Conformal field theory and the cohomology of the moduli space of stable bundles

Similar documents
Three Descriptions of the Cohomology of Bun G (X) (Lecture 4)

k=0 /D : S + S /D = K 1 2 (3.5) consistently with the relation (1.75) and the Riemann-Roch-Hirzebruch-Atiyah-Singer index formula

Chern numbers and Hilbert Modular Varieties

Flat connections on 2-manifolds Introduction Outline:

On the intersection theory of the moduli space of rank two bundles

The Yang-Mills equations over Klein surfaces

Math 248B. Applications of base change for coherent cohomology

The tangent space to an enumerative problem

Higgs Bundles and Character Varieties

Lecture on Equivariant Cohomology

Constructing compact 8-manifolds with holonomy Spin(7)

REAL INSTANTONS, DIRAC OPERATORS AND QUATERNIONIC CLASSIFYING SPACES PAUL NORBURY AND MARC SANDERS Abstract. Let M(k; SO(n)) be the moduli space of ba

Donaldson Invariants and Moduli of Yang-Mills Instantons

Knot Homology from Refined Chern-Simons Theory

As always, the story begins with Riemann surfaces or just (real) surfaces. (As we have already noted, these are nearly the same thing).

Math 797W Homework 4

The Strominger Yau Zaslow conjecture

INSTANTON MODULI AND COMPACTIFICATION MATTHEW MAHOWALD

RIEMANN S INEQUALITY AND RIEMANN-ROCH

HYPERKÄHLER MANIFOLDS

CHARACTERISTIC CLASSES

Holomorphic line bundles

Coherent sheaves on elliptic curves.

Mini-Course on Moduli Spaces

Geometric Langlands duality and the equations of Nahm and Bogomolny

REPRESENTING HOMOLOGY AUTOMORPHISMS OF NONORIENTABLE SURFACES

The Canonical Sheaf. Stefano Filipazzi. September 14, 2015

Chern-Simons gauge theory The Chern-Simons (CS) gauge theory in three dimensions is defined by the action,

REPRESENTATION THEORY WEEK 7

η = (e 1 (e 2 φ)) # = e 3

LECTURE 5: SOME BASIC CONSTRUCTIONS IN SYMPLECTIC TOPOLOGY

Invariance Theory, the Heat Equation, and the Atiyah-Singer Index Theorem

Lecture VI: Projective varieties

Geometry of Conformal Field Theory

Ω Ω /ω. To these, one wants to add a fourth condition that arises from physics, what is known as the anomaly cancellation, namely that

Period Domains. Carlson. June 24, 2010

Homotopy and geometric perspectives on string topology

arxiv: v2 [math.ag] 5 Jun 2018

Chapter 2 Linear Transformations

Demushkin s Theorem in Codimension One

14 From modular forms to automorphic representations

SYMPLECTIC GEOMETRY: LECTURE 5

Linear Algebra. Min Yan

SPECTRAL ASYMMETRY AND RIEMANNIAN GEOMETRY

Porteous s Formula for Maps between Coherent Sheaves

Representations of algebraic groups and their Lie algebras Jens Carsten Jantzen Lecture III

THE HOMOTOPY TYPES OF SU(5)-GAUGE GROUPS. Osaka Journal of Mathematics. 52(1) P.15-P.29

Conformal field theory in the sense of Segal, modified for a supersymmetric context

L E C T U R E N O T E S O N H O M O T O P Y T H E O R Y A N D A P P L I C AT I O N S

Algebraic Geometry Spring 2009

Characteristic classes in the Chow ring

Determinant lines and determinant line bundles

MILNOR SEMINAR: DIFFERENTIAL FORMS AND CHERN CLASSES

NOTES ON HOLOMORPHIC PRINCIPAL BUNDLES OVER A COMPACT KÄHLER MANIFOLD

Lecture 1. Toric Varieties: Basics

Res + X F F + is defined below in (1.3). According to [Je-Ki2, Definition 3.3 and Proposition 3.4], the value of Res + X

LECTURE 6: J-HOLOMORPHIC CURVES AND APPLICATIONS

RIEMANN SURFACES: TALK V: DOLBEAULT COHOMOLOGY

MAT 445/ INTRODUCTION TO REPRESENTATION THEORY

The symplectic structure on moduli space (in memory of Andreas Floer)

NOTES ON FIBER BUNDLES

Qualifying Exams I, 2014 Spring

CHAPTER 1. TOPOLOGY OF ALGEBRAIC VARIETIES, HODGE DECOMPOSITION, AND APPLICATIONS. Contents

16.2. Definition. Let N be the set of all nilpotent elements in g. Define N

Math. Res. Lett. 13 (2006), no. 1, c International Press 2006 ENERGY IDENTITY FOR ANTI-SELF-DUAL INSTANTONS ON C Σ.

MAPPING CLASS ACTIONS ON MODULI SPACES. Int. J. Pure Appl. Math 9 (2003),

Topics in Representation Theory: Fourier Analysis and the Peter Weyl Theorem

Monodromy of the Dwork family, following Shepherd-Barron X n+1. P 1 λ. ζ i = 1}/ (µ n+1 ) H.

On the geometric Langlands duality

Poles of Instantons and Jumping Lines of Algebraic Vector Bundles on P

THE HOMOTOPY TYPE OF THE SPACE OF RATIONAL FUNCTIONS

INTRODUCTION TO THE HODGE CONJECTURE

Math 210C. The representation ring

The Real Grassmannian Gr(2, 4)

Mathematical Research Letters 2, (1995) A VANISHING THEOREM FOR SEIBERG-WITTEN INVARIANTS. Shuguang Wang

Representation theory and quantum mechanics tutorial Spin and the hydrogen atom

THE S 1 -EQUIVARIANT COHOMOLOGY RINGS OF (n k, k) SPRINGER VARIETIES

Parameterizing orbits in flag varieties

ARITHMETICALLY COHEN-MACAULAY BUNDLES ON THREE DIMENSIONAL HYPERSURFACES

Each is equal to CP 1 minus one point, which is the origin of the other: (C =) U 1 = CP 1 the line λ (1, 0) U 0

THE REPRESENTATION THEORY, GEOMETRY, AND COMBINATORICS OF BRANCHED COVERS

Representation Theory

ON NEARLY SEMIFREE CIRCLE ACTIONS

ON A THEOREM OF CAMPANA AND PĂUN

Kähler manifolds and variations of Hodge structures

Refined Chern-Simons Theory, Topological Strings and Knot Homology

Morse theory and stable pairs

Pacific Journal of Mathematics

Towards a modular functor from quantum higher Teichmüller theory

Highest-weight Theory: Verma Modules

RIEMANN SURFACES. max(0, deg x f)x.

If F is a divisor class on the blowing up X of P 2 at n 8 general points p 1,..., p n P 2,

LECTURE 7: STABLE RATIONALITY AND DECOMPOSITION OF THE DIAGONAL

NONNEGATIVE CURVATURE AND COBORDISM TYPE. 1. Introduction

Linear connections on Lie groups

Homological mirror symmetry via families of Lagrangians

THE QUANTUM CONNECTION

Notes 10: Consequences of Eli Cartan s theorem.

Riemann Surfaces and Algebraic Curves

GEOMETRIC STRUCTURES OF SEMISIMPLE LIE ALGEBRAS

Transcription:

Conformal field theory and the cohomology of the moduli space of stable bundles Michael Thaddeus Mathematical Institute, 24-29 St Giles, Oxford OX 3LB England Introduction Let Σ g be a compact Riemann surface of genus g 2, and let Λ be a line bundle over Σ g of degree. Then the moduli space of rank 2 stable bundles V over Σ g such that Λ 2 V = Λ was shown by Seshadri [9] to be a nonsingular projective variety N g. Its rational cohomology ring H (N g ) is exceedingly rich and despite years of study has never quite been computed in full. The object of this paper is to give an essentially complete characterization of this ring, or at least to reduce the problem to a matter of linear algebra. In particular, we find a proof of Newstead s conjecture that p g (N g ) = 0. We also obtain a formula for the volume of N g, which can be regarded as a twisted version of the formula for the degree 0 moduli space recently announced by Witten. The approach we shall take is not from algebraic geometry but from mathematical physics: it relies on the SU(2) Wess-Zumino-Witten model, which is a functor Z k associating a finite-dimensional vector space to each Riemann surface with marked points. The relationship with the moduli space is that when the level k of the functor is even, the vector space associated to Σ g with no marked points can be identified with H 0 (N g ; L k/2 ), where L is a fixed line bundle over N g. Now the work of Verlinde provides us with a means of calculating the dimension of any vector space arising from our functor, and in particular dim H 0 (N g ; L k/2 ), which we shall denote D(g, k). On the other hand Newstead [4] found explicit generators for H (N g ), and we can also express D(g, k) in terms of them using a Riemann-Roch theorem. Equating the two formulas enables us to evaluate any monomial in the generators on the fundamental class of N g, and by Poincaré duality this is sufficient, at least in principle, to determine the ring structure of H (N g ). In the discussion above one important point has been skated over. The SU(2) WZW model is of course associated to bundles of degree 0, not degree, so in order to make use of Verlinde s work it is necessary to formulate a twisted version of the field theory. This is carried out in 2, but the crucial properties of the twisted theory, analogous to those which make the ordinary theory a modular functor, are not proved in this paper. Rather, we will confine ourselves to exploring the consequences of these claims, and hope to return to justify them in a later paper. An outline of the remaining sections goes as follows. In 3 we review those parts of Verlinde s work we shall need, show how they must be modified in the twisted case, and work out some explicit formulas for D(g, k). In 4 we study the cohomology of N g, which we regard throughout as a space of representations via the theorem of Narasimhan and Seshadri. We define Newstead s generators α, β, and ψ i of H (N g )

and, using Poincaré duality, show that the entire ring structure is determined by those pairings of the form (α m β n γ p )[N g ] where γ = 2 ψ i ψ i+g. A further application of Poincaré duality enables us to eliminate γ recursively from the pairings, so that we need only evaluate (α m β n )[N g ]. This is calculated in 5 using the results of 3 and the Hirzebruch-Riemann-Roch theorem. Finally we extract Newstead s conjecture and the volume formula from our results. I am very grateful to Simon Donaldson for suggesting the basic approach and most of the geometrical arguments, and to Graeme Segal for explaining Verlinde s work. 2 The WZW model We begin with a speedy review of the SU(2) Wess-Zumino-Witten model and its gluing rules. The reader is referred to [6], [7], or [7] for a more leisurely account. We must first fix a positive integer k, called the level. Let Σ g be a compact connected Riemann surface of genus g, with marked points x,..., x p Σ g. We choose a labelling of the marked points, that is, we associate to each one an irreducible representation V of SL(2, C) = SU(2) C with dim C V k +. Such representations are determined up to isomorphism by their dimension: we will write V n for the representation of dimension n +, and denote by (Σ g ; x i, n i ) the Riemann surface with points x i labelled by V ni. To each such labelled Riemann surface we wish to associate a complex vector space Z k (Σ g ; x i, n i ). Let P be a smooth principal SL(2, C) bundle over Σ g ; then P is unique up to isomorphism, and the set A of compatible holomorphic structures on the associated rank 2 vector bundle U is an affine space modelled on Ω 0, C (Σ g ; ad P ). There exists a line bundle L A whose fibre at any holomorphic structure is the determinant line of the associated -operator on U. It is easy to see that the usual action of the gauge group G = Γ(Ad P ) on A lifts to L, so we have an action of G on the tensor power L k. Next, at each marked point x i, choose an identification of the fibre (Ad P ) xi with SL(2, C). (Actually, Ad P always has a global trivialization in the present case, so we may as well just use a restriction of that.) Then G also acts on each V ni via evaluation at the point x i. We may now define our vector space at level k as a space of G -equivariant sections: Z k (Σ g ; x i, n i ) = Γ G (L k i V ni ). If the genus g is at least 2, then the set A s of semistable holomorphic structures has complement of codimension at least 2 (see p. 569 of []), so we may regard Z k (Σ g ; x i, n i ) as the space of holomorphic sections of the twisted quotient L k As G V ni, which is a vector bundle over the moduli space A s /G of semistable SL(2, C) bundles over Σ g. However, some care is needed as the moduli space has singularities. In fact, all we will really want to know about Z k (Σ g ; x i, n i ) is its dimension, which we shall denote z k (Σ g ; x i, n i ). Now it is well known to physicists that Z k is an example 2 i

of a modular functor. In particular, it satisfies the two propositions below, which have important implications for the computation of z k (Σ g ; x i, n i ). () Proposition. The value of z k (Σ g ; x i, n i ) is independent of the complex structure on Σ g. Consequently, it is independent of the positions of the x i and of the trivializations of the fibres over them, because any two choices can be exchanged by a suitable smooth map. Thus we will be justified in henceforth writing simply z k (Σ g ; n i ) for z k (Σ g ; x i, n i ). The second proposition gives two gluing axioms, which describe the behaviour of z k under (i) the addition of a handle and (ii) connect-sum. In light of (), we don t need to worry about the choice of complex structure. (2) Proposition. k (i) z k (Σ g+ ; n i ) = z k (Σ g ; n, n, n i ). n=0 k (ii) z k (Σ g+g ; n i, n j ) = z k (Σ g ; n, n i ) z k (Σ g ; n, n j ). n=0 Proofs are essentially contained in [20]. The definition of Z k given above is well known. However, as it stands the definition involves SL(2, C) bundles over Σ g, which are topologically trivial. Since we are interested in non-trivial bundles over Σ g, we shall need to modify and extend the definition slightly. Fortunately, there is a straightforward way to do this. We consider (Σ g ; x i, n i ) as before. Let ˆP be a smooth principal PSL(2, C) bundle over Σ g with w 2 ( ˆP ) 0; again, ˆP is unique up to isomorphism. The associated CP bundle lifts to a rank 2 vector bundle Û of degree. Let  denote the set of compatible holomorphic structures on Û which induce a fixed holomorphic structure ω on Λ2 Û. (We didn t need to make this explicit in the untwisted case, because Λ 2 U is canonically trivial.) As before we let ˆL be the determinant line bundle. However, we need to be careful in our definition of the gauge group; we let Ĝ = Γ(Ãd ˆP ), where Ãd ˆP is the adjoint bundle with fibre SL(2, C), not PSL(2, C). As in the untwisted case, we choose identifications of (Ãd ˆP ) xi with SL(2, C), and define Ẑ k (Σ g ; x i, n i ) = ΓĜ( ˆL k i V ni ). Again, we write ẑ k (Σ g ; x i, n i ) = dim Ẑk (Σ g ; x i, n i ). In analogy with the properties of z k discussed above, we propose the following. (3) Claim. The value of ẑ k (Σ g ; x i, n i ) is independent of the complex structure on Σ g. Consequently, it is again independent of the positions of the x i and the trivializations of the fibres, as well as the choice of the holomorphic structure ω on Λ 2 Û. As in the untwisted case, we shall henceforth write simply ẑ k (Σ g ; n i ) for ẑ k (Σ g ; x i, n i ). 3

(4) Claim. k (i) ẑ k (Σ g+ ; n i ) = ẑ k (Σ g ; n, n, n i ). n=0 k (ii) ẑ k (Σ g+g ; n i, n j ) = ẑ k (Σ g ; n, n i ) z k (Σ g ; n, n j ). n=0 Despite their crucial importance for us, we will not attempt to prove these claims here. However, it appears that the proofs of the analogous statements given by Tsuchiya et al. [20] could be adapted to this case. To prove part (ii) of (2), for example, they define a sheaf S over a local universal family of curves of genus g + g near a curve Σ g Σ g with an ordinary double crossing. The sheaf is so constructed that its stalk over the generic smooth curve is but over the singular curve is S Σg+g = Z k (Σ g+g ; n i, n j ), k S Σg Σg = Z k (Σ g ; n, n i ) Z k (Σ g ; n, n j ). n=0 The desired result is then obtained by showing that S is locally free. The key step in adapting such a proof to the twisted case would be to note that there exists a principal PSL(2, C) bundle P over the local universal family whose restriction to the generic curve is twisted, but whose restriction to the singular curve Σ g Σ g is twisted over Σ g, but untwisted over Σ g. Of course, similar reasoning would suggest that (5) k z k (Σ g+g ; n i, n j ) = ẑ k (Σ g ; n, n i ) ẑ k (Σ g ; n, n j ), n=0 and we shall see in the next section that this is indeed the case. 3 Some calculations In one respect the twisted theory described above is simpler than the familiar untwisted one: if the genus g is at least 2, then not only does the set  s have complement of codimension at least 2, as before, but now the moduli space  s /Ĝ is known [9] to be a compact complex manifold N g. The section I Ĝ now acts nontrivially on ˆL as well as the V ni, so that ˆL k V ni descends to N g if and only if k + n i is even. When this is the case, we need have no qualms about identifying Ẑk (Σ g ; n i ) with the space of holomorphic sections of a vector bundle over N g. In particular, let L denote the line bundle over N g such that ˆL2 descends to L. Then if there are no marked points at all, we have for even k Ẑ k (Σ g ) = H 0 (N g ; L k/2 ). This is the link which will enable us to use Ẑk to study the cohomology of N g. Hence our object in this section will be to give an explicit formula for ẑ k (Σ g ). 4

From (4) we deduce ẑ k (Σ g ) = n ẑ k (Σ g ; n, n) = m,n z k (Σ g ; m, n) ẑ k (Σ 0 ; m, n), where Σ 0 is of course just the 2-sphere. Thus, provided we understand ẑ k (Σ 0 ; m, n), we can reduce the computation of ẑ k to the computation of z k, which is well understood and which we shall now describe. Following Verlinde [2], we define the coefficients N m,m 2,m 3 = z k (Σ 0 ; m, m 2, m 3 ), which are symmetric in all three indices. Define the (k + ) (k + ) symmetric matrices N m by (N m ) i,j = N m,i,j. Both here and in future we index the rows and columns starting with 0 instead of, for convenience. We then obtain the following formula for the torus with two marked points. (6) Proposition. z k (Σ ; m, m 2 ) = tr N m N m2. z Proof. k (Σ ; m, m 2 ) = z k (Σ 0 ; m, m, m, m 2 ) m = z k (Σ 0 ; m, n, m ) z k (Σ 0 ; m, n, m 2 ) m,n = m,n(n m ) m,n (N m2 ) n,m = tr N m N m2. Now define another (k + ) (k + ) symmetric matrix M by (M) i,j = tr (N i N j ). This enables us to generalize the previous formula to arbitrary genus. (7) Proposition. For g, z k (Σ g ; m, m 2 ) = (M g ) m,m 2. Proof by induction. Proposition (6) is the case g =. Then, if z k (Σ g ; m, m 2 ) = (M g ) m,m 2, by attaching a handle we obtain z k (Σ g ; m, m 2 ) = n z k (Σ g ; m, n) z k (Σ ; n, m 2 ) = (M g ) m,n(m) n,m2 n = (M g ) m,m 2. (8) Corollary. z k (Σ g ) = tr M g. Proof. z k (Σ g ) = z k (Σ g ; m, m) = (M g ) m,m = tr M g. It is worth mentioning that the coefficients N m,m 2,m 3 can actually be calculated explicitly: 5

(9) Proposition. { if m m N m,m 2,m 3 = 2 m 3 min(m + m 2, k m m 2 ) and m + m 2 + m 3 is even 0 otherwise. The proof closest to our point of view appears in [7]; a fuller discussion can be found in [8]. However, we shall not use the coefficients in this form, but rather make use of the Verlinde conjecture, which states that the matrices N m can be simultaneously diagonalized. Let S be the (k + ) (k + ) matrix such that ( ) 2 (i + )(j + )π (S) ij = k + 2 sin. k + 2 Note that S is orthogonal and symmetric, so S = S. (0) Proposition. SN m S = diag ( ) (S)n,m (S) 0,m. Proofs are given in [2], [8], and [22]. () Corollary. Proof. (2) Corollary. SMS = diag ( (S) 2 0,m (SMS) i,j = m,n(s) i,m tr (N m N n )(S) n,j z k (Σ g ) = = = p m,n,p,q ( k + 2 2 ) (S) i,m N m,p,q N n,p,q (S) n,j (SN 2 p S) i,j. ) g k+ m=. (sin mπ k+2 )2g 2. Proof. (8) and (). In order to find a similar formula for ẑ k, as we said above, we need to understand ẑ k (Σ 0 ; m, m 2 ). This can be computed directly. (3) Proposition. ẑ k (Σ 0 ; m, m 2 ) = δ m +m 2 k. Proof. In this case the action of Ĝ on  is well understood []. The orbits are indexed by positive integers n;  n consists of those holomorphic structures which are isomorphic to O( n) O(n + ). The complex codimension of  n in  is 2n. In particular  0 is a dense open orbit. Hence any equivariant section over  0 will extend over  by the Hartogs theorem. That is, Ẑ k (Σ 0 ; m, m 2 ) = ΓĜ( ˆL k Â0 V m V m2 ). 6

However, since  0 is a single orbit of Ĝ, such an equivariant section is determined by its value at a single fixed a  0. Conversely, any element of ˆLk a V m V m2 which is fixed by Ĝa, the stabilizer of a, can be moved around by Ĝ to give an equivariant section. Hence (4) Ẑ k (Σ 0, m, m 2 ) = ΓĜa ( ˆL k a V m V m2 ). Now Ĝa consists precisely of the ad P -compatible automorphisms of Û which are holomorphic with respect to a. Since a  0 this means that Ĝa is isomorphic to the group of holomorphic automorphisms of O O() having determinant on each fibre. If x, x 2 are our two marked points, we may choose a basis {s, s 2 } of H 0 (Σ 0 ; O()) such that s vanishes at x and s 2 vanishes at x 2. Then we can write {( ) } z 0 (5) Ĝ a = y s + y 2 s 2 z : z C, y i C, where the rows and columns of the matrix correspond to the decomposition O O(). Now since H (Σ 0 ; O O()) = 0, the fibre of the determinant line bundle L at a is just L 2 a = Λ 3 H 0 ((Σ 0 ; O O()). A straightforward computation shows that, with respect to the presentation in (5), an element of Ĝ a acts on L a by multiplication by z. On the other hand, for i =, 2, the vanishing property of s i at x i implies that, under an appropriate identification of (Ãd ˆP ) xi with SL(2, C), Ĝ a acts on V ni via the homomorphism Ĝa SL(2, C) given by ( ) ( ) z 0 z 0 y s + y 2 s 2 z y i z. Let us try to determine the right-hand side of (4) explicitly. Fix a nonzero t ˆL k a ; then any element of ˆLk a V m V m2 is of course of the form t v for v V m V m2. The -parameter subgroups {( z 0 0 z ) : z C } and {( 0 y i s i ) : y i C } for i =, 2 generate Ĝa, so for a section φ to be Ĝa -equivariant it is necessary ( and sufficient ) to z 0 be equivariant with respect to each of these. Now we saw that 0 z t = z k t, ( ) z 0 so for equivariance we must have 0 z v = z k v for that v such that φ(a) = t v. In infinitesimal terms, this means that v (V m V m2 )( k), the k-weight space ( of V m ) V m2 as a representation of the Lie algebra sl(2, C). A typical element 0 of the second subgroup acts trivially on y s ˆL k a and on V m2 ; on V m it acts ( ) 0 as SL(2, C). Hence for equivariance v must belong to L y m V m2, where ( ) 0 L m V m is the subspace on which acts trivially. Infinitesimalizing again y 7

we recognize L m as the lowest weight space V m ( m ). In a similar way we deduce v V m V m2 ( m 2 ). Now since v determines φ, we can identify ΓĜa ( ˆL k a V m V m2 ) = (V m V m2 )( k) V m ( m ) V m2 V m V m2 ( m 2 ) = (V m V m2 )( k) V m ( m ) V m2 ( m 2 ) = (V m V m2 )( k) (V m V m2 )( m m 2 ), which clearly has dimension δ m +m 2 k, q.e.d. (6) Corollary. ẑ k (Σ g ; m, m 2 ) = (M g J) m,m 2, where J is the (k + ) (k + ) matrix with (J) i,j = δ i+j k. k Proof. ẑ k (Σ g ; m, m 2 ) = z k (Σ g ; m, m) ẑ k (Σ 0 ; m, m 2 ); then use (7) and (3). m=0 (In particular, this corollary proves (5), because J 2 = I.) (7) Corollary. For g, ẑ k (Σ g ) = tr M g, where we define tr A = k (A) i,k i. i=0 Of course, this twisted trace is not preserved by the diagonalization, so to actually evaluate ẑ k we notice instead that J is also diagonalized by S ; in fact, SJS = diag ( ) n. This gives us an analogue of (2). (8) Corollary. ẑ k (Σ g ) = ( k + 2 2 ) g k+ m= ( ) m+ (sin mπ k+2 )2g 2. This formula is the only result obtained so far which will be needed in the sequel. Hence to simplify notation we will in future denote ẑ k (Σ g ) simply by D(g, k). Although it is far from apparent from (8), for fixed g D(g, k) is a polynomial in k with rational coefficients. We shall conclude this section by proving that. We shall closely imitate similar calculations for the untwisted case which appear in an unpublished letter by Don Zagier. Similar results have also been obtained by Dowker [5]. (9) Proposition. For fixed g 2 and even k, D(g, k) equals the coefficient of x 3g 3 in the power series expansion of ( k + 2 ) g ( x 2 x sinh x ) 2g 2 (k + 2)x sinh(k + 2)x. Proof. We first rewrite (8) in terms of roots of unity: D(g, k) = ( 2k 4)g 2 8 ζ 2k+4 = ζ ± ζ k+2 (ζ ζ ) 2g 2.

Substituting ζ 2 = λ, we get D(g, k) = ( 2k 4) g λ k+2 = λ λ g λ (k+2)/2 (λ ) 2g 2. Hence D(g, k) ( 2k 4) g = λ k+2 = λ = Res z= [ Res z=λ [ z g z g (k + 2)z (k+2)/2 (z ) 2g 2 z k+2 (k + 2)z (k+2)/2 (z ) 2g 2 z k+2 ] dz z ] dz z for g 2 by the residue theorem, since then the only poles of the expression in square brackets are at the (k+2)nd roots of unity. Substituting z = e 2x in the final expression gives [ ( ) D(g, k) 2g 2 k + 2 = Res ( 2k 4) g x=0 2 sinh x sinh(k + 2)x dx = coefficient of x 2g 2 in 2 2g 2 ( x sinh x ] ) 2g 2 (k + 2)x sinh(k + 2)x. 4 The cohomology of N g Let Σ g be a Riemann surface of genus g 2. Fix a holomorphic line bundle Λ over Σ g of degree. Then the moduli space N g of stable rank 2 holomorphic vector bundles V over Σ g with Λ 2 V = Λ is a compact complex 3g 3-manifold. We wish to study the structure of the rational cohomology ring H (N g ). In 5, this will essentially be determined completely using proposition (9). First, though, we need to review what is already known about N g, and to prove some necessary lemmas. A broader discussion of some of the topics mentioned can be found in []. Roughly speaking, we wish to identify N g as a space of representations of π (Σ g ). So let us choose loops e, e 2,..., e 2g on Σ g which generate π (Σ g ) in the usual way, so that e i e i+g e i e i+g. If we cut out a small disc D Σ g, then e i e i+g e i e i+g is homotopic to the boundary circle of D, but is no longer contractible. Hence we lose the relation, and π (Σ g D) is the free group on the generators e i. Now consider the map µ : SU(2) 2g SU(2) defined by (A, A 2,..., A 2g ) g i= A i A i+g A i A i+g and in particular the subspace S g = µ ( I) SU(2) 2g. It can be shown [] that I is a regular value of µ, so S g is a smooth 6g 3-submanifold of SU(2) 2g. We may regard an element ω S g as a representation of π (Σ g D) sending the boundary 9

circle to I. It must be irreducible, since if it were reducible to an abelian subgroup, it would send the boundary circle to I. Such a representation gives us a flat connection on an SU(2) bundle over Σ g D having holonomy I around D. By passing to the associated rank 2 vector bundle and gluing in a fixed twisted unitary connection over D, we can extend our flat connection ω to a connection on a unitary vector bundle V Σ g of degree. Then the (0, ) part of the associated covariant derivative is a Cauchy-Riemann operator which induces a holomorphic structure on V. Now the diagonal conjugation action of SU(2)/ ± = SO(3) on SU(2) 2g clearly preserves S g, and by Schur s lemma the restriction of the action is free. Hence the quotient S g /SO(3) is a smooth 6g 6-manifold. But if two representations in S g are conjugate, then the induced connections on V are isomorphic, and hence so are the holomorphic structures. Hence to each point in S g /SO(3) we can associate an isomorphism class of holomorphic bundles over Σ g. In this context, the celebrated theorem of Narasimhan and Seshadri [8] can be stated as follows. (20) Theorem. All the holomorphic bundles constructed in this manner are stable, and the resulting map φ : S g /SO(3) N g is a diffeomorphism. Remark. We will in future identify N g with S g /SO(3). Under this identification, any diffeomorphism f : Σ g Σ g induces a diffeomorphism ˆf : N g N g. (To be precise, we should require that f preserves a small neighbourhood of the disc D; but this is not a serious restriction, as any diffeomorphism of Σ g is isotopic to one of this form.) To obtain distinguished cohomology classes in N g, we construct a vector bundle U over N g Σ g, as follows. Let Σ g be the universal cover of Σ g, and consider the twisted quotient Σ g π (Σ g) (S g sl(2, C)). The action on the right-hand factor is given by h(ρ, v) = (ρ, ad ρ(h) v), where we regard ρ S g as determining an SO(3)-representation of π (Σ g ). This gives us a vector bundle over S g Σ g ; to see that it descends to N g Σ g, note that the conjugation action of SO(3) on S g lifts to an action on this twisted quotient, given by T (s (ρ, v)) = (s (T ρt, ad T v)). The resulting vector bundle U is clearly natural in the sense that, for any diffeomorphism f : Σ g Σ g, we have ( ˆf f) (U ) = U. (In fact, it is also universal, that is, it has a connection in the fibre direction whose restriction to {ω} Σ g is isomorphic to ad ω. It is even possible to perform a similar construction in the holomorphic setting for details see [].) Because N g is simply connected [3], we may write c 2 (U ) = 2α σ + β 4ψ, where σ denotes the fundamental class in H 2 (Σ g ) and α H 2 (N g ); β H 4 (N g ); ψ H 3 (N g ) H (Σ g ). 0

(The scalar factors are inserted to agree with the conventions of [4].) The Poincaré duals of our loops e i form a basis e, e 2,..., e 2g of H (Σ g, Z); using this, we can decompose ψ = 2g i= ψ i e i, where ψ i H 3 (N g ). We can now state the following crucial result. (2) Theorem (Newstead). The ring H (N g ) is generated by α, β, and the ψ i. Remark. Atiyah and Bott [] later showed that the integral cohomology H (N g, Z) is torsion-free, so that rational multiples of the same classes generate the integral cohomology. Moreover, the ψ i are actually integral generators of H 3 (N g, Z), a fact which we will be needing later. Since we have a set of generators, to determine the ring structure completely it suffices to find the relations. The usual commutation relations hold, so this amounts to deciding for which nonnegative integers m j, n j, p i,j we have (22) α m j β n j ( j i ψ p i,j i ) = 0. But according to Poincaré duality, µ H m (N g ) = 0 if and only if µν = 0 for all ν H 6g 6 m (N g ). Hence in principle to decide when equality holds in (22) we need only evaluate the integers α m β n ( ψ p i (23) i )[N g ] i for those m, n, p i such that 2m + 4n + 3 i p i = 6g 6. A more general discussion of these top-dimensional pairings, and of their analogues in four dimensions, can be found in [4]. Our goal will be to find an explicit formula for (23). The first thing to notice is the following. (24) Proposition. Let i 0 g. If p i0 >, or if p i0 p i0 +g, then the pairing in (23) is zero. Proof. The first part is easy, since the commutation relations imply ψi 2 0 = 0. This only leaves the case p i0 + p i0 +g = for the second part. Now if f : Σ g Σ g is an orientation-preserving diffeomorphism, then by naturality ˆf (α) m ˆf (β ) n ( i ˆf (ψ i ) p i )[N g ] = α m β n ( i ψ p i i )[N g ]. However, since ( ˆf f) U = U, we have ˆf (α) = α, ˆf (β ) = β, i ˆf (ψ i )f (e i ) = i ψ i e i.

Now it is easy to find a diffeomorphism of Σ g such that f (e i 0 ) = e i 0, f (e i0+g ) = e i0+g, but f (e i ) = e i for all other i. For example, a half twist of the surface in Figure below the loop labelled has the desired properties. If p i0 + p i0 +g =, we conclude α m β n ( ψ p i i )[N g ] = α m β n ( ψ p i i )[N g ] = 0. i i For i g, define γ i = ψ i ψ i+g. Then the proposition above shows that all the pairings (23) are zero except those of the form (25) (α m β n γ i γ i2... γ ip )[N g ] where m + 2n + 3p = 3g 3 and i < i 2 < < i p g. Actually, the value of (25) is independent of the choice of i j. This follows from a diffeomorphism argument similar to that in the proof of (24), because any permutation of the handles in figure can be realized by a diffeomorphism. Consequently, any pairing of the form (25) is equal to (g p)!/(2 p g!) times the even simpler expression (α m β n γ p )[N g ], where γ = 2 g i= γ i. The class γ has the advantage of being independent of our choice of basis {e i }, since ψ 2 = γ σ. Still using the identification N g = S g /SO(3), we can perform another construction that would not have made sense in the holomorphic setting. The map from Σ g to Σ g which collapses the ith handle to a point (see figure 2) is not holomorphic, but it induces an embedding η i : N g N g. The image of this embedding consists precisely of those conjugacy classes of representations of π (ˆΣ g ) which send the generators e i and e i+g to the identity. That is, η i (N g ) = p(πi (I) πi+g(i)), where π i is the restriction to S g of the projection of SU(2) 2g on the ith factor, and p is the quotient map S g N g. Since dim R N g = 6g 6, N g has real codimension 6. (26) Proposition. η i (N g ) is Poincaré dual to γ i. Consequently, α m β n γ p [N g ] = 2g α m β n γ p [N g ]. Proof. If we fix an element (A, A 2,..., A 2g 2 ) S g, then the map SU(2) S g given by T (A,..., A i, T, A i,..., A i+g, T, A i+g,..., A 2g 2 ) is a right inverse for π i. There is a similar right inverse for π i+g. Hence the pullbacks by π i and π i+g of the fundamental cohomology class of SU(2) are indivisible classes χ i, χ i+g H 3 (S g, Z), Poincaré dual to πi (I) and πi+g(i) respectively. We claim that χ i = ±p (ψ i ) and χ i+g = ±p (ψ i+g ). To prove this, note that πi (I) is a union of fibres of the SU(2)-action on S g, so χ i = p ( ˆψ i ) for some ˆψ i H 3 (N g, Z). This ˆψ i is unique, because we can see from the Leray-Serre spectral sequence (or even the Gysin sequence) that the natural map H 3 (N g, Z) H 3 (S g, Z) is injective. Now for any e j in our basis for H (Σ g, Z) with 2

j i, there exist diffeomorphisms f : Σ g Σ g such that f (e i ) = e i but f (e j ) e j : the Dehn twists and half twists in figure 3 do the trick. Thus the only classes in H 3 (N g, Z) which are invariant under f for all such f are the integer multiples of ψ i. But χ i is invariant under such f by construction, and hence so is ˆψ i. Hence by the second part of (2), ˆψi is an integer multiple of ψ i. Since both ψ i and ˆψ i are indivisible, we must have ˆψ i = ±ψ i. Likewise, ˆψi+g = ±ψ i+g. We shall not bother to pin down the sign, because for our purposes it is sufficient to note that the same sign holds for ˆψ i and ˆψ i+g. This can be deduced from a diffeomorphism argument like the one above, using a Dehn twist that induces e i e i+g and e i+g e i (see figure 4). Then χ i χ i+g = p (γ i ), and so descending to N g we obtain the first result. As for the second, it is sufficient to show that η (α) = α, η (β ) = β, and η (γ ) = γ. This is straightforward from the construction of U. We can use this proposition recursively to eliminate γ from the pairings. (Actually, to bootstrap we do need to know that γ [N 2 ] = 4, but this follows from Poincaré duality.) Hence it now suffices to evaluate α m β n [N g ] when m + 2n = 3g 3. This will be carried out in the next section. 5 A formula for the pairings So far, we have practically ignored the complex structure on Σ g, and the complex structure it induces on N g. In this section we shall use that complex structure, together with what we already know, to calculate (α m β n γ p )[N g ]. Recall that in 3, we gave a formula for D(g, k) = dim H 0 (N g ; L k/2 ), where L was a positive line bundle over N g, and k was an even number. However, instead of the approach we took through mathematical physics, we could have gone through algebraic geometry, using the Hirzebruch-Riemann-Roch theorem. This tells us that (27) ( ) i dim H i (N g ; L k/2 ) = (ch L k/2 td N g )[N g ]. i Since the canonical bundle of N g is negative [4], the left-hand side equals D(g, k) by Kodaira vanishing. On the other hand, we can calculate the right-hand side in terms of α and β. Newstead [4] showed that c (N g ) = 2α and p(n g ) = ( + β ) 2g 2. By restricting L to a projective subspace of N g as in [4], it is not hard to show that c (L) = α. As for td N g, the identity x e = exp( x) x 2 x 2 sinh x 2 implies (see p. 7 of [6]) that, if y i are the Pontrjagin roots, td = exp( 2 c ) i 2 yi. sinh 2 yi Noting that 2g 2 of the Pontrjagin roots of N g are β and the rest are 0, we obtain ( ) 2g 2 td N g = exp(α) 2 β. sinh 2 β 3

Then by (27) (28) ( ) ( ) k + 2 2g 2 D(g, k) = exp α 2 β [N g ] 2 sinh 2 β for even k. For fixed g, this is a polynomial in k whose coefficients involve the pairings. On the other hand, our original formula (9) is likewise a polynomial in k. Substituting l = k + 2 in (9) and (28), and equating coefficients of l m, we get ( 2 )g (2 m g+ 2) (m g+)! B m g+ P 3g 3 m = m! 2 3g 3 P 3g 3 m (α m β 2 (3g 3 m) )[N g ], where B i is the ith Bernoulli number and P i is the coefficient of x i in the power series expansion of (x/ sinh x) 2g 2. (We use the conventions under which B 2 = /6, B 4 = /30, etc., and we interpret B i = 0 if i < 0, because (k + 2)x/ sinh(k + 2)x has no pole at 0.) Cancelling P 3g 3 m and rearranging, we obtain (29) (α m β n g m! )[N g ] = ( ) (m g+)! 22g 2 (2 m g+ 2) B m g+. In particular, we get the volume formula: vol N g = (3g 3)! α 3g 3 [N g ] = (2g 2)! 22g 2 (2 2g 2 2) B 2g 2. Finally, we combine (29) with (26) to obtain, for m + 2n + 3p = 3g 3, (α m β n γ p p g m! )[N g ] = ( ) (m+p g+)! g! (g p)! 22g 2 p (2 m+p g+ 2) B m+p g+. This agrees with the results published for genus 2 [3] and 3 [5]. For example, the number 224 obtained by Ramanan is just α 6 [N 3 ]. It also enables us to prove Newstead s conjecture that β g = 0. This conjecture first appeared in [4] and is discussed at greater length in []; there is no proof in the literature, but Frances Kirwan [2] has recently found a proof using equivariant cohomology. In any case, though, it is manifest from our formula, since n > g implies m+p g+ < 0. Indeed, we get the more general and unsuspected result that β g q γ q = 0 whenever 0 q g. However, Newstead s other conjecture, proved by Gieseker [9], that c r (N g ) = 0 for r > 2g 2, is not at all obvious from this equation, since we don t know of a systematic way to compute the Chern classes in terms of α, β, γ. It was promised in the introduction that the determination of the ring structure would be reduced to a problem in linear algebra, and this ought to be explained. To compute the relations in any degree q, we construct a matrix whose columns are indexed by the monomials in α, β, ψ i of total degree q and whose rows are indexed by similar monomials of complementary total degree 6g 6 q. The matrix entry at row µ, column ν is given by µν[n g ], as determined from the formula above and (24). The rank of this matrix is just the qth Betti number of N g. (To prove that this rank agrees with the formula for the Betti number given in [] could be regarded as a hard exercise in number theory.) Finding the relations at degree q is equivalent to finding a basis for the null space of this matrix. Such a basis could be obtained by 4

row-reduction, but without any short cuts this would be prohibitively cumbersome. Luckily, though, for many applications, such as those employing the Riemann-Roch theorem, the top-dimensional pairings are exactly what we need. Note added in proof. Since this paper was written, Don Zagier has solved the problem posed in the last paragraph, to give a complete set of explicit relations between Newstead s generators. References [] M.F. Atiyah and R. Bott, The Yang-Mills equations over Riemann surfaces, Philos. Trans. Roy. Soc. London Ser. A 308 (982) 523 65. [2] R. Dijkgraaf and E. Verlinde, Modular invariance and the fusion algebra, Nucl. Phys. B (Proc. Suppl.) 5B (988) 87 97. [3] R. Dijkgraaf and E. Witten, Topological gauge theories and group cohomology, Commun. Math. Phys. 29 (990) 393 429. [4] S.K. Donaldson, Instantons in Yang-Mills theory, in The interface of mathematics and particle physics, ed. D.G. Quillen, G.B. Segal, and Tsou S.T. (Oxford, 990). [5] J.S. Dowker, Heat kernel expansion on a generalized cone, J. Math. Phys. 30 (989) 770 773. [6] G. Felder, K. Gawȩdzki, and A. Kupiainen, The spectrum of Wess-Zumino-Witten models, Nucl. Phys. B 299 (988) 355 366. [7] K. Gawȩdzki and A. Kupiainen, SU(2) Chern-Simons theory at genus zero, IHES preprint IHES/P/90/8. [8] D. Gepner and E. Witten, String theory on group manifolds, Nucl. Phys. B 278 (986) 493 549. [9] D. Gieseker, A degeneration of the moduli space of stable bundles, J. Diffl. Geom. 9 (984) 73 206. [0] P. Griffiths and J. Harris, Principles of algebraic geometry (Wiley, 978). [] J. Igusa, On a property of commutators in the unitary group, Mem. Coll. Sci. Kyoto Univ., Ser. A 26 (950) 45 49. [2] F.C. Kirwan, The cohomology rings of moduli spaces of bundles over Riemann surfaces, to appear. [3] P.E. Newstead, Topological properties of some spaces of stable bundles, Topology 6 (967) 24 262. [4] P.E. Newstead, Characteristic classes of stable bundles over an algebraic curve, Trans. Am. Math. Soc. 69 (972) 337 345. 5

[5] S. Ramanan, The moduli spaces of vector bundles over an algebraic curve, Math. Annalen 200 (973) 69 84. [6] J. Roe, Elliptic operators, topology and asymptotic methods (Longman, 988). [7] G.B. Segal, The definition of conformal field theory, Oxford preprint. [8] C.S. Seshadri, Fibrés vectoriels sur les courbes algébriques, Astérisque 96 (982). [9] C.S. Seshadri, Space of unitary vector bundles on a compact Riemann surface, Ann. Math. 85 (967) 303 336. [20] A. Tsuchiya, K. Ueno, and Y. Yamada, Conformal field theory on universal family of stable curves with gauge symmetries, Max-Planck-Institut preprint MPI/89/8. [2] E. Verlinde, Fusion rules and modular transformations in 2d conformal field theory, Nucl. Phys. B 300 (988) 360 376. [22] E. Witten, Quantum field theory and the Jones polynomial, Commun. Math. Phys. 2 (989) 35 399. 6