows from this observation that elements of V ⊗H W that are not of trace class
are not expressible as finite sums of decomposables. The converse is not true.

2
We observe that the functor H induces an action of Diff(C) ⊗ C on H(C),
where Diff(C) denotes the Lie algebra of smooth vector fields on C. A con-
sequence of the properties already listed, that will play a crucial role in the
sequel, is

(f) Let χ be a holomorphic vector field on X, then H(χ|∂X)(A(X)) = 0.

2.1 Incoming and Outgoing Boundary Components


So far, we have taken the boundary of ∂X to be directed as the topological
boundary of S, with the orientation determined by the complex structure.
With that orientation, each component of the boundary is, by convention,
outgoing. We may choose to reverse the orientation of some boundary compo-
nents and, accordingly, regard them as incoming. In that case, we can regard
A(X) as a linear transformation of trace class from H(∂in X) to H(∂out X),
where ∂in X and ∂out X denote respectively the unions of the incoming and
outgoing boundary components.

3 SuperRiemann Surfaces
3.1 Definitions and basic properties
For our purposes, a super-Riemann surface is a Riemann surface X, with or
without boundary, together with a holomorphic line bundle S and a holo-
morphic pairing, {, } : S ⊗ S → TX . Such structures exist and are well
known to be parametrized by H 1 (X, Z2 ) if X is closed. To see that this
remains the case if X has a boundary, let {X, S, {, }} and {X, S 0 , {, }0 } be
super-Riemann surfaces with the same underlying Riemann surface X. Fix
a base point x∗ on X and let C be any curve in X beginning and ending at
x∗ . Let s and s0 be points on the fibres over x∗ of S and S 0 respectively such
that so that {s, s} = {s0 , s0 }. Since any complex line bundle over a circle
is trivial, S|C has a nowhere vanishing section σ taking the value s at x0 .
Let C be parametrized by x(t) with 0 ≤ t ≤ 1 so that x(0) = x(1) = x∗ .
Then there is a unique path σ 0 (t) in the total space of S 0 with σ 0 (0) = s0
and {σ 0 (t), σ 0 (t)} = {σ(x(t)), σ(x(t))}. Then σ 0 (1) = ±s0 , and the sign is
independent of the choice of σ. This construction defines an element of
Hom(π1 (X, x∗ ), Z2 ) = H 1 (X, Z2 ), and is precisely the obstruction to the
equivalence of the two structures.

3
We define a super-circle to be a circle, together with a real square root of
its (trivial) tangent bundle. Such a square root is either trivial or the Möbius
bundle. We denote the corresponding supercircles as T-circles or M-circles.
In either case, the circle is directed by the “squares” of spinors.
If {X, S, σ} is a super-Riemann surface and C is a smooth directed simple
closed curve in X then S|C has a real sub-bundle consisting of all s in the
fibers of S such that σ(s ⊗ s) is tangent to C in the given direction. This
gives C the structure of a super-circle.

Lemma 1 Let {X, S, σ} be a super-Riemann surface and let C bound in X.


Then the induced super-circle structure on C makes it an M-circle.

Proof: Without loss of generality, we may assume C is the boundary of


X. It follows that every spin structure on X extends uniquely to the closed
Riemann surface X ∪C D, where D is a standard disk sewn in along C. We
will continue to use S and σ to denote the extended structure. Let z be a
coordinate on D and let χ be a nowhere vanishing holomorphic section of
S|D. iz∂z is a holomorphic vector field on D that vanishes only at z = 0
and is tangent to C Then z∂z = f (z)σ(χ ⊗ χ), where f is holomorphic and
non-vanishing on D except for a simple zero at z = 0. It follows that f does
not have a single-valued square root on C; hence the induced super-circle
structure makes C an M-circle.
Let {S, σ} and {S 0 , σ 0 } be super-Riemann structures on X and let C
be a smooth simple closed curve in X. We say the two structures are C-
distinguished if C is a T-circle with respect to one and an M-circle with
respect to the other. It is evident from the foregoing discussion that this is
the case if and only if the cohomology obstruction to the equivalence of the
two structures does not annihilate C.

Corollary 1 Let {X, S, σ} be a super-Riemann structures. Then the number


of T-circles among the boundary components of X is even.

Proof: Let X̂ be X with all boundary components capped off. Then if {S, σ}
extends to X̂, all the boundary components of X are M-circles since they all
bound in X̂. Otherwise, the cohomology obstruction to the equivalence of
{S, σ} with a structure that does extend to X̂ annihilates the boundary of
X. The corollary follows. 
We end this subsection by pointing out that the automorphism group
for a connected super-circle or super-Riemann surface is a central extension

4
by Z2 of a subgroup of the automorphism group of the underlying circle or
Riemann surface. The central Z2 is generated by multiplication by −1 in
the spinor bundle. In the sequel, we will confine ourselves to super-Riemann
surfaces all of whose boundary components are M-circles.

3.2 The super-Lie algebra of a super-Riemann surface


We begin by recalling that a super-Lie algebra L consists of a Lie algebra L0 ,
an L0 -module L1 and a symmetric pairing from L1 ⊗L1 to L0 , with respect to
which the elements of L0 act as derivations. We will emphasize this point of
view by writing the action of L0 on L1 simply as adjacency with the element
of L0 on the left. Finally, it is required that the triple product from L1 to
L0 given by {α, β}γ satisfies a Jacobi identity, which is equivalent to the
requirement that {α, α}α = 0.
We define the super-Lie algebra associated with either a super-circle or
a super-Riemann surface. Let X = {X, S, {, }} (resp C = {C, S, {, }}) be a
super-Riemann surface (resp super-circle). We define the super-Lie algebra
L(X) (resp L(C))associated with it by setting L0 to be the Lie algebra of
holomorphic vector fields on X (resp smooth vector fields on C) and L1 to be
the complex (resp real) vector space of holomorphic (resp smooth) sections
of S. The pairing {, } is the fiberwise bi-linear pairing already defined as part
of the super-Riemann surface (resp super-circle) structure. What remains to
be defined is the action of L0 on L1 . We use sigma for a generic section of
S in either context and recall that we must have {σ, σ}σ = 0. We complete
the definition by setting
1
(f {σ, σ})(gσ) = (f {σ, σ}g − ({σ, σ}f )g)σ,
2
where f and g are holomorphic (resp smooth real-valued) functions. This
definition, while not fiberwise, is local and is independent of the choice of σ.

Lemma 2 L(X) and L(C) as defined above are super Lie algebras over C
and R respectively. Moreover, if X is a super-Riemann surface and C is
an smooth directed simple closed curve in X with the induced super-circle
structure, then there is a natural restriction homomorphism of super-algebras
from L(X) to L(C) ⊗ C.

5
3.3 Representations of super-Lie algebras
Let L be a super-Lie algebra and let V = V0 ⊕ V1 be a Z2 -graded vector
space. A representation ρ of L on V consists of representations of L0 on V0
and V1 together with pairings L1 ⊗ V0 → V1 and L1 ⊗ V1 → V0 such that
• The elements of L0 act as derivations of the pairings involving L1 .

• 2ρ(α)2 = ρ({α, α}) for α ∈ L1 .


We will be particularly interested in the case in which L admits a complex
conjugation, V is a complex Hilbert space in which V1 and V0 are orthogonal,
and

• ρ(α)† = iρ(α) for α ∈ L1 .

• ρ(χ)† = −ρ(χ) for χ ∈ L0 .

We remark that a tensor product V ⊗ V 0 of Z2 -graded vector spaces is


graded by setting (V ⊗V 0 )0 = V0 ⊗V00 +V1 ⊗V10 and (V ⊗V 0 )1 = V1 ⊗V00 +V0 ⊗V10 .

3.4 Proposed definition of a super-conformal field the-


ory
The definition of a conformal field theory, given above, can easily be modified
to define a super conformal field theory. In the first place, the functor H
becomes a functor from the category of disjoint unions of super-circles and
super-diffeomorphisms to the category of Z2 -graded Hilbert spaces, and the
action of Diff(C) on H(C) is extended to an action of L(C). The only change
to (a)-(c) is that the tensor product becomes a graded tensor product.
We require A(X) ∈ H(∂X)0 , where X is now a super-Riemann surface,
require f in (e) to be a superdiffeomorphism, and extend (f) to all elements
of L(X).

4 Krichever-Novikov algebras and supersym-


metric generalizations
We recall that the Krichever-Novikov algebra KN(X, F ) of a closed Riemann
surface X with respect to a distinguished finite subset F is the Lie Algebra

6
of meromorphic fields holomorphic on the complement of the distinguished
subset. There is an immediate and straightforward extension to the case of a
closed super-Riemannian manifold, again with a distinguished finite subset.
Let X be a closed Riemann surface and F a finite subset. Let the bundle
S define a spin–structure on X. We will define the super-Krichever-Novikov
algebra SKN(X, F ) by setting SKN0 (X, F ) = KN(X, F ) while SKN1 (X, F )
consists of meromorphic sections η of S.

5 The semigroup of twice bounded Riemann


surfaces
One of the principal difficulties in working with Segal’s definition of a con-
formal field theory is the relative intractability of the infinite dimensional
diffeomorphism group of the circle. A partial way around this difficulty is
provided by the fact that a Riemann surface with two boundary curves and
a preferred point on each admits a canonical parametrization, which makes
the set of such Riemann surfaces into a holomorphic semigroup G, that acts
on H(S 1 ) where H is the Hilbert space valued functor of a conformal field
theory and S 1 is the unit circle in C.
More explicitly, let X be a Riemann surface with two boundary compo-
nents. We direct one of them as incoming and the other as outgoingWe also
choose points xin and xout on ∂in (X) and ∂out (X) respectively. We now ob-
serve that there is a uniqueRharmonic function
R t on X such that t|∂in (X) = 0,
t|∂out (X) is constant, and ∂out (X) ∗dt = ∂in (X) ∗dt = 2π.
The one-form ∗dt, together with the choice of xin and xout define canon-
ical diffeomorphisms of each boundary component of X with the standard
unit circle S 1 in C. We use θ as a parameter on each circle, taking θ = 0 at
the marked point and dθ = ∗dt. Using these identifications, one can interpret
A(X) as a trace-class operator on H = H(S 1 ), and it follows that a confor-
mal field theory induces a representation of the semigroup G by trace-class
operators on a Hilbert space. Moreover, observation (f) of Section 2 tells
us that KN (X̂, 0, ∞) partially intertwines the actions of Diff(S 1 ) with G on
H, where X̂ is the closed Riemann surface obtained from X by using the
canonical parametrizations to sew disks to both boundary components, and
0 and ∞ are the respective centers of the disks sewn to the incoming and
outgoing boundary components.

7
An important sub-semigroup of G is the punctured open unit disk D◦ .For
any conformal field theory, H decomposes into orthogonal subspaces on which
z ∈ D◦ operates as multiplication by z p z q where p and q are non-negative
real numbers whose difference is an integer. We will concern ourselves mainly
with chiral theories, for which only
S the case q = 0 occurs.
We observe also that G = g Gg where Gg denotes the component of
G consisting of Riemann surfaces of genus g. Gg is a complex manifold of
dimension 3g + 1, the composition is additive with respect to genus, and the
composition Gg × Gg0 → Gg+g0 is holomorphic. In particular, G0 = D◦ .
We can define an analogous semigroup of superRiemann surfaces, recall-
ing that we are requiring the boundary components to be M-circles. In
addition to the marked point on each boundary component, we must choose
a square root in S of ∂θ at each marked point. With these choices made,
we obtain a semigroup G s . Moreover, G s g is a 2g+1 -fold covers of Gg . In
particular, G s 0 = Ds is the non-trivial two-fold cover of D◦ .

5.1 The Hyperelliptic Semigroup


We recall that a hyperelliptic Riemann surface is one that can be realized as
a branched double cover of the Riemann sphere. We will consider specifically
double covers of the form {(z, w)|w2 = zp(z), 0 < |z0 | ≤ |z| ≤ |z∞ } where
p(z) is a polynomial of degree 2g without multiple roots, and all the roots
of p(z) have absolute values strictly between |z0 | and |z∞ |. the Riemann
surface thus defined has genus g, and two boundary components respectively
double-covered by the circles |z| = |z0 | and |z| = |z∞ |. In this case, we can
set dt = 21 log | zz0 |, and we choose the marked points on the two boundary
components to lie over z0 and z∞ respectively. The deck transformation
reduces this from a fourfold S choice to a twofold choice. This construction
defines a sub-semigroup S = Sg of G with S0 = G0 = D◦ .
One of the reasons that the hyperelliptic semigroup is of interest is the
form of the Krichever-Novikov algebra of a hyperelliptic curve. We are in-
terested in meromorphic vector fields on a hyperelliptic Riemann surface S
with poles only at one or both of the points w = z = 0 and w = z = ∞.
Such vector fields have the form z k wa (2w∂z + p0 (z)∂w ), where k is any inte-
ger and a is 0 or 1. Vector fields for which a = 1 are invariant under the
deck transformation, while those for which a = 0 are reversed by the deck
transformation. Only among the latter are the squares of spinors, which we
take to be sections of a preferred spinor bundle corresponding to the divisor

8
(1 − g)P where P is the point w = z = ∞ and g is the genus of the hyper-
elliptic curve in question. In particular, a spinor field corresponding to the
divisor (1 − g)P is a square root of 2w∂z + p0 (z)∂w .

You might also like