Quiver varieties and finite dimensional representations of quantum affine algebras

By Hiraku Nakajima

Abstract

We study finite dimensional representations of the quantum affine algebra using geometry of quiver varieties introduced by the author. As an application, we obtain character formulas expressed in terms of intersection cohomologies of quiver varieties.

Introduction

Let be a simple finite dimensional Lie algebra of type , let be the corresponding (untwisted) affine Lie algebra, and let be its quantum enveloping algebra of Drinfel’d-Jimbo, or the quantum affine algebra for short. In this paper we study finite dimensional representations of , using geometry of quiver varieties which were introduced in Reference 29Reference 44Reference 45.

There is a large amount of literature on finite dimensional representations of ; see for example Reference 1Reference 10Reference 18Reference 25Reference 28 and the references therein. A basic result relevant to us is due to Chari-Pressley Reference 11: irreducible finite dimensional representations are classified by an -tuple of polynomials, where is the rank of . This result was announced for Yangian earlier by Drinfel’d Reference 15. Hence the polynomials are called Drinfel’d polynomials. However, not much is known about the properties of irreducible finite dimensional representations, say their dimensions, tensor product decomposition, etc.

Quiver varieties are generalizations of moduli spaces of instantons (anti-self-dual connections) on certain classes of real -dimensional hyper-Kähler manifolds, called ALE spaces Reference 29. They can be defined for any finite graph, but we are concerned for the moment with the Dynkin graph of type corresponding to . Motivated by results of Ringel Reference 47 and Lusztig Reference 33, the author has been studying their properties Reference 44Reference 45. In particular, it was shown that there is a homomorphism

where is the universal enveloping algebra of , is a certain lagrangian subvariety of the product of quiver varieties (the quiver variety depends on a choice of a dominant weight ), and denotes the top degree homology group with complex coefficients. The multiplication on the right hand side is defined by the convolution product.

During the study, it became clear that the quiver varieties are analogous to the cotangent bundle of the flag variety . The lagrangian subvariety is an analogue of the Steinberg variety , where is the nilpotent cone and is the Springer resolution. The above mentioned result is an analogue of Ginzburg’s lagrangian construction of the Weyl group Reference 20. If we replace homology group by equivariant -homology group in the case of , we get the affine Hecke algebra instead of as was shown by Kazhdan-Lusztig Reference 26 and Ginzburg Reference 13. Thus it became natural to conjecture that an equivariant -homology group of the quiver variety gave us the quantum affine algebra . After the author wrote Reference 44, many people suggested this conjecture to him, for example Kashiwara, Ginzburg, Lusztig and Vasserot.

A geometric approach to finite dimensional representations of (when ) was given by Ginzburg-Vasserot Reference 21Reference 58. They used the cotangent bundle of the -step partial flag variety, which is an example of a quiver variety of type . Thus their result can be considered as a partial solution to the conjecture.

In Reference 23 Grojnowski constructed the lower-half part of on equivariant -homology of a certain lagrangian subvariety of the cotangent bundle of a variety . This was used earlier by Lusztig for the construction of canonical bases on the lower-half part of the quantized enveloping algebra . Grojnowski’s construction was motivated in part by Tanisaki’s result Reference 52: a homomorphism from the finite Hecke algebra to the equivariant -homology of the Steinberg variety is defined by assigning to perverse sheaves (or more precisely Hodge modules) on their characteristic cycles. In the same way, he considered characteristic cycles of perverse sheaves on . Thus he obtained a homomorphism from to -homology of the lagrangian subvariety. This lagrangian subvariety contains a lagrangian subvariety of the quiver variety as an open subvariety. Thus his construction was a solution to ‘half’ of the conjecture.

Later Grojnowski wrote an ‘advertisement’ of his book on the full conjecture Reference 24. Unfortunately, details were not explained, and his book is not published yet.

The purpose of this paper is to solve the conjecture affirmatively, and to derive results whose analogues are known for . Recall that Kazhdan-Lusztig Reference 26 gave a classification of simple modules of , using the above mentioned -theoretic construction. Our analogue is the Drinfel’d-Chari-Pressley classification. Also Ginzburg gave a character formula, called a -adic analogue of the Kazhdan-Lusztig multiplicity formula Reference 13. (See the introduction in Reference 13 for a more detailed account and historical comments.) We prove a similar formula for in this paper.

Let us describe the contents of this paper in more detail. In §1 we recall a new realization of , called Drinfel’d realization Reference 15. It is more suitable than the original one for our purpose, or rather, we can consider it as a definition of . We also introduce the quantum loop algebra , which is a subquotient of , i.e., the quantum affine algebra without central extension and the degree operator. Since the central extension acts trivially on finite dimensional representations, we study rather than . Introducing a certain -subalgebra of , we define a specialization of at . This was originally introduced by Chari-Pressley Reference 12 for the study of finite dimensional representations of when is a root of unity. Then we recall basic results on finite dimensional representations of . We introduce several concepts, such as l-weights, l-dominant, l-highest weight modules, l-fundamental representation, etc. These are analogues of the same concepts without l for -modules.l’ stands for the loop. In the literature, some of these concepts were used without ‘l’.

In §2 we introduce two types of quiver varieties, , (both depend on a choice of a dominant weight ). They are analogues of and the nilpotent cone respectively, and have the following properties:

(1)

is a nonsingular quasi-projective variety, having many components of various dimensions.

(2)

is an affine algebraic variety, not necessarily irreducible.

(3)

Both and have a -action, where .

(4)

There is a -equivariant projective morphism .

In §3–§8 we prepare some results on quiver varieties and -theory which we use in later sections.

In §9–§11 we consider an analogue of the Steinberg variety and its equivariant -homology . We construct an algebra homomorphism

We first define images of generators in §9, and check the defining relations in §10 and §11. Unlike the case of the affine Hecke algebra, where is isomorphic to ( the Steinberg variety), this homomorphism is not an isomorphism, neither injective nor surjective.

In §12 we show that the above homomorphism induces a homomorphism

(It is natural to expect that is an integral form of and that is torsion-free, but we do not have the proofs.)

In §13 we introduce a standard module . It depends on the choice of a point and a semisimple element such that is fixed by . The parameter corresponds to the specialization , while corresponds to Drinfel’d polynomials. In this paper, we assume is not a root of unity, although most of our results hold even in that case (see Remark 14.3.9). Let be the Zariski closure of . We define as the specialized equivariant -homology , where is a fiber of at , and is an -algebra structure on determined by . By the convolution product, has a -module structure. Thus it has a -module structure by the above homomorphism. By the localization theorem of equivariant -homology due to Thomason Reference 55, is isomorphic to the complexified (non-equivariant) -homology of the fixed point set . Moreover, it is isomorphic to via the Chern character homomorphism thanks to a result in §7. We also show that is a finite dimensional l-highest weight module. As a usual argument for Verma modules, has the unique (nonzero) simple quotient. The author conjectures that is a tensor product of l-fundamental representations in some order. This is proved when the parameter is generic in §14.1.

In §14 we show that the standard modules and are isomorphic if and only if and are contained in the same stratum. Here the fixed point set has a stratification defined in §4. Furthermore, we show that the index set of the stratum coincides with the set of l-dominant l-weights of , the standard module corresponding to the central fiber . Let us denote by the index corresponding to . Thus we may denote and its unique simple quotient by and respectively if is contained in the stratum corresponding to an l-dominant l-weight . We prove the multiplicity formula

where is a point in , is the inclusion, and is the intersection cohomology complex attached to and the constant local system .

Our result is simpler than the case of the affine Hecke algebra: nonconstant local systems never appear. This phenomenon corresponds to an algebraic result that all modules are l-highest weight. It compensates for the difference of and during the proof of the multiplicity formula.

If is of type , then coincides with a product of varieties studied by Lusztig Reference 33, where the underlying graph is of type . In particular, the Poincaré polynomial of is a Kazhdan-Lusztig polynomial for a Weyl group of type . We should have a combinatorial algorithm to compute Poincaré polynomials of for general .

Once we know , information about can be deduced from information about , which is easier to study. For example, consider the following problems:

(1)

Compute Frenkel-Reshetikhin’s -characters Reference 18.

(2)

Decompose restrictions of finite dimensional -modules to -modules (see Reference 28).

These problems for are easier than those for , and we have the following answers.

Frenkel-Reshetikhin’s -characters are generating functions of dimensions of l-weight spaces (see §13.5). In §13.5 we show that these dimensions are Euler numbers of connected components of for standard modules . As an application, we prove a conjecture in Reference 18 for of type (Proposition 13.5.2). These Euler numbers should be computable.

Let be the restriction of to a -module. In §15 we show the multiplicity formula

where is a weight such that is dominant, is the corresponding irreducible finite dimensional module (these are concepts for usual without ‘l’), is a point in , is the inclusion, is a stratum of , and is the intersection cohomology complex attached to and the constant local system .

If is of type , then the stratum coincides with a nilpotent orbit cut out by Slodowy’s transversal slice Reference 44, 8.4. The Poincaré polynomials of were calculated by Lusztig Reference 30 and coincide with Kostka polynomials. This result is compatible with the conjecture that is a tensor product of l-fundamental representations, for the restriction of an l-fundamental representation is simple for type , and Kostka polynomials give tensor product decompositions. We should have a combinatorial algorithm to compute Poincaré polynomials of for general .

We give two examples where can be described explicitly.

Consider the case that is a fundamental weight of type , or more generally a fundamental weight such that the label of the corresponding vertex of the Dynkin diagram is . Then it is easy to see that the corresponding quiver variety consists of a single point . Thus remains irreducible in this case.

If is the highest weight of the adjoint representation, the corresponding is a simple singularity , where is a finite subgroup of of the type corresponding to . Then has two strata and . The intersection cohomology complexes are constant sheaves. Hence we have

These two results were shown by Chari-Pressley Reference 9 by a totally different method.

As we mentioned, the quantum affine algebra has another realization, called the Drinfel’d new realization. This Drinfel’d construction can be applied to any symmetrizable Kac-Moody algebra , not necessarily a finite dimensional one. This generalization also fits our result, since quiver varieties can be defined for arbitrary finite graphs. If we replace finite dimensional representations by l-integrable representations, parts of our result can be generalized to a Kac-Moody algebra , at least when it is symmetric. For example, we generalize the Drinfel’d-Chari-Pressley parametrization. A generalization of the multiplicity formula requires further study.

If is an affine Lie algebra, then is the quantum affinization of the affine Lie algebra. It is called a double loop algebra, or toroidal algebra, and has been studied by various people; see for example Reference 22Reference 48Reference 49Reference 56 and the references therein. A first step to a geometric approach to the toroidal algebra using quiver varieties for the affine Dynkin graph of type was given by M. Varagnolo and E. Vasserot Reference 57. In fact, quiver varieties for affine Dynkin graphs are moduli spaces of instantons (or torsion free sheave) on ALE spaces. Thus these cases are relevant to the original motivation, i.e., a study of the relation between -dimensional gauge theory and representation theory. In some cases, these quiver varieties coincide with Hilbert schemes of points on ALE spaces, for which many results have been obtained (see Reference 46). We will return to this in the future.

If we replace equivariant -homology by equivariant homology, we should get the Yangian instead of . This conjecture is motivated again by the analogy of quiver varieties with . The equivariant homology of gives the graded Hecke algebra Reference 32, which is an analogue of for . As an application, the affirmative solution of the conjecture implies that the representation theory of and that of the Yangian are the same. This has been believed by many people, but there is no written proof.

While the author was preparing this paper, he was informed that Frenkel-Mukhin Reference 17 proved the conjecture in Reference 18 (Proposition 13.5.2) for general .

Acknowledgement.

Part of this work was done while the author enjoyed the hospitality of the Institute for Advanced Study. The author is grateful to G. Lusztig for his interest and encouragement.

1. Quantum affine algebra

In this section, we give a quick review for the definitions of the quantized universal enveloping algebra of the Kac-Moody algebra associated with a symmetrizable generalized Cartan matrix, its affinization , and the associated loop algebra . Although the algebras defined via quiver varieties are automatically symmetric, we treat the nonsymmetric case also for completeness.

1.1. Quantized universal enveloping algebra

Let be an indeterminate. For nonnegative integers , define

Suppose that the following data are given:

(1)

: free -module (weight lattice),

(2)

with a natural pairing ,

(3)

an index set of simple roots

(4)

() (simple root),

(5)

() (simple coroot),

(6)

a symmetric bilinear form on .

These are required to satisfy the following:

(a)

for and ,

(b)

is a symmetrizable generalized Cartan matrix, i.e., , and and for ,

(c)

,

(d)

are linearly independent,

(e)

there exists () such that (fundamental weight).

The quantized universal enveloping algebra of the Kac-Moody algebra is the -algebra generated by , (), () with relations

where , .

Let (resp. ) be the -subalgebra of generated by the elements (resp. ). Let be the -subalgebra generated by elements (). Then we have the triangle decomposition Reference 36, 3.2.5:

Let and . Let be the -subalgebra of generated by elements , , for , , . It is known that is an integral form of , i.e., the natural map is an isomorphism. (See Reference 10, 9.3.1.) For , let us define as via the algebra homomorphism that takes to . It will be called the specialized quantized enveloping algebra. We say a -module (defined over ) is a highest weight module with highest weight if there exists a vector such that

Then there exists a direct sum decomposition (weight space decomposition) where for any }. By using the triangular decomposition Equation 1.1.6, one can show that the simple highest weight -module is determined uniquely by .

We say a -module (defined over ) is integrable if has a weight space decomposition with , and for any , there exists such that for all and .

The (unique) simple highest weight -module with highest weight is integrable if and only if is a dominant integral weight , i.e., for any (Reference 36, 3.5.6, 3.5.8). In this case, the integrable highest weight -module with highest weight is denoted by .

For a -module (defined over ), we define highest weight modules, integrable modules, etc. in a similar way.

Suppose is dominant. Let , where is the highest weight vector. It is known that the natural map is an isomorphism and is the simple integrable highest weight module of the corresponding Kac-Moody algebra with highest weight , where is the homomorphism that sends to (Reference 36, Chapter 14 and 33.1.3). Unless is a root of unity, the simple integrable highest weight -module is the specialization of (Reference 10, 10.1.14, 10.1.15).

1.2. Quantum affine algebra

The quantum affinization of (or simply quantum affine algebra) is an associative algebra over generated by (, ), (), , , and (, ) with the following defining relations:

where , , , , and is the symmetric group of letters. Here , , , are generating functions defined by

We will also need the following generating function later:

We have

Remark 1.2.12.

When is finite dimensional, then . Then the relation Equation 1.2.10 reduces to the one in literature. Our generalization seems natural since we will check it later, at least for symmetric .

Let (resp. ) be the -subalgebra of generated by the elements (resp. ). Let be the -subalgebra generated by the elements , .

The quantum loop algebra is the subalgebra of generated by (, ), (), and (, ), i.e., generators other than , . We will be concerned only with the quantum loop algebra, and not with the quantum affine algebra in the sequel.

There is a homomorphism defined by

Let and . Let be the -subalgebra generated by , , and the coefficients of for , , , . (It should be true that is free over and that the natural map is an isomorphism. But the author does not know how to prove this.) This subalgebra was introduced by Chari-Pressley Reference 12. Let (resp. ) be the -subalgebra generated by (resp. ) for , , . We have . Let be the -subalgebra generated by , the coefficients of and

for all , , , . One can easily show that (see, e.g., Reference 36, 3.1.9).

For , let be the specialized quantum loop algebra defined by via the algebra homomorphism that takes to . We assume is not a root of unity in this paper. Let and be the specializations of and respectively. We have a weak form of the triangular decomposition

which follows from the definition (cf. Reference 12, 6.1).

We say a -module is an l-highest weight module (‘l’ stands for the loop) with l-highest weight (where , ) if there exists a vector such that

By using Equation 1.2.13 and a standard argument, one can show that there is a simple l-highest weight module of with l-highest weight vector satisfying the above for any with , . Moreover, such is unique up to isomorphism. For abuse of notation, we denote the pair simply by the symbol .

A -module is said to be l-integrable if

(a)

has a weight space decomposition as a -module such that ,

(b)

for any , there exists such that for all , …, , and .

For example, if is finite dimensional, and is a finite dimensional module, then satisfies the above conditions after twisting with a certain automorphism of (Reference 10, 12.2.3).

Proposition 1.2.16.

Assume that is symmetric. The simple l-highest weight -module with l-highest weight is l-integrable if and only if is dominant and there exist polynomials for with such that

where , and denotes the expansion at and respectively.

This result was announced by Drinfel’d for the Yangian Reference 15. The proof of the ‘only if’ part when is finite dimensional was given by Chari-Pressley Reference 10, 12.2.6. Since the proof is based on a reduction to the case , it can be applied to a general Kac-Moody algebra (not necessarily symmetric). The ‘if’ part was proved by them later in Reference 11 when is finite dimensional, again not necessarily symmetric. As an application of the main result of this paper, we will prove the converse for a symmetric Kac-Moody algebra in §13. Our proof is independent of Chari-Pressley’s proof.

Remark 1.2.18.

The polynomials are called Drinfel’d polynomials.

When the Drinfel’d polynomials are given by

for some , , the corresponding simple l-highest weight module is called an l-fundamental representation. When is finite dimensional, is a Hopf algebra since Drinfel’d Reference 15 announced and Beck Reference 5 proved that can be identified with (a quotient of) the specialized quantized enveloping algebra associated with Cartan data of affine type. Thus a tensor product of -modules is again a -module. We have the following:

Proposition 1.2.19 (Reference 10, 12.2.6,12.2.8).

Suppose is finite dimensional.

(1) If and are simple l-highest weight -modules with Drinfel’d polynomials , such that is simple, then its Drinfel’d polynomial is given by

(2) Every simple l-highest weight -module is a subquotient of a tensor product of l-fundamental representations.

Unfortunately the coproduct is not defined for general as far as the author knows. Thus the above results do not make sense for general .

1.3. An l-weight space decomposition

Let be an l-integrable -module with the weight space decomposition . Since the commutative subalgebra preserves each , we can further decompose into a sum of generalized simultaneous eigenspaces for :

where is a pair as before and

If , we call an l-weight space, and the corresponding an l-weight. This is a refinement of the weight space decomposition. A further study of the l-weight space decomposition will be given in §13.5.

Motivated by Proposition 1.2.16, we introduce the following notion:

Definition 1.3.2.

An l-weight is said to be l-dominant if is dominant and there exists a polynomial for with such that Equation 1.2.17 holds.

Thus Proposition 1.2.16 means that an l-highest weight module is l-integrable if and only if the l-highest weight is l-dominant.

2. Quiver variety

2.1. Notation

Suppose that a finite graph is given and assume that there are no edge loops, i.e., no edges joining a vertex with itself. Let be the set of vertices and the set of edges. Let be the adjacency matrix of the graph, namely

We associate with the graph a symmetric generalized Cartan matrix , where is the identity matrix. This gives a bijection between the finite graphs without edge loops and symmetric Cartan matrices. We have the corresponding symmetric Kac-Moody algebra , the quantized enveloping algebra , the quantum affine algebra and the quantum loop algebra . Let be the set of pairs consisting of an edge together with its orientation. For , we denote by (resp. ) the incoming (resp. outgoing) vertex of . For we denote by the same edge as with the reverse orientation. Choose and fix an orientation of the graph, i.e., a subset such that , . The pair is called a quiver. Let us define matrices and by

So we have , .

Let be a collection of finite-dimensional vector spaces over for each vertex . The dimension of is a vector

If and are such collections, we define vector spaces by

For and , let us define a multiplication of and by

Multiplications , of , , are defined in an obvious manner. If , its trace is understood as .

For two collections , of vector spaces with , , we consider the vector space given by

where we use the notation unless we want to specify dimensions of , . The above three components for an element of will be denoted by , , respectively. An element of will be called an ADHM datum.

Usually a point in is called a representation of the quiver in the literature. Thus is the product of the space of representations of and that of . On the other hand, the factor or has never appeared in the literature.

Convention 2.1.4.

When we relate the quiver varieties to the quantum affine algebra, the dimension vectors will be mapped into the weight lattice in the following way:

where (resp. ) is the th component of (resp. ). Since and are both linearly independent, these maps are injective. We consider and as elements of the weight lattice in this way hereafter.

For a collection of subspaces of and , we say is -invariant if .

Fix a function such that for all . In Reference 44Reference 45, it was assumed that takes its value , but this assumption is not necessary as remarked by Lusztig Reference 38. For , let us denote by data given by for .

Let us define a symplectic form on by

Let be the algebraic group defined by

where we use the notation when we want to emphasize the dimension. It acts on by

preserving the symplectic form . The moment map vanishing at the origin is given by

where the dual of the Lie algebra of is identified with the Lie algebra via the trace. Let be an affine algebraic variety (not necessarily irreducible) defined as the zero set of .

For , we consider the complex

where is the differential of at , and is given by

If we identify with its dual via the symplectic form , is the transpose of .

2.2. Two quotients and

We consider two types of quotients of by the group . The first one is the affine algebro-geometric quotient given as follows. Let be the coordinate ring of the affine algebraic variety . Then is defined as a variety whose coordinate ring is the invariant part of :

As before, we use the notation unless we need to specify the dimension vectors , . By the geometric invariant theory Reference 43, this is an affine algebraic variety. It is also known that the geometric points of are closed -orbits.

For the second quotient we follow A. King’s approach Reference 27. Let us define a character by for . Set

The direct sum with respect to is a graded algebra, hence we can define

These are what we call quiver varieties.

2.3. Stability condition

In this subsection, we shall give a description of the quiver variety which is easier to deal with. We again follow King’s work Reference 27.

Definition 2.3.1.

A point is said to be stable if the following condition holds:

if a collection of subspaces of is -invariant and contained in , then .

Let us denote by the set of stable points.

Clearly, the stability condition is invariant under the action of . Hence we may say an orbit is stable or not.

Let us lift the -action on to the trivial line bundle by .

We have the following:

Proposition 2.3.2.

(1) A point is stable if and only if the closure of does not intersect with the zero section of for .

(2) If is stable, then the differential is surjective. In particular, is a nonsingular variety.

(3) If is stable, then in Equation 2.1.8 is injective.

(4) The quotient has a structure of nonsingular quasi-projective variety of dimension , and is a principal -bundle over .

(5) The tangent space of at the orbit is isomorphic to the middle cohomology group of Equation 2.1.8.

(6) The variety is isomorphic to .

(7) has a holomorphic symplectic structure as a symplectic quotient.

Proof.

See Reference 45, 3.ii and Reference 44, 2.8.

Notation 2.3.3.

For a stable point , its -orbit considered as a geometric point in the quiver variety is denoted by . If has a closed -orbit, then the corresponding geometric point in will be denoted also by .

From the definition, we have a natural projective morphism (see Reference 45, 3.18)

If , then is the unique closed orbit contained in the closure of . For , let

If we want to specify the dimension, we denote the above by . Unfortunately, this notation conflicts with the previous notation when . And the central fiber plays an important role later. We shall always write for and not use the notation Equation 2.3.5 with .

In order to explain a more precise relation between and , we need the following notion.

Definition 2.3.6.

Suppose that and a -invariant increasing filtration

with are given. Then set and . Let denote the endomorphism which induces on . For , let be such that its composition with the inclusion is , and let be the restriction of to . For , set and . Let be the direct sum of considered as data on .

Proposition 2.3.7.

Suppose . Then there exist a representative of and a -invariant increasing filtration as in Definition 2.3.6 such that is a representative of on .

Proof.

See Reference 45, 3.20

Proposition 2.3.8.

is a Lagrangian subvariety which is homotopic to .

Proof.

See Reference 44, 5.5, 5.8.

2.4. Hyper-Kähler structure

We briefly recall hyper-Kähler structures on , . This viewpoint was used for the study of , in Reference 44. (Caution: The following notation is different from the original one. and were denoted by and respectively in Reference 44. in Equation 2.1.7 was denoted by and the pair was denoted by in Reference 44.)

Put and fix hermitian inner products on and . They together with an orientation induce a hermitian inner product and a quaternion structure on (Reference 44, p.370). Let be a compact Lie group defined by . This is a maximal compact subgroup of , and acts on preserving the hermitian and quaternion structures. The corresponding hyper-Kähler moment map vanishing at the origin decomposes into the complex part (defined in Equation 2.1.7) and the real part , where

Proposition 2.4.1.

(1) A -orbit in intersects with if and only if it is closed. The map

is a homeomorphism.

(2) Choose a parameter so that . Then a -orbit in intersects with if and only if it is stable. The map

is a homeomorphism.

Proof.

See Reference 44, 3.1,3.2,3.5.

2.5

Suppose is a collection of subspaces of and is given. We can extend to by letting it equal on a complementary subspace of in . This operation induces a natural morphism

where . This induces a morphism

Moreover, we also have a map

Thus closed -orbits in are mapped to closed -orbits in by Proposition 2.4.1(1).

The following lemma was stated in Reference 45, p.529 without proof.

Lemma 2.5.3.

The morphism Equation 2.5.2 is injective.

Proof.

Suppose , have the same image under Equation 2.5.2. We choose representatives , which have closed -orbits.

Let us define () by

Choose complementary subspaces of in . We choose a -parameter subgroup as follows: on and on . Then the limit exists and its restriction to is . Since has a closed orbit, we may assume that the restriction of to is . Note that is a subspace of by the construction.

Suppose that there exists such that . We want to construct such that . Since we have , the restriction of to is invertible. Let be an extension of the restriction to so that is mapped to . Then maps to .

Hereafter, we consider as a subset of . It is clearly a closed subvariety. Let

If the graph is of finite type, stabilizes at some (see Proposition 2.6.3 and Lemma 2.9.4(2) below). This is not true in general. However, it presents no harm in this paper. We use to simplify the notation, and do not need any structures on it. We can always work on individual , not on .

Later, we shall also study for various simultaneously. We introduce the following notation:

Note that there are no obvious morphisms between and since the stability condition is not preserved under Equation 2.5.1.

2.6. Definition of

Let us introduce an open subset of (possibly empty):

Proposition 2.6.2.

If , then it is stable. Moreover, induces an isomorphism .

Proof.

See Reference 45, 3.24 or Reference 44, 4.1(2).

As in §2.5, we consider as a subset of when . Then we have

Proposition 2.6.3.

If the graph is of type , then

where the summation runs over the set of such that .

Proof.

See Reference 44, 6.7, Reference 45, 3.28.

Definition 2.6.4.

We say a point is regular if it is contained in for some . The above proposition says that all points are regular if the graph is of type . But this is not true in general (see Reference 45, 10.10).

2.7. -action

Let us define a -action on and , where . (Caution: We use the same notation and , but their roles are totally different.)

The -action is simply defined by its natural action on . It preserves the equation and commutes with the -action given by Equation 2.1.6. Hence it induces an action on and .

The -action is slightly different from the one given in Reference 45, 3.iv, and we need extra data. For each pair such that , we introduce and fix a numbering , , …, on edges joining and . It induces a numbering , …, , , …, on oriented edges between and . Let us define by

Then we define a -action on by

The equation is preserved since the left hand side is multiplied by . It commutes with the -action and preserves the stability condition. Hence it induces a -action on and . This -action makes the projective morphism equivariant.

In order to distinguish this -action from the -action Equation 2.1.6, we denote it as

2.8. Notation for -action

For an integer , we define a -module structure on by

and denote it by . For a -module , we use the following notational convention:

We use the same notation later when is an element of -equivariant -theory.

2.9. Tautological bundles

By the construction of , we have a natural vector bundle

associated with the principal -bundle . For abuse of notation, we denote it also by . It naturally has the structure of a -equivariant vector bundle. Letting act trivially, we make it a -equivariant vector bundle.

The vector space is also considered as a -equivariant vector bundle, where acts naturally and acts trivially.

We call and tautological bundles.

We consider , , as vector bundles defined by the same formula as in Equation 2.1.2. By the definition of tautological bundles, , , can be considered as sections of those bundles. Those bundles naturally have structures of -equivariant vector bundles. But we modify the -action on by letting act by on the component . This makes an equivariant section of .

We consider the following -equivariant complex over (cf. Reference 45, 4.2):

where

Let us explain the factor . Set . Since the -action in Equation 2.7.2 is defined so that

has weights , , …, , the above can be written as

in the notation Equation 2.8.2. By the same reason is an equivariant complex.

We assign degree to the middle term. (This complex is the complex in Reference 45, 4.2 with a modification of the -action.)

Lemma 2.9.2.

Fix a point and consider as a complex of vector spaces. Then is injective.

Proof.

See Reference 45, p.530. (Lemma 54 therein is a misprint of Lemma 5.2.)

Note that is not surjective in general. In fact, the following notion will play a crucial role later. Let be an irreducible component of for . Considering at a generic element of , we set

Lemma 2.9.4.

(1) Take and fix a point . Let be as in Equation 2.9.1. If , then we have

Moreover, the converse holds if we assume is regular in the sense of Definition 2.6.4. Namely under this assumption, if and only if Equation 2.9.5 holds.

(2) If , then is dominant.

Proof.

(1) See Reference 45, 4.7 for the first assertion. During the proof of Reference 45, 7.2, we have shown the second assertion, using Reference 45, 3.10 = Proposition 2.3.7.

(2) Consider the alternating sum of dimensions of the complex . It is equal to the alternating sum of dimensions of cohomology groups. It is nonnegative, if by Lemma 2.9.2 and (1). On the other hand, it is equal to

Thus we have the assertion.

3. Stratification of

As was shown in Reference 44, §6 and Reference 45, 3.v, there exists a natural stratification of by conjugacy classes of stabilizers. A local topological structure of a neighborhood of a point in a stratum (e.g., the homology group of the fiber of ) was studied in Reference 44, 6.10. We give a refinement in this section. We define a slice to a stratum, and study a local structure as a complex analytic space. Our technique is based on work of Sjamaar-Lerman Reference 50 in the symplectic geometry and hence our transversal slice may not be algebraic. It is desirable to have a purely algebraic construction of a transversal slice, as Maffei did in a special case Reference 42.

We fix dimension vectors , and denote , by , in this section.

3.1. Stratification

Definition 3.1.1 (cf. Sjamaar-Lerman Reference 50).

For a subgroup of denote by the set of all points in whose stabilizer is conjugate to . A point is said to be of -orbit type if its representative is in . The set of all points of orbit type is denoted by .

The stratum corresponding to the trivial subgroup is by definition. We have the following decomposition of :

where the summation runs over the set of all conjugacy classes of subgroups of .

For a more detailed description of , see Reference 44, 6.5, Reference 45, 3.27.

3.2. Local normal form of the moment map

Let us recall the local normal form of the moment map following Sjamaar-Lerman Reference 50.

Take and fix its representative . We suppose has a closed -orbit and satisfies by Proposition 2.4.1(1). Let be the stabilizer of . It is the complexification of the stabilizer in (see, e.g., Reference 51, 1.6). Since , the -orbit through is an isotropic submanifold of . Let be the quotient vector space , where is the tangent space of the orbit , and is the symplectic perpendicular of in , i.e., . This is naturally a symplectic vector space. A vector bundle over is called the symplectic normal bundle. (In general, the symplectic normal bundle of an isotropic submanifold is defined by .) It is isomorphic to . (In Reference 44, p.388, was defined as the orthogonal complement of the quaternion vector subspace spanned by with respect to the Riemannian metric.) The action of on preserves the induced symplectic structure on . Let be the corresponding moment map vanishing at the origin.

We choose an -invariant splitting and its dual splitting . Let us consider the natural action of on the product . With the natural symplectic structure on , we have the moment map

where is the projection of to . Zero is a regular value of , hence the symplectic quotient is a symplectic manifold. It can be identified with via the map

The embedding into is isotropic and its symplectic normal bundle is . Thus two embeddings of , one into and the other into , have the isomorphic symplectic normal bundles.

The -equivariant version of Darboux-Moser-Weinstein’s isotropic embedding theorem (a special case of Reference 50, 2.2) says the following:

Lemma 3.2.1.

A neighborhood of (in ) is -equivalently symplectomorphic to a neighborhood of embedded as the zero section of with the -moment map given by the formula

(Here ‘symplectomorphic’ means that there exists a biholomorphism intertwining symplectic structures.)

Note that Sjamaar-Lerman worked on a real symplectic manifold with a compact Lie group action. Thus we need to take care when applying their result to our situation. Darboux-Moser-Weinstein’s theorem is based on the inverse function theorem, which we have both in the category of -manifolds and in that of complex manifolds. A problem is that the domain of the symplectomorphism may not be chosen so that it covers the whole as it is noncompact. We can overcome this problem by taking a symplectomorphism defined in a neighborhood of the compact orbit first, and then extending it to a neighborhood of , as explained in the next three paragraphs. This approach is based on a result in Reference 51.

A subset of a -space is called orbitally convex with respect to the -action if it is invariant under (= maximal compact subgroup of ) and for all and all we have that both and in implies that for all . By Reference 51, 1.4, if and are complex manifolds with -actions, and if is an orbitally convex open subset of and is a -equivariant holomorphic map, then can be uniquely extended to a -equivariant holomorphic map.

Suppose that is a Kähler manifold with a (real) moment map and that is a point such that is fixed under the coadjoint action of . Then Reference 51, Claim 1.13 says that the compact orbit possesses a basis of orbitally convex open neighborhoods.

In our situation, we have a Kähler metric (§2.4) and we have assumed . Thus possesses a basis of orbitally convex open neighborhoods, and we have Lemma 3.2.1.

Now we want to study local structures of , using Lemma 3.2.1. First the equation implies , . Thus and are locally isomorphic to ‘quotients’ of by , i.e., ‘quotients’ of by . Here the ‘quotients’ are taken in the sense of the geometric invariant theory. Following Proposition 2.3.2(1), we say a point is stable if the closure of does not intersect with the zero section of for . Here we lift the -action to the trivial line bundle by , where is the restriction of the one-parameter subgroup used in §2.2. Let be the set of stable points. As in §2.3, we have a morphism , which we denote by . By Reference 51, Proposition 2.7, we may assume that the neighborhood of in Lemma 3.2.1 is saturated, i.e., the closure of the -orbit of a point in the neighborhood is contained in the neighborhood. Thus under the symplectomorphism in Lemma 3.2.1, (i) closed -orbits are mapped to closed -orbits, and (ii) the stability conditions are interchanged.

Proposition 3.2.2.

There exist a neighborhood (resp. ) of (resp. ) and biholomorphic maps , such that the following diagram commutes:

In particular, is biholomorphic to .

Furthermore, under , a stratum of is mapped to a stratum , which is defined as in Definition 3.1.1. (If intersects with , then is conjugate to a subgroup of .)

The above discussion shows Proposition 3.2.2 except for the last assertion. The last assertion follows from the argument in Reference 50, p.386.

3.3. Slice

By Reference 44, p.391, we have a -invariant splitting , where is the tangent space of the stratum containing , and acts trivially on . Thus we have

Furthermore, it was proved that and are quiver varieties associated with a certain graph possibly different from the original one, and possibly with edge loops. Replacing if necessary, we may assume that is a product of a neighborhood of in and of in We define a transversal slice to at as

Since is a local biholomorphism, this slice satisfies the properties in Reference 13, 3.2.19, i.e., there exists a biholomorphism

which induces biholomorphisms between factors

Remark 3.3.1.

Our construction gives a slice to a stratum in

for general . (See Reference 44, p.371 and Theorem 3.1 for the definition of .) In particular, the fiber of is isomorphic to the fiber of at . This is a refinement of Reference 44, 6.10, where an isomorphism between homology groups was obtained. We also remark that this gives a proof of smallness of

which was observed by Lusztig when is of type Reference 40. An essential point is, as remarked in Reference 44, 6.11, that is diffeomorphic to an affine algebraic variety, and its homology group vanishes for degree greater than its complex dimension.

For our application, we only need the case when is regular, i.e., for some . Then, by Reference 44, p.392, and are isomorphic to the quiver varieties and , associated with the original graph with dimension vector

where

in Convention 2.1.4.

Theorem 3.3.2.

Suppose that as above. Then there exist neighborhoods , , of , , respectively, and biholomorphic maps , such that the following diagram commutes:

In particular, is biholomorphic to .

Furthermore, a stratum of is mapped to a product of and a stratum of .

Remark 3.3.3.

Suppose that is a subgroup of fixing . Since the -action commutes with the -action, has an -action. The above construction can be made -equivariant. In particular, the diagram in Theorem 3.3.2 can be restricted to a diagram for -fixed point sets.

4. Fixed point subvariety

Let be an abelian reductive subgroup of . In this section, we study the -fixed point subvarieties and of and .

4.1. A homomorphism attached to a component of

Suppose that is fixed by . Take a representative of . For every , there exists such that

where the left hand side is the action defined in Equation 2.7.2 and the right hand side is the action defined in Equation 2.1.6. By the freeness of the -action on (see Proposition 2.3.2), is uniquely determined by . In particular, the map is a homomorphism.

Let be the set of fixed points such that Equation 4.1.1 holds for some representative of . Note that depends only on the -conjugacy class of . Since the -conjugacy class of is locally constant on , is a union of connected components of . Later we show that is connected under some assumptions (see Theorem 5.5.6). As in Proposition 2.3.8, we have

Proposition 4.1.2.

is homotopic to .

We regard as an -module via and consider the weight space which corresponds to :

We denote by the component of at the vertex . We have . We regard as an -module via the composition

We also have the weight space decomposition , . We denote by the composition

Then Equation 4.1.1 is equivalent to

where is as in Equation 2.7.1.

Lemma 4.1.4.

If , then for some and .

Proof.

Consider satisfying for any , . If we set

then is -invariant and contained in by Equation 4.1.3. Thus we have by the stability condition.

The restriction of tautological bundles , to are bundles of -modules. We have the weight decomposition , . We consider , as vector bundles over .

Similarly, the restriction of the complex in Equation 2.9.1 decomposes as , where

Here , are restrictions of , . When we want to emphasize that this is a complex over , we denote this by .

The tangent space of at is the -fixed part of the tangent space of . Since the latter is the middle cohomology group of Equation 2.1.8, the former is the middle cohomology group of the complex

where the differentials are the restrictions of , in Equation 2.1.8. Those restrictions are injective and surjective respectively by Proposition 2.3.2. Hence we have the following dimension formula:

Recall that we have an isomorphism (Proposition 2.6.2). Let

By definition, is an open subvariety of which is isomorphic to under .

4.2. A sufficient condition for

Let be a semisimple element in and let be the Zariski closure of .

Definition 4.2.1.

We say is generic if for any . (This condition depends on .)

Proposition 4.2.2.

Assume that there is at most one edge joining two vertices of , and that

for any pair of eigenvalues of . (The condition for the special case implies that is not a root of unity.) Then is generic.

Proof.

We prove by induction on . The assertion is trivial when .

Take a point in and its representative . As in Equation 4.1.1, there exists such that

We decompose into eigenspaces of :

We also decompose into eigenspaces of as . Then Equation 4.1.3 holds where is replaced by .

Choose and fix an eigenvalue of . First suppose for some . Let

Since is not a root of unity, we have for . Hence the above is well defined. By Equation 4.1.3 (and from the assumption), we have . By the assumption, we have , and hence again by Equation 4.1.3. Then we may assume the restriction of to is as in the proof of Lemma 2.5.3.

Thus the data is defined on the smaller subspace . Thus by the induction hypothesis.

If for any , we replace . If we can find a so that for some , we are done. Otherwise, we have for any , , and we have by Equation 4.1.3. Then we choose , which may not be an eigenvalue of , so that and repeat the above argument. (This is possible since we may assume .) We have and the data is defined on the smaller subspace as above.

5. Hecke correspondence and induction of quiver varieties

5.1. Hecke correspondence

Take dimension vectors , , such that . Choose collections of vector spaces , , , with , .

Let us consider the product . We denote by (resp. ) the vector bundle (resp. ). A point in is denoted by . We regard , , () as homomorphisms between tautological bundles.

We define a three-term sequence of vector bundles over by

where

This is a complex, that is, , thanks to the equations and . Moreover, it is an equivariant complex with respect to the -action.

By Reference 45, 5.2, is injective and is surjective. Hence the quotient is a -equivariant vector bundle. Let us define an equivariant section of by

where follows from and . The point is contained in the zero locus of if and only if there exists such that

Moreover, is zero by the stability condition for . Hence is a subspace of with dimension which is -invariant and contains . Moreover, such is unique if we fix representatives and . Hence we have an isomorphism between and the variety of all pairs and (modulo -action) such that

(a)

is stable, and

(b)

is a -invariant subspace containing the image of with .

Definition 5.1.4.

We call the Hecke correspondence, and denote it by . It is a -invariant closed subvariety.

Introducing a connection on , we consider the differential

of the section . Its restriction to is independent of the connection. By Reference 45, 5.7, the differential is surjective over . Hence, is nonsingular.

By the definition, the quotient defines a line bundle over .

5.2. Hecke correspondence and fixed point subvariety

Let be as in §4 and let be as in §4.1 for .

Let us consider the intersection . It decomposes as

Take a point . Then we have

and there exists such that

By the uniqueness of , we must have , that is, is -equivariant. Since is injective, can be considered as an -submodule of .

If and are the weight decompositions, then there exists such that

(a)

is an isomorphism if or ,

(b)

is a codimension embedding.

5.3

We introduce a generalization of the Hecke correspondence. Let us define as

(a)

,

(b)

is a -invariant subspace containing the image of with .

For , it is just . We consider as a closed subvariety of by setting

We have a vector bundle of rank defined by .

We shall show that is nonsingular later (see the proof of Lemma 11.2.3).

5.4. Induction

We recall some results in Reference 45, §4. Let be the middle cohomology of the complex Equation 2.9.1, i.e.,

We introduce the following subsets of (cf. Reference 34, 12.2):

Since is an open subset of , is a locally closed subvariety. The restriction of to is a -equivariant vector bundle of rank , where we used Convention 2.1.4.

Replacing by , we have a natural map

Note that the projection Equation 2.3.4 factors through . In particular, the fiber of is preserved under .

Proposition 5.4.3.

Let be the Grassmann bundle of -planes in the vector bundle obtained by restricting to . Then we have the following diagram:

where is the natural projection, and and are restrictions of the projections to the first and second factors. The kernel of the natural surjective homomorphism is isomorphic to the tautological vector bundle of the Grassmann bundle of the first row, and also to the the restriction of the vector bundle over in the third row.

Proof.

The proof is essentially contained in Reference 45, 4.5. See also Proposition 5.5.2 for a similar result.

5.5. Induction for fixed point subvarieties

We consider an analogue of the results in the previous subsection for fixed point subvariety . Let us use notation as in §4.1, and suppose that is the Zariski closure of a semisimple element .

Let be the middle cohomology of the complex in Equation 4.1.5, i.e.,

Let

Replacing by , we have a natural map

where () is the homomorphism obtained from by replacing by its codimension subspace. Its conjugacy class is independent of the choice of the subspace. This map is just the restriction of in the previous subsection.

For each , let denote the Grassmann bundle of -planes in the vector bundle obtained by restricting to . Let

be their fiber product over .

We have the following analogue of Proposition 5.4.3:

Proposition 5.5.2.

Suppose that . We have the following diagram:

where is the natural projection. For each , the kernel of the natural surjective homomorphism is isomorphic to the tautological vector bundle of the Grassmann bundle. Moreover, we have

(Here of a complex means the alternating sum of dimensions of cohomology groups.)

Proof.

We have a surjective homomorphism of codimension over . This gives a morphism from to the fiber product of Grassmann bundles. By a straightforward modification of the arguments in Reference 45, 4.5, one can show that it is an isomorphism. The details are left to the reader. The assumption is used to distinguish and .

Let us prove the remaining part Equation 5.5.3, Equation 5.5.4. First note that

Since we have an -dimensional subspace in , we must have

Replacing by , we get Equation 5.5.3.

Moreover, we have

On the other hand, the dimension formula Equation 4.1.6 implies

Since is an open subset of , we get Equation 5.5.4.

Note that the inequality Equation 5.5.3 implies that

and the equality holds if and only if

In particular, we have the following analog of Reference 45, 4.6.

Corollary 5.5.5.

Suppose . On a nonempty open subset , we have

for each . Also, the complement is a lower dimensional subvariety of .

As an application of this induction, we prove the following:

Theorem 5.5.6.

Assume that is not a root of unity and there is at most one edge joining two vertices of . Then is connected if it is a nonempty set.

Proof.

We prove the assertion by induction on , . (The result is trivial when .)

We first make a reduction to the case when

Fix a and consider

Since is not a root of unity, we have for . Hence the above is well defined.

Suppose . By Equation 4.1.3 and the choice of , we have

Let us replace by . Namely we change to and all other data are unchanged. The equation and the stability condition are preserved by the replacement. Thus we have a morphism

where is a fixed point subvariety of obtained by the replacement. (This notation will not be used elsewhere. The data is fixed elsewhere.)

Conversely, we can put any homomorphism to get a point in starting from a point in . This shows that is the total space of the vector bundle over , where is considered as a trivial bundle. In particular, is (nonempty and) connected if and only if is also. By the induction hypothesis, is connected and we are done.

Thus we may assume . Then consists of the last term by the choice of . (Note under the assumption that there is at most one edge joining two vertices of .) Hence we have Equation 5.5.7 with .

Now let us prove the connectedness of under Equation 5.5.7. By Corollary 5.5.5, we have

unless for each . Hence it is enough to prove the connectedness of for .

Let us consider the map . By Equation 5.5.7, becomes smaller for . Hence is connected by the induction hypothesis. Again by Corollary 5.5.5, is also connected. Since is a fiber product of Grassmann bundles, is connected.

6. Equivariant -theory

In this section, we review the equivariant -theory of a quasi-projective variety with a group action. See Reference 13, Chapter 5 for further details.

6.1. Definitions

Let be a quasi-projective variety over . Suppose that a linear algebraic group acts algebraically on . Let be the Grothendieck group of the abelian category of -equivariant coherent sheaves on . It is a module over , the representation ring of .

A class in represented by a -equivariant sheaf will be denoted by , or simply by if there is no fear of confusion.

The trivial line bundle of rank , i.e., the structure sheaf, is denoted by . If the underlying space is clear, we simply write .

Let be the Grothendieck group of the abelian category of -equivariant algebraic vector bundles on . This is also an -module. The tensor product defines a structure of an -algebra on . Also, has a structure of a -module by the tensor product:

Suppose that is a -invariant closed subvariety of and let be the complement. Two inclusions

induce an exact sequence

where is given by and is given by . (See Reference 53.)

Suppose that is a -invariant closed subvariety of and that is nonsingular. Let be the Grothendieck group of the derived category of -equivariant complexes of algebraic vector bundles over , which are exact outside (see Reference 4, §1). We have a natural homomorphism by setting

Here is the th cohomology sheaf of , which is a -equivariant coherent sheaf on which is supported on . If is the defining ideal of , we have for sufficiently large . Then

is a sheaf on , and defines an element in . Conversely if a -equivariant coherent sheaf on is given, we can take a resolution by a finite -equivariant complex of algebraic vector bundles:

where denotes the inclusion. (See Reference 13, 5.1.28.) This shows that the homomorphism is an isomorphism. This relative -group was not used in Reference 13 explicitly, but many operations were defined by using it implicitly. When , is isomorphic to . In particular, we have an isomorphism if is nonsingular.

We shall also use equivariant topological -homology . There are several approaches for the definition, but we take the one in Reference 54, 5.3. There is a comparison map

which satisfies obvious functorial properties.

Occasionally, we also consider the higher equivariant topological -homology group . (See Reference 54, 5.3 again.) In this circumstance, may be written as . But we do not use higher equivariant algebraic -homology .

Suppose that is a -invariant closed subvariety of and let be the complement. Two inclusions

induce a natural exact hexagon

for suitably defined , .

6.2. Operations on -theory of vector bundles

If is a -equivariant vector bundle, its rank and dual vector bundle will be denoted by and respectively.

We extend and to operations on :

where is the set of the connected components of . Note that the rank of a vector bundle may not be a constant, when has several connected components. But we assume is connected in this subsection for simplicity. In general, operators below can be defined component-wisely.

If is a -equivariant line bundle, we define for . Thus we have for .

If is a vector bundle, we define

These operations can be extended to of -equivariant algebraic vector bundles:

This is well defined since we have , for an exact sequence .

Note the formula

for a vector bundle . Using this formula, we expand into the Laurent expansion also at :

6.3. Tor-product

(Cf. Reference 4, 1.3, Reference 13, 5.2.11.) Let be a nonsingular quasi-projective variety with a -action. Let , be -invariant closed subvarieties of . Suppose that (resp. ) is a -equivariant complex of vector bundles over which is exact outside (resp. ). Then we can construct a complex

with suitably defined differentials from the double complex . It is exact outside . This construction defines an -bilinear pairing

Since we assume is nonsingular, we have , , . Thus we also have an -bilinear pairing

We denote these operations by . (It is denoted by in Reference 13.)

Lemma 6.3.1 (Reference 13, 5.4.10, Reference 58, Lemma 1).

Let , be nonsingular -subvarieties with conormal bundles , . Suppose that is nonsingular and , where means the restriction to . Then for any , , we have

where .

6.4. Pull-back with support

(Cf. Reference 4, 1.2, Reference 13, 5.2.5.) Let be a -equivariant morphism between nonsingular -varieties. Suppose that and are -invariant closed subvarieties of and respectively satisfying . Then the pull-back

induces a homomorphism . Via isomorphisms , , we get a homomorphism . Note that this depends on the ambient spaces , .

Let as above. Suppose that , are -invariant closed subvarieties such that for . Then we have

for ().

6.5. Push-forward

Let be a proper -equivariant morphism between -varieties (not necessarily nonsingular). Then we have a push-forward homomorphism defined by

Suppose further that and are nonsingular. If , are -invariant closed subvarieties, we have the following projection formula (Reference 13, 5.3.13):

for , .

6.6. Chow group and homology group

Let be the integral Borel-Moore homology of . Let be the Chow group of . We have a cycle map

which has certain functorial properties (see Reference 19, Chapter 19).

If is a closed subvariety of and is its complement, then we have exact sequences which are analogues of Equation 6.1.2, Equation 6.1.3:

We have operations on and which are analogues of those in §6.3, §6.4, §6.5. (See Reference 19.)

In the next section, we prove results for -homology and the Chow group in parallel arguments. It is the reason why we avoid higher algebraic -homology. There is no analogue for the Chow group.

7. Freeness

7.1. Properties , ,

Following Reference 14Reference 39, we say that an algebraic variety has property if

(a)

and is a free abelian group.

(b)

The cycle map is an isomorphism.

Similarly, we say has property if

(a)

and is a free abelian group.

(b)

The comparison map is an isomorphism.

Suppose that is a closed subvariety of a nonsingular variety . We have a diagram (see Reference 4)

where the horizontal arrows are local Chern character homomorphisms in algebraic and topological -homologies respectively, the left vertical arrow is a comparison map, and the right vertical arrow is the cycle map. It is known that the upper horizontal arrow is an isomorphism (Reference 19, 15.2.16). Thus the composite is an isomorphism if has property .

Assume that is nonsingular and projective. We define the bilinear pairing by

where is the canonical map from to the point.

We say that has property if has property and the pairing Equation 7.1.1 is perfect. (In Reference 39, this property is called .)

Let be a linear algebraic group. Let be an algebraic variety with a -action. We say that has property if

(a)

and is a free -module.

(b)

The natural map is an isomorphism.

(c)

For a closed algebraic subgroup , -equivariant -theories satisfy the above properties (a), (b), and the natural homomorphism is an isomorphism.

Suppose further that is smooth and projective. By the same formula as Equation 7.1.1, we have a bilinear pairing . We say that has property if has property and this pairing is perfect.

A finite partition of a variety into locally closed subvarieties is said to be an -partition if the subvarieties in the partition can be indexed , …, in such a way that is closed in for , …, . The following is proved in Reference 14, Lemma 1.8.

Lemma 7.1.2.

If has an -partition into pieces which have property , then has property .

The proof is based on exact sequences Equation 6.6.1,Equation 6.6.2 in homology groups and Chow groups. Since we have corresponding exact sequences Equation 6.1.2,Equation 6.1.3 in -theory, we have the following.

Lemma 7.1.3.

Suppose that an algebraic variety has an action of a linear algebraic group . If has an -partition into -invariant locally closed subvarieties which have property , then has property .

Lemma 7.1.4.

Let be a -equivariant fiber bundle with affine spaces as fibers. Suppose that is locally a trivial -equivariant vector bundle, i.e., a product of base and a vector space with a linear -action. If has property (resp. ), then also has property (resp. ).

Proof.

We first show that is surjective. Choose a closed subvariety of so that is a trivial -bundle over . There is a diagram

with exact rows by Equation 6.1.2. By a diagram chase it suffices to prove the surjectivity for the restrictions of to and to . By repeating the process on , it suffices to prove it for the case when is a trivial -equivariant bundle. By Thom isomorphism Reference 53, 4.1 is an isomorphism if is a -equivariant bundle. Thus we prove the assertion.

Let us repeat the same argument for and by replacing Equation 6.1.2 by Equation 6.1.3. By the five lemma both are isomorphisms. In particular, we have by assumption.

Consider the diagram

where the vertical arrows are comparison maps. The left vertical arrow is an isomorphism by assumption. Thus the right vertical arrow is also an isomorphism by the commutativity of the diagram and what we just proved above. Condition (c) for can be checked in the same way, and has property .

Property can be checked in the same way.

Lemma 7.1.5.

Let be a nonsingular quasi-projective variety with -action with a Kähler metric such that

(a)

is complete,

(b)

is invariant under the maximal compact subgroup of ,

(c)

there exists a moment map associated with the Kähler metric and the -action (the maximal compact subgroup of the second factor), and it is proper.

Let

If the fixed point set has property (resp. ), then both and have property (resp. ).

Furthermore, the bilinear pairing

is nondegenerate if has property . A similar intersection pairing between and is nondegenerate if has property . Here is the canonical map from to the point.

Proof.

By Reference 2, 2.2 the moment map is a Bott-Morse function, and critical manifolds are the fixed point . Let , , …be the components of . By Reference 2, §3, stable and unstable manifolds for the gradient flow of coincide with -attracting sets of Bialynicki-Birula decomposition Reference 7:

These are invariant under the -action since the -action commutes with the -action.

Note that results in Reference 2 are stated for compact manifolds, but the argument can be modified to our setting. A difference is that is not unless is compact. On the other hand, is since is proper.

As in Reference 3, we can introduce an ordering on the index set of components of such that is an -partition and is an -partition with respect to the reversed order.

By Reference 7 (see also Reference 8 for analytic arguments), the maps

are fiber bundles with affine spaces as fibers. Furthermore, (resp. ) is locally isomorphic to a -equivariant vector bundle by the proof. Thus and have properties and by Lemma 7.1.4. Hence and have properties and by Lemmas 7.1.2 and 7.1.3.

By the argument in Reference 39, 1.7, 2.5, the pairing Equation 7.1.6 can be identified with a pairing

of the form

for some pairing such that is the pairing Equation 7.1.1 for . Since is nondegenerate for all by the assumption, Equation 7.1.6 is also nondegenerate.

The proof of the statement for , is similar. One uses the fact that the intersection pairing is nondegenerate under property .

7.2. Decomposition of the diagonal

Proposition 7.2.1 (cf. Reference 16, Reference 13, 5.6.1).

Let be a nonsingular projective variety.

(1) Let be the structure sheaf of the diagonal and the corresponding element in . Assume that

holds for some , . Then has property .

(2) Let be a linear algebraic group. Suppose that has -action and that Equation 7.2.2 holds in for some , . Then has property .

(3) Let be the class of the diagonal in . Assume that

holds for some , . Then has property .

Proof.

Let denote the projection to the th factor (). Let be the diagonal embedding . Then we have . Hence

If we substitute Equation 7.2.2 into the above, we get

In particular, is spanned by ’s.

If for some , then . Hence we have . The above equality Equation 7.2.4 implies . This means that is torsion-free. Thus we could assume the ’s are linearly independent in Equation 7.2.2. Under this assumption, is a basis of , and Equation 7.2.4 implies that is the dual basis.

If we perform the same computation in , we get the same result. In particular, is a basis of . However, , are in , thus we have . We also have is an isomorphism. Thus has property .

If has -action and Equation 7.2.2 holds in the equivariant -group, we do the same calculation in the equivariant -groups. Then the same argument shows that has property .

The assertion for Chow groups and homology groups can be proved in the same way.

7.3. Diagonal of the quiver variety

Let us recall the decomposition of the diagonal of the quiver variety defined in Reference 45, Sect. 6. In this section, we fix dimension vectors , and use the notation instead of .

Let us consider the product . We denote by (resp. ) the vector bundle (resp. ). A point in is denoted by . We regard , , () as homomorphisms between tautological bundles.

We consider the following -equivariant complex of vector bundles over :

where

It was shown that is injective and is surjective (cf. Reference 45, 5.2). Thus is an equivariant vector bundle. We define an equivariant section of by

Then is contained in the zero locus of if and only if there exists such that

Moreover is an isomorphism by the stability condition. Hence is equal to the diagonal . If is a connection on , the differential is surjective on (cf. Reference 45, 5.7). In particular, we have an exact sequence

In , is equal to the alternating sum of terms of Equation 7.3.1 which has a form for some . Hence satisfies the conditions of Proposition 7.2.1 except the projectivity. Unfortunately, the projectivity is essential in the proof of Proposition 7.2.1. (We could not define otherwise.) Thus Proposition 7.2.1 is not directly applicable to . In order to get rid of this difficulty, we consider the fixed point set with respect to the -action.

For technical reasons, we need to use a -action, which is different from Equation 2.7.2. Let act on by

This induces a -action on and which commutes with the previous -action. (If the adjacency matrix satisfies for any , then the new -action coincides with the old one.) The tautological bundles , become -equivariant vector bundles as before.

We consider the fixed point set . is a fixed point if and only if there exists a homomorphism such that

as in §4.1. Here denotes the new -action. We decompose the fixed point set according to the conjugacy class of :

Lemma 7.3.3.

is a nonsingular projective variety.

Proof.

Since is a union of connected components (possibly single component) of the fixed point set of the -action on a nonsingular variety , is nonsingular.

Suppose that is a fixed point of the -action. It means that lies in the closed orbit . But converges to as . Hence the closed orbit must be . Since is equivariant, is contained in . In particular, is projective.

This lemma is not true for the original -action.

We restrict the complex Equation 7.3.1 to . Then fibers of and become -modules and hence we can take the -fixed part of Equation 7.3.1:

where (resp. ) is the restriction of (resp. ) to the -fixed part. Then is injective and is surjective, and is a vector bundle which is the -fixed part of .

The section takes values in . Considering it as a section of , we denote it by . The zero locus is which is the diagonal of . Furthermore, the differential is surjective on .

Our original -action (defined in §2.7) commutes with the new -action. Thus has an induced -action. By the construction, is a -equivariant vector bundle, and is an equivariant section.

Proposition 7.3.4.

has properties and . Moreover, is connected.

Proof.

Let be the structure sheaf of the diagonal considered as a sheaf on . By the above argument, the Koszul complex of gives a resolution of :

where . Thus we have the following equality in the Grothendieck group :

Since is injective and is surjective, we have

Each factor of the right hand side can be written in the form for some . For example, the first factor is equal to

where is the weight space of , i.e.,

The remaining factors have a similar description. Thus by Proposition 7.2.1, has property .

Moreover, the above shows that is generated by exterior powers of , and its duals (as an -algebra). Note that these bundles have constant rank on . If have components , , …, the structure sheaf of (extended to by setting outside) cannot be represented by , . This contradiction shows that is connected.

The assertion for Chow groups can be proved in exactly the same way. By the above argument, the fundamental class is the top Chern class of , which can be represented as for some , .

Theorem 7.3.5.

and have properties and . Moreover, the bilinear pairing

is nondegenerate. A similar pairing between and is also nondegenerate. Here is the canonical map from to the point.

Proof.

We apply Lemma 7.1.5. By Reference 44, 2.8, the metric on defined in §2.4 is complete. By the construction, it is invariant under , where is the maximal compact subgroup of . (Note that the hyper-Kähler structure is not invariant under the -action, but the metric is invariant.) The moment map for the -actions is given by

This is a proper function on . Thus Lemma 7.1.5 is applicable. Note that we have as in Reference 44, 5.8. (Though our -action is different from the one in Reference 44, the same proof works.)

7.4. Fixed point subvariety

Let be an abelian reductive subgroup of as in §4. Let and be the fixed point set in and respectively. Exactly as in the previous subsection, we have the following generalization of Theorem 7.3.5.

Theorem 7.4.1.

and have properties and . Moreover, the bilinear pairing

is nondegenerate. A similar pairing between and is also nondegenerate. Here is the canonical map from to the point.

7.5. Connectedness of

Let us consider a natural homomorphism

which sends representations to bundles associated with tautological bundles. If we can apply Proposition 7.2.1 to , then this homomorphism is surjective. Unfortunately we cannot apply Proposition 7.2.1 since is not projective. However, it seems reasonable to conjecture that the homomorphism Equation 7.5.1 is surjective. In particular, it implies that is connected as in the proof of Proposition 7.3.4. This was stated in Reference 45, 6.2. But the proof contains a gap since the function may not be proper in general.

8. Convolution

Let , , be a nonsingular quasi-projective variety, and write for the projection ().

Suppose (resp. ) is a closed subvariety of (resp. ) such that the restriction of the projection is proper. Let . We can define the convolution product by

for .

Note that the convolution product depends on the ambient spaces , and . When we want to specify them, we say the convolution product relative to , , .

In this section, we study what happens when , , are replaced by

(a)

submanifolds , , of , , ,

(b)

principal -bundles , , over , , .

Although we work on nonequivariant -theory, the results extend to the case of equivariant -theory, the Borel-Moore homology group, or any other reasonable theory in a straightforward way.

8.1

Before studying the above problem, we recall the following lemma which will be used several times.

Lemma 8.1.1.

In the above setting, we further assume that and , where is the diagonal embedding . Then we have

for a vector bundle over , where is the projection, and in the right hand side is the tensor product Equation 6.1.1 between and .

The proof is obvious from the definition, and is omitted.

8.2. Restriction of the convolution to submanifolds

Suppose we have nonsingular closed submanifolds , , of , , such that

By this assumption, we have

Let (resp. ) be the intersection (resp. ). By Equation 8.2.2, we have . We have the convolution product relative to :

where is the projection .

We want to relate two convolution products and via pull-back homomorphisms. For this purpose, we consider the inclusion , where is the inclusion (). By Equation 8.2.1, we have a pull-back homomorphism

Similarly, we have

Proposition 8.2.3.

For , we have

Namely, the following diagram commutes:

Example 8.2.5.

Suppose , and , where is the diagonal embedding . Then the above assumption is satisfied, and we have , where is the diagonal embedding . If is a vector bundle over , we have by the base change Reference 13, 5.3.15. By Lemma 8.1.1, we have

where , are the projections. Note

by Equation 6.4.1. Hence we have Equation 8.2.4 in this case.

Proof of Proposition 8.2.3.

In order to relate relative to , , and relative to , , , we replace by factor by factor.

Step 1. First we want to replace by . We consider the following fiber square:

where is the projection. We have

where we have used the base change (Reference 13, 5.3.15) in the second equality and Equation 6.4.1 in the third equality. If denotes the projection, we have . Hence we get

Similarly, we have

where is the projection. Substituting this into Equation 8.2.6, we obtain

Step 2. Next we replace by . By Equation 8.2.1, we have a homomorphism

which is just the identity operator. We will consider as an element of or interchangeably. We consider the fiber square

where is the projection. By base change Reference 13, 5.3.15, we get

By the projection formula Equation 6.5.1, we get

Substituting this into Equation 8.2.7, we have

where , are the projections, and we have used and .

Step 3. We finally replace by . By Equation 8.2.1, we have a homomorphism

which is just the identity operator. We consider the fiber square

By base change Reference 13, 5.3.15, we get

Substituting this into Equation 8.2.8, we obtain

where we have used Equation 6.5.1 in the second equality and , in the third equality. Finally, by Equation 8.2.2, the homomorphism is just the identity operator . Thus we have the assertion.

8.3. Convolution and principal bundles

Let be a linear algebraic group and suppose that we have principal -bundles over for . Consider the restriction of the principal -bundle to for . Then the pull-back homomorphism gives a canonical isomorphism

Similarly we have an isomorphism

Let act on diagonally. We assume that there exists a closed -invariant subvariety of such that

for .

Let be the projection . Since the restriction of the projection is proper, we have the convolution product given by

for , . We want to compare this convolution product with that on .

By Equation 8.3.3, we have

Via Equation 8.3.1, we define

Thus we can consider the convolution product .

By the construction, we have

Noticing that the restriction of to is proper, we have

Combining this with Equation 8.3.2, we have

Proposition 8.3.5.

In the above setup, we have

Namely the following diagram commutes:

where the left vertical arrow is

and the right vertical arrow is

Example 8.3.7.

Suppose , and , where is the diagonal embedding . If we take , where is the diagonal embedding , the assumption Equation 8.3.3 is satisfied. In fact, the restriction of to is an isomorphism. Take a vector bundle and consider . By the isomorphism , we can define . Then both and is where is the natural isomorphism. Hence Equation 8.3.4 holds for and . By Lemma 8.1.1, we have

where , are the projections. We can directly check Equation 8.3.6 in this case.

Proof of Proposition 8.3.5.

As in the proof of Proposition 8.2.3, we replace by factor by factor.

Step 1. First we replace by . Consider the following fiber square:

where is the projection. By base change Reference 13, 5.3.15 and Equation 6.4.1, we have

where and are projections.

Step 2. Consider the fiber square

where is the projection. By base change Reference 13, 5.3.15, we have

where we have used Equation 8.3.4 for . Substituting this into Equation 8.3.8, we get

where we have used Equation 6.5.1 in the second equality. Let be the projection. By , we have

We also have

where is the projection. Substituting these two equalities into Equation 8.3.9, we get

Step 3. Consider the fiber square

By base change Reference 13, 5.3.15, we have

where we have used Equation 8.3.4 for in the second equality. Substituting this into Equation 8.3.10, we have

By and , we get

This proves our assertion.

9. A homomorphism

9.1

Let us define an analogue of the Steinberg variety by

Here means that is equal to if we regard both as elements of by Equation 2.5.4. This is a closed subvariety of .

The map is proper and its image is contained in . Hence we can define the convolution product on the equivariant -theory:

Let

be the subspace of consisting of elements such that

(1)

for fixed , for all but finitely many choices of ,

(2)

for fixed , for all but finitely many choices of .

The convolution product is well defined on . When the underlying graph is of type , is empty for all but finitely many choices of , so is just the direct product .

Let denote the disjoint union . When we write

we mean as convention.

The second projection induces a homomorphism . Thus is an -algebra. Moreover, is isomorphic to where corresponds to in Equation 2.8.1. Thus is a -algebra.

The aim of this section and the next two sections is to define the homomorphism from into . We first define the map on generators of , and then check the defining relation.

9.2

First we want to define the image of , .

Let be the -equivariant complex over defined in Equation 2.9.1. We consider as an element of by identifying it with the alternating sum

The rank of the complex Equation 2.9.1, as an element of (see §6.2), is given by

Let denote the diagonal embedding ;

where denotes the expansion at , respectively. Note that

9.3

Next we define the images of and . They are given by line bundles over Hecke correspondences.

Let , and be as in §5.1. By the definition, the quotient defines a line bundle over . The generator is very roughly defined as the th power of , but we need a certain modification in order to have the correct commutation relation.

For the modification, we need to consider the following variants of :

where means that we consider the complex consisting of in degrees and for other degrees. Since is injective, we have in . We have the corresponding decomposition of the Cartan matrix :

We identify (resp. ) with a map given by

We also need matrices , given by

where , are as in Equation 2.1.1. We also identify them with maps exactly as above.

Let be the exchange of factors. Let us denote by . As on , we have a natural line bundle over . Let us denote it by .

Now we define the images of , by

where and are the inclusions. Hereafter, we may omit or , hoping that it causes no confusion.

9.4

Theorem 9.4.1.

The assignments Equation 9.2.1, Equation 9.3.2 define a homomorphism

of -algebras.

We need to check the defining relations Equation 1.2.1, Equation 1.2.2, Equation 1.2.3, Equation 1.2.4, Equation 1.2.5, Equation 1.2.6, Equation 1.2.7, Equation 1.2.8, Equation 1.2.9, Equation 1.2.10, and Equation 1.2.11. We do not need to consider the relations Equation 1.2.1, Equation 1.2.5 because we are considering instead of . The relations Equation 1.2.2, Equation 1.2.3 and Equation 1.2.4 follow from Lemma 8.1.1 and the fact that . The relation Equation 1.2.6 also follows from Lemma 8.1.1. The remaining relations will be checked in the next two sections.

10. Relations (I)

10.1. Relation Equation 1.2.7

Fix a vertex and take , . Let be the inclusion and let and be the projections and respectively.

By Lemma 8.1.1, we have

We have the following equality in :

Hence we have

Substituting this into Equation 10.1.1, we get

This is equivalent to Equation 1.2.7 for . The relation Equation 1.2.7 for can be proved in the same way.

10.2. Relation Equation 1.2.8 for

Fix two vertices . Let , , , be dimension vectors such that

We want to compute and in the component .

Let us consider the intersection

in (resp. ). On the intersection, we have the inclusion of restrictions of tautological bundles

Lemma 10.2.1.

The above two intersections are transversal, and there is a -equivariant isomorphism between them such that

(a)

it is the identity operators on the factor and ,

(b)

it induces isomorphisms and .

Proof.

See Reference 45, Lemmas 9.8, 9.9, 9.10 and their proofs.

Since the intersection is transversal, we have

where is the following line bundle over :

Similarly, we have

where is the following line bundle over :

Let us compare Equation 10.2.2 and Equation 10.2.3. On , we have

On the other hand, we have

on . Hence under the isomorphism in Lemma 10.2.1, we obtain

where we have used

By

we have

Thus we have .

10.3. Relation Equation 1.2.10

We give the proof of Equation 1.2.10 for in this subsection. The relation Equation 1.2.10 for can be proved in a similar way, and hence is omitted.

Fix two vertices . Let , , , be dimension vectors such that

We want to compute and in the component .

Let us consider the intersection

in (resp. ).

Lemma 10.3.1.

The above two intersections are transversal respectively.

Proof.

The proof below is modeled on Reference 45, 9.8, 9.9. We give the proof for . Then the same result for follows by , .

We consider the complex Equation 5.1.1 for and :

where we use suffixes , to distinguish endomorphisms. We have sections and of and respectively.

Identifying these vector bundles and sections with those of pull-backs to , we consider their zero loci and .

As in the proof of Reference 45, 5.7, we consider the transpose of , via the symplectic form. Their sum gives a vector bundle endomorphism

It is enough to show that the kernel of is zero at .

Take representatives of (). Then we have , which satisfy Equation 5.1.3 for . Suppose that

lies in the kernel. Then there exist () such that

Then we have

Hence we have

by the stability condition.

Consider the equation Equation 10.3.3 at the vertex . Since , is an isomorphism. Hence Equation 10.3.3 implies that is invariant under . Since is a codimension subspace, the induced map is a scalar which we denote by . Moreover, there exists a homomorphism such that

For another vertex , is an isomorphism, hence we can define so that the same equation holds also for the vertex . Thus we have

Substituting Equation 10.3.4 into Equation 10.3.2 and using the injectivity of , we get

This means that .

Substituting Equation 10.3.4 into Equation 10.3.3 and noticing is injective, we obtain

Thus is invariant under Arguing as above, we can find a constant and a homomorphism such that

Substituting this equation into Equation 10.3.2, we get

This means that . Hence is injective.

Let us consider the variety (resp. ) of all pairs and satisfying the following:

(a)

is a subspace with ,

(b)

is -stable,

(c)

,

(d)

the induced homomorphism (resp. ) is zero for with , .

Then (resp. ) is isomorphic to (resp. ). The isomorphism is given by defining

where (resp. ) is given by

It is also clear that the restriction of to (resp. ) is an isomorphism onto its image. Hereafter, we identify (resp. ) with the image. Then and are closed subvarieties of . Let (resp. ) denote the inclusion.

The quotient (resp. ) forms a line bundle over

(resp. By the above consideration, (resp. ) is represented by

(resp.

Note that we have

Set . We consider

as a section of the vector bundle over . Let us denote it by . Similarly is a section (denoted by ) of the vector bundle over .

Lemma 10.3.9.

The section (resp. ) is transversal to the zero section (if it vanishes somewhere).

Proof.

Fix a subspace with . Let be the parabolic subgroup of consisting of elements which preserve . We also fix a complementary subspace . Thus we have . We will check the assertion for . The assertion for follows if we exchange and .

We consider

It is a linear subspace of . Let be the composition of the restriction of the moment map to and the projection . Let denote the set of which is stable. It is preserved under the action of and we have a -equivariant isomorphism

Note that the -part of the moment map vanishes on thanks to the definition of .

The assertion follows if we check that

is surjective at . Here is the natural projection. Thus it is enough to show that is surjective.

Suppose that is orthogonal to , namely

for any . Hence we have

where is the restriction of to . Therefore the image of is invariant under and contained in . By the stability condition, we have . Thus we have proved the assertion.

Let (resp. ) be the zero locus of (resp. ). By Lemma 10.3.9, we have the following exact sequence (Koszul complex) on (resp. ):

(resp.

Hence we have the following equality in (resp. ):

Both and consist of all pairs and satisfying the following:

(a)

is a subspace with ,

(b)

is -stable,

(c)

,

(d)

the induced homomorphism is zero,

modulo the action of . In particular, we have . Hence we have

This implies

by the projection formula Equation 6.5.1. Multiplying this equality by and taking the sum with respect to and , we get

Comparing this with Equation 10.3.7 and using Equation 10.3.8, we get Equation 1.2.10.

10.4. Relation Equation 1.2.11

We give the proof of Equation 1.2.11 for in this subsection, assuming other relations. (The relations Equation 1.2.8 with and Equation 1.2.9 will be checked in the next section, but its proof is independent of results in this subsection.) The relation Equation 1.2.11 for can be proved in a similar way, and hence is omitted.

By the proof of Reference 45, 9.3, operators , acting on are locally nilpotent. (See also Lemma 13.2.4 below.) It is known that the constant term of Equation 1.2.11, i.e.,

can be deduced from the other relations and the local nilpotency of , (see, e.g., Reference 13, 4.3.2 for the proof for ). Thus our task is to reduce

to Equation 10.4.1. This reduction was done by Grojnowski Reference 23, but we reproduce it here for the sake of completeness.

For , let be dimension vectors with

Let

be the projection. Let

This is equal to

(a)

(in ),

(b)

is a -invariant subspace containing the image of with .

In particular, we have line bundles on (, …, ). By the definition, there exists a line bundle such that

Moreover, we have

Now consider the symmetrization. By Equation 10.4.4, we have

for some tensor product of exterior products of the bundle and its dual. (In the notation in §11.4 below, corresponds to the symmetric function .) Note that we have . Thus can be considered as a vector bundle over . Then the projection formula implies that

Noticing that is independent of , we can derive Equation 10.4.2 from Equation 10.4.1.

11. Relations (II)

The purpose of this section is to check the relations Equation 1.2.8 with and Equation 1.2.9. Our strategy is the following. We first reduce the computation of the convolution product to the case of the graph of type using results in §8 and introducing modifications of quiver varieties and Hecke correspondences. Then we perform the computation using the explicit description of the equivariant -theory for quiver varieties for the graph of type .

In this section we fix a vertex .

11.1. Modifications of quiver varieties

We take a collection of vector spaces with . Let be the -component of , i.e.,

Let

This is a product of the quiver variety for the graph of type and the affine space:

where

Moreover, the variety is isomorphic to the cotangent bundle of the Grassmann manifold of -dimensional subspaces in the -dimensional space. (See Reference 44, Chap. 7 for details.) The isomorphism is given as follows: is the set of -orbits of , such that

(a)

,

(b)

is injective.

The action is given by . Then

defines a map , and the linear map

defines a cotangent vector at .

Let be as in Definition 2.3.1 and let be the set of stable points in . Although the stability condition 2.3.1 was defined only for , it can be defined for any . Let

We have a natural action of

on , and . We have the following relations between these varieties:

(a)

is an open subvariety of ,

(b)

is a nonsingular closed subvariety of (defined by the equation for ),

(c)

is a principal -bundle over .

The vector space defines vector bundles , , and . We denote all of them by for brevity, hoping that it causes no confusion.

11.2. Modifications of Hecke correspondences

Fix . Take collections of vector spaces , whose dimension vectors , satisfy . (For the proof of Theorem 9.4.1, it is enough to consider the case . But we study general for a later purpose.) These data will be fixed throughout this subsection, and we use the following notation:

Consider and . These varieties are products of quiver varieties for the graph of type and the affine space. We fix an isomorphism for . Then we have identifications and . We write them as and respectively for brevity. We write for and for . Let us define a subvariety as the product of the Hecke correspondence for the graph of type and the diagonal for the affine space. Namely

where and

Here is the generalization of the Hecke correspondence introduced in Equation 5.3.1. Since the graph is of type , it is isomorphic to the conormal bundle of

The quotient defines a vector bundle over of rank .

We have

The latter inclusion is obvious from the definition of stability, and the former one follows from the argument in Reference 45, Proof of 4.5. Let

For , the first factor satisfies if and only if the other factor also satisfies . This implies that

Let

The quotient defines vector bundles over , and . For brevity, all are simply denoted by .

Let us denote by the inclusion for . By Equation 11.2.2, the inclusion map induces the pull-back homomorphism with support

Similarly, we have

Lemma 11.2.3.

We have

More generally, if denotes a tensor product of exterior products of the bundle and its dual, we have

Proof.

The latter statement follows from the first statement and the formula Equation 6.4.1. Thus it is enough to check the first statement, and the first statement follows from the transversality of intersections Equation 11.2.2 in .

Let be the -part of the moment map . It induces a map for . Let us denote it by . Thus we have . Composing with the projection , we have a map for . We denote it also by for brevity. It is enough to show that the restriction of the differential to is surjective on .

We consider the homomorphisms , defined in Equation 2.9.1 where is replaced by (). We denote them by and respectively.

Take a point . Then

and there exists such that

The tangent space is isomorphic to the space of such that

modulo the image of

where

and we have used the identification for .

Now suppose that is orthogonal to the image of . Putting , we consider as an element of . Then

for any . Since the image of Equation 11.2.5 lies in the kernel of , the above equality holds for any satisfying Equation 11.2.4.

Taking from , we find

Next taking from the other component (data related to the vertex ), we get

Comparing -components, we find

Comparing -components, we have

If we define

Equation 11.2.6, Equation 11.2.7 and Equation 11.2.8 imply that is -invariant and contained in . Thus by the stability condition. In particular, we have . This means that is surjective.

Let be the projection (). Then we have

By these properties, we have homomorphisms

Lemma 11.2.10.

We have

More generally, if denotes a tensor product of exterior products of the bundle and its dual, we have

Proof.

The latter statement follows from the former one together with the projection formula Equation 6.5.1. Thus it is enough to prove the former statement.

By definition, consists of

such that there exists satisfying Equation 5.1.3. We fix representatives , . Then the above is uniquely determined. Recall that we have chosen the identification for over . Let us define by

We define a new datum

By definition, we have

Hence is contained in . Moreover, is independent of the choice of the representative . Thus we have defined a map by

which is the inverse of the restriction of . In particular, this implies

Since is a principal -bundle, we have

Thus we have proved the first equation. The second equation can be proved in a similar way.

11.3. Reduction to the rank case

First consider the relation Equation 1.2.8 for . Let , , , be dimension vectors such that

We want to compute and in the component

and then compare it with the right hand side of Equation 1.2.8 with in the same component.

Let , , , and be as in §11.1. Let , , , be the Hecke correspondence and its modifications introduced in §11.2. (We drop the superscript and write the dimension vector , .) Let be the exchange of factors as before. Let , , be subvarieties in , , defined in the same way as .

We have the following commutative diagram:

The horizontal arrows are convolution products relative to

(1)

, , ,

(2)

, , ,

(3)

, , ,

(4)

, , .

The vertical arrows between the first and the second rows are homomorphisms given in Proposition 8.3.5. The arrows between the second and the third are homomorphisms given in Proposition 8.2.3. By the properties Equation 11.2.2, Equation 11.2.9 and

(a)

,

(b)

the restriction of to is proper,

(c)

,

those homomorphisms can be defined. Finally the arrows between the third and the fourth are restrictions to open subvarieties.

The commutativity for the first and the second squares follows from Propositions 8.3.5 and 8.2.3 respectively. The last square is also commutative since is an open subvariety of and since we have Equation 11.2.1.

Recall that the modified Hecke correspondence in the last row is the product of the Hecke correspondence for type and the diagonal . Under the composite of vertical homomorphisms, , at the upper left are the images of the exterior products of the corresponding elements for type and at the lower left, except for the following two differences:

(a)

the groups acting on varieties are different,

(b)

the sign factors in Equation 9.3.2, which involve the orientation , are different.

For the quiver varieties of type , the group is

But, if we define a homomorphism by

we have an induced homomorphism in equivariant -groups: . (Here is as in Equation 2.7.1.) It is compatible with the convolution product, hence it is enough to check the relation in .

Furthermore, the sign factor cancels out in . Thus the above differences make no effect when we check the relation Equation 1.2.8.

By the commutativity of the diagram, is the image of the corresponding element in the lower right.

We have a similar commutative diagram to compute . Hence the commutator is the image of the corresponding commutator in the lower right. In the next section, we will check the relation Equation 1.2.8 for type . In particular, the commutator in the lower right is represented by tautological bundles, considered as an element of the -theory of the diagonal . Note that is mapped to by examples in §8, and that the tautological bundles on are restricted to tautological bundles on . Hence we have exactly the same relation Equation 1.2.8 for the general case.

Similarly, we can reduce the check of the relation Equation 1.2.9 to the case of type .

11.4. Rank case

In this subsection, we check the relation when the graph is of type . This calculation is essentially the same as the one by Vasserot Reference 58, but we reproduce it here for the convenience of the reader. (Remark that our -action is different from the one in Reference 58. The definition of , etc. is also different.) We drop the subscript as usual.

We prepare several notations. For , let

Let

For a partition of the set into subsets, let be the subgroups of consisting of permutations which preserve each subset. For a subgroup , let be the subring of consisting of elements which are fixed by the action of . If is another partition of , we define the symmetrizer by

where is the quotient field of . For each , let be the partition . If is a partition of into subsets and (resp. ), we define a new partition (resp. ) by

If is a partition and , we define

where , .

Let be the quiver variety for the graph of type with dimension vectors , . It is isomorphic to the cotangent bundle of the Grassmann variety of -dimensional subspaces of an -dimensional space. Let denote the Grassmann variety contained in as the -section. Let be the analogue of Steinberg’s variety as before. The following lemma is crucial.

Lemma 11.4.1 (Reference 58, Lemma 13, Reference 13, Claim 7.6.7).

The representation of

on by convolution is faithful.

Thus it is enough to check the relation in .

Let be as in Equation 5.3.1. It is the conormal bundle of

We denote the projections for by , , and the projections for by , . Note that both , are smooth and proper. Let , denote the projections , .

Lemma 11.4.2 (Reference 58, Corollary 4).

For (resp. ), we have

where (resp. ) is the relative tangent bundle along the fibers of (resp. ).

Proof.

As explained in Reference 58, Corollary 4, the result follows from Lemma 6.3.1. The factor is introduced to make the differential in the Koszul complex equivariant.

By the Thom isomorphism Reference 13, 5.4.17,

is an isomorphism. Moreover, we have the following explicit description of the -group of the Grassmann variety (cf. Reference 13, 6.1.6):

where acts as permutations of , …, and , …, . If denotes the tautological rank vector bundle over and denotes the quotient bundle , the isomorphism is given by

where denotes the th elementary symmetric polynomial.

The tautological vector bundle is isomorphic to , and is isomorphic to the trivial bundle . Let (resp. , ) be the complex Equation 2.9.1 (resp. Equation 9.3.1) over . In the description above, we have

We also have

where

acts as permutations of , …, , , …, , and , …, . The natural vector bundle is . The relative tangent bundles , are

Lemma 11.4.4 (Reference 58, Proposition 6).

(1) The pull-back homomorphisms

are identified with the natural homomorphisms

respectively.

(2) The push-forward homomorphisms

are identified with the natural homomorphisms

respectively. (The right hand sides are a priori in , but they are in fact in .)

Using the above lemmas, we can write the operators explicitly as:

for . Similarly,

for .

Let us compare with in the component

as follows:

Hence we have

The relation Equation 1.2.9 for , can be proved in the same way.

Let us compare and in the component

as follows:

Terms with cancel out for and . Thus

Let

Then we have

Applying the residue theorem to we get

where denotes the Laurent expansion of at and respectively. Since

by Equation 11.4.3, we have completed the proof of Theorem 9.4.1.

12. Integral structure

In this section, we compare with . In the case of the affine Hecke algebra, the equivariant -group of the Steinberg variety is isomorphic to the integral form of the affine Hecke algebra (see Reference 13, 7.2.5). We shall prove a weaker form of the corresponding result for quiver varieties in this section.

12.1. Rank case

We first consider the case when the graph is of type . We drop the subscript . We use the notation in §11.4. We also consider where is as in Equation 5.3.1 and is the exchange of factors. We identify its equivariant -group with as in §11.4. In particular, the vector bundle is identified with .

Lemma 12.1.1.

(1) Let be an increasing sequence of integers and let , …, be a sequence of positive integers such that . Let be the partition

Then for , we have

for some . Here is the Hall-Littlewood polynomial (see Reference 41, III(2.1)), and the summation runs over the set of unordered -tuples such that for .

(2) Let us consider a tensor product of exterior products of the bundle and its dual over , and denote by the corresponding element in the equivariant -group. Then for , we have the following formula:

where the summation runs over the set of unordered -tuples such that for .

Proof.

(1) Generalizing Equation 11.4.5, we have the following formula for :

where the summation runs over the set of ordered -tuples such that , for .

Choose so that

Consider the following term which appeared in the above formula:

By Reference 41, III(2.1) it is equal to

where is the Hall-Littlewood polynomial and

Thus we have the assertion.

(2) By Lemmas 11.4.2, 11.4.4 we have

12.2

Let be

(It seems reasonable to conjecture that is free over since it is true for type . But I do not know how to prove it in general.)

Theorem 12.2.1.

The homomorphism in Theorem 9.4.1 induces a homomorphism .

Remark 12.2.2.

The homomorphism is neither injective nor surjective. It is likely that there exists a surjective homomorphism from a modification of to for a suitable subset of , as in Reference 45, 9.5, 10.15.

Proof of Theorem 12.2.1.

It is enough to check that , , and the coefficients of are mapped to . For and the coefficients of , the assertion is clear from the definition.

For and , we can use a reduction to the rank case as in §11. Namely, it is enough to show the assertion when the graph is of type .

Now if the graph is of type , Lemma 12.1.1 together with Lemma 11.4.1 show that is represented by a certain line bundle over extended to by . We leave the proof for as an exercise. The only thing we need is to write down an analogue of Lemma 12.1.1 for . It is straightforward.

12.3. The module

By Theorems 12.2.1 and 7.3.5, is a -module. We show that it is an l-highest weight module in this subsection.

Lemma 12.3.1.

Let be as in Equation 5.3.1 and let denote the exchange of factors. Let be a tensor product of exterior products of the vector bundle and its dual over . Let us consider it as an element of . Then can be written as a linear combination (over ) of elements of the form

where is the diagonal in .

Proof.

As in §11, we may assume that the graph is of type . Now Lemma 12.1.1 together with the fact that Hall-Littlewood polynomials form a basis of symmetric polynomials implies the assertion.

Proposition 12.3.2.

Let be the class represented by the structure sheaf of . Then

Proof.

The following proof is an adaptation of the proof of Reference 45, 10.2, which was inspired by Reference 35, 3.6 in turn.

We need the following notation:

We prove by induction on the dimension vector . When , the result is trivial since . Consider and suppose that

Take . We want to show

We may assume that the support of is contained in an irreducible component of without loss of generality. In fact, suppose that such that is an irreducible component. Since is a closed subvariety of and since is a closed subvariety of , we have the diagram

where the first and the second row are exact by Equation 6.1.2. Thus there exists such that . Then , and therefore there exists such that . By induction on the number of irreducible components in the support, we may assume that the support of is contained in an irreducible component, which is denoted by .

Let us consider defined in Equation 2.9.3. If for all , must be by Lemma 2.9.4. We have nothing to prove in this case. Thus there exists such that . Set . By the descending induction on , we may assume that

Since is an open subvariety of , we have an exact sequence

by Equation 6.1.2. Consider . By Equation 12.3.4, it is enough to show that

Since , the support of is contained in . We have a map

which is the restriction of the map Equation 5.4.2. Recall that this map is a Grassmann bundle (see Proposition 5.4.3). Let us denote its tautological bundle by . Then can be written as a linear combination of elements of the form

where is a tensor product of exterior powers of the tautological bundle and . Since the homomorphism

is surjective by Equation 6.1.2, there exists such that .

Consider . By Proposition 5.4.3, it is isomorphic to and the map can be identified with the projection to the second factor. Moreover, the tautological bundle is identified with the restriction of the natural vector bundle . Hence we have

where is considered as an element of . By Lemma 12.3.1, can be written as a linear combination of elements

where is the projection. By Equation 12.3.3,

Hence . Thus we have shown Equation 12.3.5.

13. Standard modules

In this section, we start the study of the representation theory of using and the homomorphism in 12.2.1. We shall define certain modules called standard modules, and study their properties. Results in this section hold even if is a root of unity.

Note that is contained in the center of by

Hence a -module (over ), which is l-integrable as a -module, decomposes as , where is a homomorphism from to and is the corresponding simultaneous generalized eigenspace, i.e., some powers of the kernel of act as on . Such a homomorphism is given by the evaluation of the character at a semisimple element in . (This gives us a bijection between homomorphisms and semisimple elements.)

What is the meaning of the choice of when we consider as a -module? The role of is clear. It is a specialization , and we get -modules. It will become clear later that corresponds to the Drinfel’d polynomials by

13.1. Fixed data

Let be a semisimple element in and let be the Zariski closure of . Let be the homomorphism given by the evaluation at . Considering as an -module by this evaluation homomorphism, we denote it by . Via the homomorphism , we consider also as an -module. We consider as a -algebra, where is a -algebra as in §9.1.

Let , be the fixed point subvarieties of , respectively. Let us take a point which is regular, i.e., for some .

The data , will be fixed throughout this section.

13.2. Definition

As in Equation 2.3.5, let denote the inverse image of under the map . It is invariant under the -action. Let be . We set

as convention.

Let be

Let

By Theorem 7.3.5 together with Theorem 3.3.2, is a free -module. Thus the -module structure on descends to a -module structure. Hence is a -module via the composition of

We call the standard module.

It has a decomposition , and each summand is a weight space:

Thus has the weight decomposition as a -module.

In the remainder of this section, we study properties of . The first one is the following.

Lemma 13.2.4.

As a -module, is l-integrable.

Proof.

The assertion is proved exactly as in Reference 45, 9.3. Note that the regularity assumption of is not used here.

13.3. Highest weight vector

Recall that is an isomorphism on (Proposition 2.6.2). Under this isomorphism, we can consider as a point in . Then consists of the single point , thus we have a canonical generator of . We denote it by .

Since is fixed by , the fibers , of tautological bundles at are -modules. Then the restriction of the complex to can be considered as a complex of -modules. In particular, it defines an element in . Let us denote it by .

Let us spell out more explicitly. Since is fixed by , we have a homomorphism by §4.1. It is uniquely determined by up to the conjugacy. Then a virtual -module

can be considered as a virtual -module via . Its isomorphism class is independent of and coincides with . Note that the first and third terms in Equation 2.9.1 are absorbed in the term .

Proposition 13.3.1.

The standard module is an l-highest weight module with l-highest weight . Namely, the following hold:

(1) is a polynomial in of degree .

(2)

(3)

Proof.

(1) If we restrict the complex to , is surjective and is injective by Lemma 2.9.2. Thus is represented by a genuine -module, and is a polynomial in . The degree is equal to by the definition of .

(2) The first equation is the consequence of , which follows from Lemma 2.9.4. The remaining equations follow from the definition and Lemma 8.1.1.

(3) The assertion is proved exactly as in Proposition 12.3.2. Note that the assumption is used here in order to apply Lemma 2.9.4.

Remark 13.3.2.

is the Drinfel’d polynomial attached to the simple quotient of , which we will study later.

We give a proof of Proposition 1.2.16 as promised:

Proof of Proposition 1.2.16.

It is enough to show that there exists a simple l-integrable l-highest module with given Drinfel’d polynomials . We can construct it as the quotient of the standard module by the unique maximal proper submodule. (The uniqueness can be proved as in the case of Verma modules.) Here the parameter is chosen so that , i.e., is a normalization of the characteristic polynomial of .

13.4. Localization

Let denote the localization of with respect to .

Let denote the fixed point set of on , and let be the inclusion. Note that it induces an inclusion which we also denote by . By the concentration theorem Reference 53

is an isomorphism. Let

be the pull-back with support map. Then is given by multiplication by , where is the normal bundle of in . By Reference 13, 5.11.3, becomes invertible in the localized -group. Thus is an isomorphism on the localized -group. As in Reference 13, 5.11.10, we introduce a correction factor to :

Then is an algebra isomorphism with respect to the convolution.

Since acts trivially on , we have

Thus we have the evaluation map

by sending to .

By the bivariant Riemann-Roch theorem Reference 13, 5.11.11,

is an algebra homomorphism with respect to the convolution. Here is the local Chern character homomorphism with respect to and is the Todd genus of .

Composing Equation 13.2.2 with all these homomorphisms, we have a homomorphism

Note that the torsion part in Equation 13.2.2 disappears in the right hand side of Equation 13.4.1 after tensoring with .

We have similar -linear maps for :

where is an isomorphism by the concentration theorem Reference 53 and the invertibility of in the localized -homology group, is an isomorphism since acts trivially on , and is an isomorphism by Theorem 7.4.1 and Theorem 3.3.2. The composition is compatible with the -module structure, where is a -module via the convolution together with Equation 13.4.2.

Recall that we have the decomposition

where runs over the set of homomorphisms (with various )4.1). Let

Thus we have the canonical decomposition

Each summand in Equation 13.4.4 is an l-weight space with respect to the -action in the sense that operators act on as scalars plus nilpotent transformations. More precisely, we have

Proposition 13.4.5.

(1) Let be the tautological vector bundle over . Viewing as an element of , we consider it as an operator on . Then we have

where is the evaluation at of , considered as an -module via .

(2) Let us consider

as a virtual -module via . Then operators act on by

plus nilpotent transformations.

Proof.

(2) follows from (1). We show (1).

Note that is mapped to under , and is mapped to the fundamental class under . Combining with the projection formula Equation 6.5.1, we find that the operator is mapped to

under the homomorphism Equation 13.4.2. Thus as an operator on , it is equal to

where is the inclusion.

Now, on a connected space , any acts on as a scalar plus nilpotent operator, where the scalar is the -part of . In our situation, the -part of Equation 13.4.7 is given by . (Although we do not prove is connected, the -part is the same on any component.)

Furthermore, determines all eigenvalues of the operator acting on . Hence it determines the conjugacy class of the homomorphism . Thus the generalized eigenspace of with the eigenvalue coincides with .

13.5. Frenkel-Reshetikhin’s -character

In this subsection, we study Frenkel-Reshetikhin’s -character for the standard module . The result is a simple application of Proposition 13.4.5. Results in this subsection will not be used in the rest of the paper.

We assume is of type and is not a root of unity in this subsection.

Let us recall the definition of -character. It is a map from the Grothendieck group of finite dimensional -modules . As we shall see later in §14.3, standard modules ( is fixed, and are moving) give a basis of the Grothendieck group, thus it is enough to define the -character for standard modules . We decompose as

as in Equation 1.3.1. Moreover, by Proposition 13.4.5, have the form

where , are polynomials in with constant term . (Compare with Reference 18, Proposition 1. Note .) Suppose

Then the -character is defined by

where , are formal variables and takes its value in . ( should not be confused with .)

Let

Proposition 13.5.2 (cf. Conjecture 1 in Reference 18).

Let be a standard module with . Suppose that in Proposition 13.3.1 equals

for . Then the -character of has the following form:

where each is a product of with .

Proof.

By Proposition 13.4.5, is a generalized eigenspace for for a homomorphism . Thus it is enough to study the eigenvalue. We consider , as -modules via as before. Let , be weight spaces as in §4.1.

By the definition of and , we have

By Proposition 13.4.5 we have

where , are defined by Equation 13.5.1.

Let be the set of eigenvalues of on counted with multiplicities. Then we have

Thus we have

Note that the term for with has the contribution

and any other terms are monomials of which are not constant.

Moreover, we have by Lemma 4.1.4. This completes the proof.

14. Simple modules

The purpose of this section is to study simple modules of . Our discussion relies on Ginzburg’s classification of simple modules of the convolution algebra Reference 13, Chapter 9. (See also Reference 37.) He applied his classification to the affine Hecke algebra. However, unlike the case of the affine Hecke algebra, his classification does not directly imply a classification of simple modules of , and we need an extra argument. A difficulty lies in the fact that the homomorphism in Equation 13.4.2 is not necessarily an isomorphism. Our additional input is Proposition 13.3.1(3). In order to illustrate its usage, we first consider the special case when is generic in the first subsection. In this case, Ginzburg’s classification becomes trivial. Then we shall review Ginzburg’s classification in §14.2, and finally we shall study the general case in the last subsection.

We preserve the setup in §13.

14.1

Let us identify , with their image under Equation 13.4.2. Let denote the fundamental class of the diagonal of .

Lemma 14.1.1.

Let us consider

and let be the weight of determined by and as in §5.2. Then we have the following equality in :

Proof.

We have the following equality in :

where is (the restriction of) the natural line bundle over . The assertion follows immediately.

Theorem 14.1.2.

Suppose that is generic in the sense of Definition 4.2.1. (Hence, .) Then the standard module

is a simple -module. Its Drinfel’d polynomial is given by

where is the -component of . Moreover, is isomorphic to a tensor product of l-fundamental representations when is finite dimensional.

Proof.

Recall that we have a distinguished vector (we denote it by ) in the standard module 13.3). It has the properties listed in Proposition 13.3.1. In particular, it is the eigenvector for , and the eigenvalues are given in terms of therein. In the present setting, is equal to .

Let

We have . We want to show that any nonzero submodule of is itself. The weight space decomposition (as a -module) Equation 13.2.3 of induces that of . Since the set of weights of is bounded from with respect to the dominance order, there exists a maximal weight of . Then a vector in the corresponding weight space is killed by all by the maximality. Thus contains a nonzero vector . Hence it is enough to show that since we have already shown that in Proposition 13.3.1(3).

Let us consider the operator

where is the diagonal embedding. If we consider such operators for various , , they form a commuting family. Moreover, is invariant under them since we have the relation

where , are tautological bundles over , respectively. (Here .) Thus is a direct sum of generalized eigenspaces for . Let us take a direct summand . By Proposition 13.4.5(1), is contained in for some . If we can show , then we get since is an arbitrary direct summand.

Since is generic, we have . Hence

and is a nonsingular projective variety (having possibly infinitely many components). By the Poincaré duality, the intersection pairing

is nondegenerate.

Let denote the transpose of with respect to the pairing , namely

By the definition of the convolution, is equal to where is the map exchanging the first and second factors and is the induced homomorphism on .

Let us consider , where is as above and is any other homomorphism. It is just the projection of to the component . We have

for some , , and . These and come from asymmetry in the definition of , and in the homomorphism Equation 13.4.2. We do not give their explicit forms, although it is possible. What we need is for to be written by tensor powers of exterior products of for various , . Thus we can write

for some by Lemma 14.1.1. Therefore, for , we have

for any , , , . Here we have used , , . Since , we have one of the following:

(a)

for any ,

(b)

.

The first case is excluded by the nondegeneracy of . Thus we have .

Let us prove the last assertion. First consider the case for some . If is not a root of unity, is generic for the quiver variety . Hence the above shows that the standard module for is simple, and hence gives an l-fundamental representation.

Let us return to the case for general . Let , …, be eigenvalues of counted with multiplicities. By Proposition 1.2.19, it is enough to show that

where is the standard module for with . Since has no odd homology groups (Theorem 7.4.1), we have

where denotes the topological Euler number. By a property of the Euler number, we have

If we take a maximal torus of , the fixed point set is isomorphic to Again by a property of the Euler number, we have

Since we have

we get Equation 14.1.3.

14.2. Simple modules of the convolution algebra

We briefly recall Ginzburg’s classification of simple modules of the convolution algebra Reference 13, §8.6. (See also Reference 37.)

Let be a complex algebraic variety. We consider the derived category of complexes of sheaves with constructible cohomology sheaves, and denote it by . We use the notation in Reference 13. For example, we put

is an algebra by the Yoneda product. The Verdier duality operator is denoted by . Given graded vector spaces , , we write if there exists a linear isomorphism which does not necessarily preserve the gradings. We will also use the same notation to denote that two objects are quasi-isomorphic up to a shift in the derived category.

Let be a projective morphism between algebraic varieties , , and assume that is nonsingular. Then we are in the setting for the convolution in §8 with and , where

Since , we have the convolution product

Let be the algebra . Set . Then the convolution defines an -module structure on . More generally, if is a locally closed subset of , then has an -module structure via convolution.

By Reference 13, 8.6.7, we have an algebra isomorphism, which does not necessarily preserve gradings,

where is the constant sheaf on .

We apply the decomposition theorem Reference 6 to . There exists an isomorphism in :

where is the set of isomorphism classes of simple perverse sheaves on such that some shift is a direct summand of . We thus have an isomorphism

Set and

so that . By definition, under the multiplication of . By a property of perverse sheaves, we have for and . Hence,

In particular, the projection is an algebra homomorphism. Furthermore, is a semisimple algebra, and the kernel of the projection, i.e., , consists of nilpotent elements, thus it is precisely the radical of . In particular,

is a complete set of mutually nonisomorphic simple -modules.

For , let denote the inclusion. Then is an -module. More generally, if is a locally closed embedding, then the hyper-cohomology groups and are -modules. It is known (see Reference 13, 8.6.16, 8.6.35) that is isomorphic to as an -module.

For , we write instead of , and instead of . By applying to Equation 14.2.1, we get an isomorphism

Let

By definition, we have under the -module on . In particular, is an -submodule for each . Hence

is an -module, on which acts as . By definition,

where the -module structure on the right hand side is given by . Thus we have

Theorem 14.2.3.

In the Grothendieck group of -modules of finite dimension over , we have

where the -module structure on the right hand side is given by .

Proof.

Since is equal to in the Grothendieck group, the assertion follows from the discussion above.

14.3

In this subsection, we assume that the graph is of type , and is not a root of unity. We apply the results in the previous subsection to our quiver varieties.

Recall

where is the restriction of . Thus the results in the previous subsection are applicable to this setting. We have an algebra isomorphism

Let us denote this algebra by as in the previous subsection.

Since the graph is of type , we have by Proposition 2.6.3. Thus we have the stratification . (In fact, this holds under the same assumption as in Theorem 5.5.6: If we decompose as in Reference 45, 3.27, then the restriction of to for is zero by Proposition 4.2.2 with .) Since the restriction

is a locally trivial topological fibration by Theorem 3.3.2, all the complexes in the right hand side of Equation 14.2.1 (applied to , ) have locally constant cohomology sheaves along each stratum . Since is irreducible by Theorem 5.5.6, it implies that is the intersection cohomology complex associated with an irreducible local system on . Thus we have

for some finite dimensional vector space . Let . By a discussion in the previous subsection, is a complete set of mutually non-isomorphic simple -modules. Via the homomorphism Equation 13.4.2, is considered also as a -module.

Theorem 14.3.2.

Assume is not a root of unity.

(1) Simple perverse sheaves whose shift appears in a direct summand of are the intersection cohomology complexes associated with the constant local system on various .

(2) Let us denote the constant local system by for simplicity. Then is nonzero if and only if . Moreover, there is a bijection between the set and the set of l-weights of which are l-dominant.

(3) The simple -module is also simple as a -module, and its Drinfel’d polynomial is in Proposition 13.3.1 for .

(4) is the simple quotient of , where is a point in a stratum .

(5) Standard modules and are isomorphic as -modules if and only if and are contained in the same stratum.

Proof.

We use the transversal slice in §3.3. The idea to use transversal slices is taken from Reference 13, §8.5.

Choose and fix a point . Suppose that is contained in a stratum for some . We first show

Claim.

If denotes the constant local system on , the corresponding vector space is nonzero.

If we restrict to the component , then we have

where is a direct summand of . The summation runs over the set of pairs such that is contained in . (In fact, Equation 14.3.1 was obtained by applying the decomposition theorem to each component and taking the direct sum.) If we restrict Equation 14.3.3 to the open stratum of , the right hand side of Equation 14.3.3 becomes

where the summation runs over the set of isomorphism classes of irreducible local systems on . On the other hand, induces an isomorphism between and by Proposition 2.6.2. This means that the restriction of the left hand side of Equation 14.3.3 is the constant local system . Hence we have , and is nonzero. This is the end of the proof of the claim.

The claim implies the first assertion of (2). Let us prove the latter assertion of (2). Suppose . Then we have and by Proposition 4.1.2. By Proposition 13.4.5, the corresponding l-weight space is nonzero, where the l-weight is given by Equation 13.4.6. Furthermore, since can be represented by a genuine -module over a point in by Lemma 2.9.2, is a polynomial in . Thus is l-dominant.

Conversely suppose that we have the l-weight space with the l-weight Equation 13.4.6 nonzero. Since is not a root of unity, the -analogue of the Cartan matrix is invertible. Hence Equation 13.4.6 determines , and a homomorphism . Moreover, the l-weight space is precisely by Proposition 13.4.5(1). In particular, we have , and hence . Furthermore, if we decompose into as in §4.1, we have since the l-weight Equation 13.4.6 is l-dominant. By Corollary 5.5.5, is surjective for any , on a nonempty open subset of . By Lemma 2.9.4, is nonempty. This shows the latter half of (2).

Let denote the dimension vector corresponding to , i.e., . Take a transversal slice to at as in §3.3. Let be its intersection with . Since the transversal slice in §3.3 can be made -equivariant (Remark 3.3.3), it is a transversal slice to (at ) in . Let . Let , be the inclusions.

The stratification induces by restriction a stratification where . Any intersection complex restricts (up to shift) to the intersection complex by transversality. Here is the restriction of to . Taking of Equation 14.3.1, we get

Let be the inclusion. It induces two pull-back homomorphisms , , and there is a natural morphism for any . We apply these functors to both sides of Equation 14.3.4 and take cohomology groups. By a property of intersection cohomology sheaves (see Reference 13, 8.5.3), the homomorphism

is zero unless (or equivalently ), in which case it is a quasi-isomorphism. Thus

where the summation runs over isomorphism classes of irreducible local systems on , and is the fiber of the local system at . Moreover, Equation 14.3.5 is a homomorphism of -modules, and Equation 14.3.6 is an isomorphism of -modules, where the module structure on the right hand side is given by .

On the other hand, we have

As shown in Equation 13.4.3, the right hand side is isomorphic to the standard module . Thus the left hand side of Equation 14.3.6 is a quotient of , and it is indecomposable by Proposition 13.3.1(3). Thus the right hand side of Equation 14.3.6 consists of at most a single direct summand. Since we have already shown that in the claim, we get if is a nonconstant irreducible local system. Since was an arbitrary point, we have the statement (1).

Let us prove (3). For the proof, we need a further study of Equation 14.3.6. By the above discussion, we have

By the base change theorem, we have where is the restriction of to . Further, we have since is a nonsingular submanifold of . Applying the Verdier duality, we have

Hence Equation 14.3.7 becomes

where is the dual space of as a complex vector space. Let us introduce an -module on by

where denotes the dual pairing, is the exchange of two factors of , and is the induced homomorphism on . Then Equation 14.3.8 is compatible with -module structures (cf. Reference 13, paragraphs preceding 8.6.25).

The decomposition Equation 13.4.4 induces a similar one for :

The homomorphism respects the decomposition, and induces a decomposition on Equation 14.3.8.

Recall that we have the distinguished vector in . The component of is -dimensional space . (See §13.3.) By the above discussion, is not annihilated by the above homomorphism . Thus we may consider also as an element of .

We want to show that any nonzero -submodule of is itself. Our strategy is the same as in the proof of Theorem 14.1.2. Since we already show that is a quotient of , Proposition 13.3.1(3) implies . Thus it is enough to show that contains . To show this, consider

By the argument as in the proof of Theorem 14.1.2, contains a nonzero vector in . Hence it is enough to show that .

As in the proof of Theorem 14.1.2 above, is a direct sum of generalized eigenspaces for . Let us choose and fix a direct summand contained in . Then satisfies

for any , . Since by Proposition 13.3.1(3), the above equation implies that . Thus we get as desired.

We have shown the statement (4) during the above discussion.

Let us prove (5). Since is a locally trivial topological fibration on each stratum , and are isomorphic if both and are contained in . Conversely, if and are isomorphic as -modules, the corresponding l-highest weights and are equal. Since determines the homomorphism as in the proof of (2), and are in the same stratum.

Remark 14.3.9.

The assumption that is not a root of unity is used to apply Theorem 5.5.6 and to have the invertibility of the -analogue of the Cartan matrix. It seems likely that Theorem 5.5.6 holds even if is a root of unity. The latter condition was used to parametrize the index set of (i.e., Theorem 14.3.2(2)). But one should have a similar parametrization if one replaces a notion of l-weights in a suitable way. Thus Theorem 14.3.2 should hold even if is a root of unity, if one replaces the statement (2).

Let be the set of l-weights of , which are l-dominant. Since the index set of the stratum coincides with , we may write as , when corresponds to . The standard module depends only on the stratum containing , so we may also write as . We have an analogue of the Kazhdan-Lusztig multiplicity formula:

Theorem 14.3.10.

Assume is not a root of unity.

For , let denote the inclusion. Let be the l-weight corresponding to the stratum containing . In the Grothendieck group of finite dimensional -modules, we have

where is the stratum corresponding to , and is the intersection cohomology complex attached to and the constant local system . Here the -module structure on the right hand side is given by .

This follows from Theorem 14.3.2 and a result in the previous subsection.

Remark 14.3.11.

By Reference 13, 8.7.8 and Theorem 7.4.1 with Theorem 3.3.2, the cohomology group vanishes for all odd , where is the dimension of .

15. The -module structure

In this section, we assume the graph is of type . The result of this section holds even if is a root of unity, if we replace the simple module by the corresponding Weyl module (see Reference 10, 11.2 for the definition).

15.1

For a given , let be the finite set consisting of all such that is dominant and the weight space with weight is nonzero in the simple highest weight -module .

Let be the union of all for various . It is the set consisting of all such that is dominant.

Since the graph is of type , we have . Since is isomorphic to an open subvariety of , is irreducible if is connected. Although we do not know whether is connected or not (see §7.5), we consider the intersection cohomology complex attached to and the constant local system . It may not be a simple perverse sheaf if is not connected.

We prove the following in this section:

Theorem 15.1.1.

As a -module, we have the following decomposition:

where is the inclusion, and acts trivially on the factor .

Remark 15.1.2.

By Reference 13, 8.7.8 and Theorem 7.3.5 with Theorem 3.3.2, the cohomology group vanishes.

15.2. Reduction to

Suppose that is contained in a stratum . Take a representative of and define as in Equation 4.1.1. Here is fixed and we do not consider . We choose , , so that , , , where . Let for . Then we have

from Equation 4.1.1. If denotes the stabilizer of in , the above equation means that .

Let us consider a -module

parametrized by , where is an -module given by the evaluation at as in §13. When , we can replace by by Theorem 7.3.5 and Theorem 3.3.2, hence the module coincides with . Moreover, it depends continuously on , also by Theorem 7.3.5.

Let us consider as a -module by the restriction. Since finite dimensional -modules are classified by discrete data (highest weights), it is independent of . (Simple modules of depend continuously on .) Thus it is enough to decompose when , i.e., , . By Theorem 7.3.5 and Theorem 3.3.2, is specialized to at , . Thus our task now becomes the decomposition of into simple -modules.

15.3

When is pure dimensional, we denote by the top degree part of , that is, the subspace spanned by the fundamental classes of irreducible components of . Suppose that has several connected components , , … such that each is pure dimensional, but may change for different . Then we define as . Note that the degree may differ for different since the dimensions are changing.

By Reference 45, 9.4, there is a homomorphism

In fact, it is the restriction of the homomorphism in Equation 13.4.2 for , , composed with the projection

For each , we take a point . (By Lemma 2.9.4(2), is dominant if is nonempty.) By Reference 45, 10.2, is the simple highest weight module via this homomorphism. (In fact, we have already proved a similar result, i.e., Proposition 13.3.1.)

Proposition 15.3.1.

Consider the map . Then , as a map into , is semi-small and all strata are relevant, namely

where is the codimension in .

Proof.

See Reference 44, 6.11 and Reference 45, 10.11.

Proposition 15.3.2.

We have

where is taken from . (By Theorem 3.3.2 is independent of the choice of .)

Proof.

By the decomposition theorem for a semi-small map Reference 13, 8.9.3, the left hand side of Equation 15.3.3 decomposes as

where is a component of and is the intersection complex associated with an irreducible local system on . Moreover, by Reference 13, 8.9.9, we have

where runs over the set of irreducible local systems on the component of containing . But as argued in the proof of Theorem 14.3.2, the indecomposability of implies that no intersection complex associated with a nontrivial local system appears in the summand. Moreover the left hand side of Equation 15.3.4 is independent of the choice of the component by Theorem 3.3.2. Thus we can combine the summation over together as

Our remaining task is to identify the index set of . The fundamental class is nonzero if is nonempty. Thus is nonempty if and only if

By Reference 45, 10.2 and the construction, is isomorphic to the weight space of weight in . Thus it is nonzero if and only if .

Take and consider the inclusion . Applying to Equation 15.3.3 and then summing with respect to , we get

By the convolution product, is a module of . By Reference 13, §8.9, the decomposition Equation 15.3.5 is compatible with the module structure, where acts on by . This completes the proof of Theorem 15.1.1.

Added in proof

Crawley-Boevey recently proved that is connected (see §7.5) in “Geometry of the moment map for representations of quivers”, to appear in Compositio Math.

Table of Contents

  1. Abstract
  2. Introduction
  3. 1. Quantum affine algebra
    1. 1.1. Quantized universal enveloping algebra
    2. 1.2. Quantum affine algebra
    3. Proposition 1.2.16.
    4. Proposition 1.2.19 (10, 12.2.6,12.2.8).
    5. 1.3. An l-weight space decomposition
    6. Definition 1.3.2.
  4. 2. Quiver variety
    1. 2.1. Notation
    2. Convention 2.1.4.
    3. 2.2. Two quotients and
    4. 2.3. Stability condition
    5. Definition 2.3.1.
    6. Proposition 2.3.2.
    7. Notation 2.3.3.
    8. Definition 2.3.6.
    9. Proposition 2.3.7.
    10. Proposition 2.3.8.
    11. 2.4. Hyper-Kähler structure
    12. Proposition 2.4.1.
    13. 2.5
    14. Lemma 2.5.3.
    15. 2.6. Definition of
    16. Proposition 2.6.2.
    17. Proposition 2.6.3.
    18. Definition 2.6.4.
    19. 2.7. -action
    20. 2.8. Notation for -action
    21. 2.9. Tautological bundles
    22. Lemma 2.9.2.
    23. Lemma 2.9.4.
  5. 3. Stratification of
    1. 3.1. Stratification
    2. Definition 3.1.1 (cf. Sjamaar-Lerman 50).
    3. 3.2. Local normal form of the moment map
    4. Lemma 3.2.1.
    5. Proposition 3.2.2.
    6. 3.3. Slice
    7. Theorem 3.3.2.
  6. 4. Fixed point subvariety
    1. 4.1. A homomorphism attached to a component of
    2. Proposition 4.1.2.
    3. Lemma 4.1.4.
    4. 4.2. A sufficient condition for
    5. Definition 4.2.1.
    6. Proposition 4.2.2.
  7. 5. Hecke correspondence and induction of quiver varieties
    1. 5.1. Hecke correspondence
    2. Definition 5.1.4.
    3. 5.2. Hecke correspondence and fixed point subvariety
    4. 5.3
    5. 5.4. Induction
    6. Proposition 5.4.3.
    7. 5.5. Induction for fixed point subvarieties
    8. Proposition 5.5.2.
    9. Corollary 5.5.5.
    10. Theorem 5.5.6.
  8. 6. Equivariant -theory
    1. 6.1. Definitions
    2. 6.2. Operations on -theory of vector bundles
    3. 6.3. Tor-product
    4. Lemma 6.3.1 (13, 5.4.10, 58, Lemma 1).
    5. 6.4. Pull-back with support
    6. 6.5. Push-forward
    7. 6.6. Chow group and homology group
  9. 7. Freeness
    1. 7.1. Properties , ,
    2. Lemma 7.1.2.
    3. Lemma 7.1.3.
    4. Lemma 7.1.4.
    5. Lemma 7.1.5.
    6. 7.2. Decomposition of the diagonal
    7. Proposition 7.2.1 (cf. 16, 13, 5.6.1).
    8. 7.3. Diagonal of the quiver variety
    9. Lemma 7.3.3.
    10. Proposition 7.3.4.
    11. Theorem 7.3.5.
    12. 7.4. Fixed point subvariety
    13. Theorem 7.4.1.
    14. 7.5. Connectedness of
  10. 8. Convolution
    1. 8.1
    2. Lemma 8.1.1.
    3. 8.2. Restriction of the convolution to submanifolds
    4. Proposition 8.2.3.
    5. Example 8.2.5.
    6. 8.3. Convolution and principal bundles
    7. Proposition 8.3.5.
    8. Example 8.3.7.
  11. 9. A homomorphism
    1. 9.1
    2. 9.2
    3. 9.3
    4. 9.4
    5. Theorem 9.4.1.
  12. 10. Relations (I)
    1. 10.1. Relation 1.2.7
    2. 10.2. Relation 1.2.8 for
    3. Lemma 10.2.1.
    4. 10.3. Relation 1.2.10
    5. Lemma 10.3.1.
    6. Lemma 10.3.9.
    7. 10.4. Relation 1.2.11
  13. 11. Relations (II)
    1. 11.1. Modifications of quiver varieties
    2. 11.2. Modifications of Hecke correspondences
    3. Lemma 11.2.3.
    4. Lemma 11.2.10.
    5. 11.3. Reduction to the rank case
    6. 11.4. Rank case
    7. Lemma 11.4.1 (58, Lemma 13, 13, Claim 7.6.7).
    8. Lemma 11.4.2 (58, Corollary 4).
    9. Lemma 11.4.4 (58, Proposition 6).
  14. 12. Integral structure
    1. 12.1. Rank case
    2. Lemma 12.1.1.
    3. 12.2
    4. Theorem 12.2.1.
    5. 12.3. The module
    6. Lemma 12.3.1.
    7. Proposition 12.3.2.
  15. 13. Standard modules
    1. 13.1. Fixed data
    2. 13.2. Definition
    3. Lemma 13.2.4.
    4. 13.3. Highest weight vector
    5. Proposition 13.3.1.
    6. 13.4. Localization
    7. Proposition 13.4.5.
    8. 13.5. Frenkel-Reshetikhin’s -character
    9. Proposition 13.5.2 (cf. Conjecture 1 in 18).
  16. 14. Simple modules
    1. 14.1
    2. Lemma 14.1.1.
    3. Theorem 14.1.2.
    4. 14.2. Simple modules of the convolution algebra
    5. Theorem 14.2.3.
    6. 14.3
    7. Theorem 14.3.2.
    8. Theorem 14.3.10.
  17. 15. The -module structure
    1. 15.1
    2. Theorem 15.1.1.
    3. 15.2. Reduction to
    4. 15.3
    5. Proposition 15.3.1.
    6. Proposition 15.3.2.
  18. Added in proof

Mathematical Fragments

Equation (1.1.6)
Equations (1.2.1), (1.2.2), (1.2.3), (1.2.4), (1.2.5), (1.2.6), (1.2.7), (1.2.8), (1.2.9), (1.2.10)
Equation (1.2.11)
Equation (1.2.13)
Proposition 1.2.16.

Assume that is symmetric. The simple l-highest weight -module with l-highest weight is l-integrable if and only if is dominant and there exist polynomials for with such that

where , and denotes the expansion at and respectively.

Proposition 1.2.19 (Reference 10, 12.2.6,12.2.8).

Suppose is finite dimensional.

(1) If and are simple l-highest weight -modules with Drinfel’d polynomials , such that is simple, then its Drinfel’d polynomial is given by

(2) Every simple l-highest weight -module is a subquotient of a tensor product of l-fundamental representations.

Equation (1.3.1)
Equation (2.1.1)
Equation (2.1.2)
Convention 2.1.4.

When we relate the quiver varieties to the quantum affine algebra, the dimension vectors will be mapped into the weight lattice in the following way:

where (resp. ) is the th component of (resp. ). Since and are both linearly independent, these maps are injective. We consider and as elements of the weight lattice in this way hereafter.

Equation (2.1.6)
Equation (2.1.7)
Equation (2.1.8)
Definition 2.3.1.

A point is said to be stable if the following condition holds:

if a collection of subspaces of is -invariant and contained in , then .

Let us denote by the set of stable points.

Proposition 2.3.2.

(1) A point is stable if and only if the closure of does not intersect with the zero section of for .

(2) If is stable, then the differential is surjective. In particular, is a nonsingular variety.

(3) If is stable, then in Equation 2.1.8 is injective.

(4) The quotient has a structure of nonsingular quasi-projective variety of dimension , and is a principal -bundle over .

(5) The tangent space of at the orbit is isomorphic to the middle cohomology group of Equation 2.1.8.

(6) The variety is isomorphic to .

(7) has a holomorphic symplectic structure as a symplectic quotient.

Equation (2.3.4)
Equation (2.3.5)
Definition 2.3.6.

Suppose that and a -invariant increasing filtration

with are given. Then set and . Let denote the endomorphism which induces on . For , let be such that its composition with the inclusion is , and let be the restriction of to . For , set and . Let be the direct sum of considered as data on .

Proposition 2.3.7.

Suppose . Then there exist a representative of and a -invariant increasing filtration as in Definition 2.3.6 such that is a representative of on .

Proposition 2.3.8.

is a Lagrangian subvariety which is homotopic to .

Proposition 2.4.1.

(1) A -orbit in intersects with if and only if it is closed. The map

is a homeomorphism.

(2) Choose a parameter so that . Then a -orbit in intersects with if and only if it is stable. The map

is a homeomorphism.

Equation (2.5.1)
Equation (2.5.2)
Lemma 2.5.3.

The morphism Equation 2.5.2 is injective.

Equation (2.5.4)
Proposition 2.6.2.

If , then it is stable. Moreover, induces an isomorphism .

Proposition 2.6.3.

If the graph is of type , then

where the summation runs over the set of such that .

Definition 2.6.4.

We say a point is regular if it is contained in for some . The above proposition says that all points are regular if the graph is of type . But this is not true in general (see Reference 45, 10.10).

Equation (2.7.1)
Equation (2.7.2)
Equation (2.8.1)
Equation (2.8.2)
Equation (2.9.1)
Lemma 2.9.2.

Fix a point and consider as a complex of vector spaces. Then is injective.

Equation (2.9.3)
Lemma 2.9.4.

(1) Take and fix a point . Let be as in Equation 2.9.1. If , then we have

Moreover, the converse holds if we assume is regular in the sense of Definition 2.6.4. Namely under this assumption, if and only if 2.9.5 holds.

(2) If , then is dominant.

Definition 3.1.1 (cf. Sjamaar-Lerman Reference 50).

For a subgroup of denote by the set of all points in whose stabilizer is conjugate to . A point is said to be of -orbit type if its representative is in . The set of all points of orbit type is denoted by .

Lemma 3.2.1.

A neighborhood of (in ) is -equivalently symplectomorphic to a neighborhood of embedded as the zero section of with the -moment map given by the formula

(Here ‘symplectomorphic’ means that there exists a biholomorphism intertwining symplectic structures.)

Proposition 3.2.2.

There exist a neighborhood (resp. ) of (resp. ) and biholomorphic maps , such that the following diagram commutes:

In particular, is biholomorphic to .

Furthermore, under , a stratum of is mapped to a stratum , which is defined as in Definition 3.1.1. (If intersects with , then is conjugate to a subgroup of .)

Theorem 3.3.2.

Suppose that as above. Then there exist neighborhoods , , of , , respectively, and biholomorphic maps , such that the following diagram commutes:

In particular, is biholomorphic to .

Furthermore, a stratum of is mapped to a product of and a stratum of .

Remark 3.3.3.

Suppose that is a subgroup of fixing . Since the -action commutes with the -action, has an -action. The above construction can be made -equivariant. In particular, the diagram in Theorem 3.3.2 can be restricted to a diagram for -fixed point sets.

Equation (4.1.1)
Proposition 4.1.2.

is homotopic to .

Equation (4.1.3)
Lemma 4.1.4.

If , then for some and .

Equation (4.1.5)
Equation (4.1.6)
Definition 4.2.1.

We say is generic if for any . (This condition depends on .)

Proposition 4.2.2.

Assume that there is at most one edge joining two vertices of , and that

for any pair of eigenvalues of . (The condition for the special case implies that is not a root of unity.) Then is generic.

Equation (5.1.1)
Equation (5.1.3)
Equation (5.3.1)
Equation (5.4.2)
Proposition 5.4.3.

Let be the Grassmann bundle of -planes in the vector bundle obtained by restricting to . Then we have the following diagram:

where is the natural projection, and and are restrictions of the projections to the first and second factors. The kernel of the natural surjective homomorphism is isomorphic to the tautological vector bundle of the Grassmann bundle of the first row, and also to the the restriction of the vector bundle over in the third row.

Proposition 5.5.2.

Suppose that . We have the following diagram:

where is the natural projection. For each , the kernel of the natural surjective homomorphism is isomorphic to the tautological vector bundle of the Grassmann bundle. Moreover, we have

(Here of a complex means the alternating sum of dimensions of cohomology groups.)

Corollary 5.5.5.

Suppose . On a nonempty open subset , we have

for each . Also, the complement is a lower dimensional subvariety of .

Theorem 5.5.6.

Assume that is not a root of unity and there is at most one edge joining two vertices of . Then is connected if it is a nonempty set.

Equation (5.5.7)
Equation (6.1.1)
Equation (6.1.2)
Equation (6.1.3)
Lemma 6.3.1 (Reference 13, 5.4.10, Reference 58, Lemma 1).

Let , be nonsingular -subvarieties with conormal bundles , . Suppose that is nonsingular and , where means the restriction to . Then for any , , we have

where .

Equation (6.4.1)
Equation (6.5.1)
Equations (6.6.1), (6.6.2)
Equation (7.1.1)
Lemma 7.1.2.

If has an -partition into pieces which have property , then has property .

Lemma 7.1.3.

Suppose that an algebraic variety has an action of a linear algebraic group . If has an -partition into -invariant locally closed subvarieties which have property , then has property .

Lemma 7.1.4.

Let be a -equivariant fiber bundle with affine spaces as fibers. Suppose that is locally a trivial -equivariant vector bundle, i.e., a product of base and a vector space with a linear -action. If has property (resp. ), then also has property (resp. ).

Lemma 7.1.5.

Let be a nonsingular quasi-projective variety with -action with a Kähler metric such that

(a)

is complete,

(b)

is invariant under the maximal compact subgroup of ,

(c)

there exists a moment map associated with the Kähler metric and the -action (the maximal compact subgroup of the second factor), and it is proper.

Let

If the fixed point set has property (resp. ), then both and have property (resp. ).

Furthermore, the bilinear pairing

is nondegenerate if has property . A similar intersection pairing between and is nondegenerate if has property . Here is the canonical map from to the point.

Proposition 7.2.1 (cf. Reference 16, Reference 13, 5.6.1).

Let be a nonsingular projective variety.

(1) Let be the structure sheaf of the diagonal and the corresponding element in . Assume that

holds for some , . Then has property .

(2) Let be a linear algebraic group. Suppose that has -action and that 7.2.2 holds in for some , . Then has property .

(3) Let be the class of the diagonal in . Assume that

holds for some , . Then has property .

Equation (7.2.4)
Equation (7.3.1)
Proposition 7.3.4.

has properties and . Moreover, is connected.

Theorem 7.3.5.

and have properties and . Moreover, the bilinear pairing

is nondegenerate. A similar pairing between and is also nondegenerate. Here is the canonical map from to the point.

Theorem 7.4.1.

and have properties and . Moreover, the bilinear pairing

is nondegenerate. A similar pairing between and is also nondegenerate. Here is the canonical map from to the point.

Equation (7.5.1)
Lemma 8.1.1.

In the above setting, we further assume that and , where is the diagonal embedding . Then we have

for a vector bundle over , where is the projection, and in the right hand side is the tensor product Equation 6.1.1 between and .

Equation (8.2.1)
Equation (8.2.2)
Proposition 8.2.3.

For , we have

Namely, the following diagram commutes:

Equation (8.2.6)
Equation (8.2.7)
Equation (8.2.8)
Equation (8.3.1)
Equation (8.3.2)
Equation (8.3.3)
Equation (8.3.4)
Proposition 8.3.5.

In the above setup, we have

Namely the following diagram commutes:

where the left vertical arrow is

and the right vertical arrow is

Equation (8.3.8)
Equation (8.3.9)
Equation (8.3.10)
Equation (9.2.1)
Equation (9.3.1)
Equation (9.3.2)
Theorem 9.4.1.

The assignments Equation 9.2.1, Equation 9.3.2 define a homomorphism

of -algebras.

Equation (10.1.1)
Lemma 10.2.1.

The above two intersections are transversal, and there is a -equivariant isomorphism between them such that

(a)

it is the identity operators on the factor and ,

(b)

it induces isomorphisms and .

Equation (10.2.2)
Equation (10.2.3)
Equation (10.3.2)
Equation (10.3.3)
Equation (10.3.4)
Equation (10.3.7)
Equation (10.3.8)
Lemma 10.3.9.

The section (resp. ) is transversal to the zero section (if it vanishes somewhere).

Equation (10.4.1)
Equation (10.4.2)
Equation (10.4.4)
Equation (11.2.1)
Equation (11.2.2)
Lemma 11.2.3.

We have

More generally, if denotes a tensor product of exterior products of the bundle and its dual, we have

Equation (11.2.4)
Equation (11.2.5)
Equation (11.2.6)
Equation (11.2.7)
Equation (11.2.8)
Equation (11.2.9)
Lemma 11.4.1 (Reference 58, Lemma 13, Reference 13, Claim 7.6.7).

The representation of

on by convolution is faithful.

Lemma 11.4.2 (Reference 58, Corollary 4).

For (resp. ), we have

where (resp. ) is the relative tangent bundle along the fibers of (resp. ).

Equation (11.4.3)
Lemma 11.4.4 (Reference 58, Proposition 6).

(1) The pull-back homomorphisms

are identified with the natural homomorphisms

respectively.

(2) The push-forward homomorphisms

are identified with the natural homomorphisms

respectively. (The right hand sides are a priori in , but they are in fact in .)

Equation (11.4.5)
Lemma 12.1.1.

(1) Let be an increasing sequence of integers and let , …, be a sequence of positive integers such that . Let be the partition

Then for , we have

for some . Here is the Hall-Littlewood polynomial (see Reference 41, III(2.1)), and the summation runs over the set of unordered -tuples such that for .

(2) Let us consider a tensor product of exterior products of the bundle and its dual over , and denote by the corresponding element in the equivariant -group. Then for , we have the following formula:

where the summation runs over the set of unordered -tuples such that for .

Theorem 12.2.1.

The homomorphism in Theorem 9.4.1 induces a homomorphism .

Lemma 12.3.1.

Let be as in Equation 5.3.1 and let denote the exchange of factors. Let be a tensor product of exterior products of the vector bundle and its dual over . Let us consider it as an element of . Then can be written as a linear combination (over ) of elements of the form

where is the diagonal in .

Proposition 12.3.2.

Let be the class represented by the structure sheaf of . Then

Equation (12.3.3)
Equation (12.3.4)
Equation (12.3.5)
Equation (13.2.2)
Equation (13.2.3)
Lemma 13.2.4.

As a -module, is l-integrable.

Proposition 13.3.1.

The standard module is an l-highest weight module with l-highest weight . Namely, the following hold:

(1) is a polynomial in of degree .

(2)

(3)

Equation (13.4.1)
Equation (13.4.2)
Equation (13.4.3)
Equation (13.4.4)
Proposition 13.4.5.

(1) Let be the tautological vector bundle over . Viewing as an element of , we consider it as an operator on . Then we have

where is the evaluation at of , considered as an -module via .

(2) Let us consider

as a virtual -module via . Then operators act on by

plus nilpotent transformations.

Equation (13.4.7)
Equation (13.5.1)
Proposition 13.5.2 (cf. Conjecture 1 in Reference 18).

Let be a standard module with . Suppose that in Proposition 13.3.1 equals

for . Then the -character of has the following form:

where each is a product of with .

Lemma 14.1.1.

Let us consider

and let be the weight of determined by and as in §5.2. Then we have the following equality in :

Theorem 14.1.2.

Suppose that is generic in the sense of Definition 4.2.1. (Hence, .) Then the standard module

is a simple -module. Its Drinfel’d polynomial is given by

where is the -component of . Moreover, is isomorphic to a tensor product of l-fundamental representations when is finite dimensional.

Equation (14.1.3)
Equation (14.2.1)
Equation (14.3.1)
Theorem 14.3.2.

Assume is not a root of unity.

(1) Simple perverse sheaves whose shift appears in a direct summand of are the intersection cohomology complexes associated with the constant local system on various .

(2) Let us denote the constant local system by for simplicity. Then is nonzero if and only if . Moreover, there is a bijection between the set and the set of l-weights of which are l-dominant.

(3) The simple -module is also simple as a -module, and its Drinfel’d polynomial is in Proposition 13.3.1 for .

(4) is the simple quotient of , where is a point in a stratum .

(5) Standard modules and are isomorphic as -modules if and only if and are contained in the same stratum.

Equation (14.3.3)
Equation (14.3.4)
Equation (14.3.5)
Equation (14.3.6)
Equation (14.3.7)
Equation (14.3.8)
Remark 14.3.9.

The assumption that is not a root of unity is used to apply Theorem 5.5.6 and to have the invertibility of the -analogue of the Cartan matrix. It seems likely that Theorem 5.5.6 holds even if is a root of unity. The latter condition was used to parametrize the index set of (i.e., Theorem 14.3.2(2)). But one should have a similar parametrization if one replaces a notion of l-weights in a suitable way. Thus Theorem 14.3.2 should hold even if is a root of unity, if one replaces the statement (2).

Theorem 15.1.1.

As a -module, we have the following decomposition:

where is the inclusion, and acts trivially on the factor .

Proposition 15.3.2.

We have

where is taken from . (By Theorem 3.3.2 is independent of the choice of .)

Equation (15.3.4)
Equation (15.3.5)

References

Reference [1]
T. Akasaka and M. Kashiwara, Finite-dimensional representations of quantum affine algebras, Publ. RIMS 33 (1997), 839–867. MR 99d:17017
Reference [2]
M.F. Atiyah, Convexity and commuting hamiltonians, Bull. London Math. Soc. 14 (1982), 1–15. MR 83e:53037
Reference [3]
M.F. Atiyah and R. Bott, The Yang-Mills equations over Riemann surfaces, Phil. Trans. Roy. Soc. London A 308 (1982), 524–615. MR 85k:14006
Reference [4]
P. Baum, W. Fulton and R. MacPherson, Riemann-Roch and topological -theory for singular varieties, Acta. Math. 143 (1979), 155–192. MR 82c:14021
Reference [5]
J. Beck, Braid group action and quantum affine algebras, Comm. Math. Phys. 165 (1994), 555–568. MR 95i:17011
Reference [6]
A. Beilinson, J. Bernstein and P. Deligne, Faisceaux pervers, Astérisque 100 (1982). MR 86g:32015
Reference [7]
A. Bialynicki-Birula, Some theorems on actions of algebraic groups, Ann. of Math. 98 (1973), 480–497. MR 51:3186
Reference [8]
J. Carrell and A. Sommese, -actions, Math. Scand. 43 (1978/79), 49–59. MR 80h:32053
Reference [9]
V. Chari and A. Pressley, Fundamental representations of Yangians and singularities of -matrices, J. reine angew Math. 417 (1991), 87–128. MR 92h:17010
Reference [10]
—, A guide to quantum groups, Cambridge University Press, Cambridge, 1994. MR 95j:17010
Reference [11]
—, Quantum affine algebras and their representations in “Representation of Groups”, CMS Conf. Proc., 16, AMS, 1995, 59–78. MR 96j:17009
Reference [12]
—, Quantum affine algebras at roots of unity, Representation Theory 1 (1997), 280–328. MR 98e:17018
Reference [13]
N. Chriss and V. Ginzburg, Representation theory and complex geometry, Progress in Math. Birkhäuser, 1997. MR 98i:22021
Reference [14]
C. De Concini, G. Lusztig and C. Procesi, Homology of the zero-set of a nilpotent vector field on a flag manifold, J. Amer. Math. Soc. 1 (1988), 15–34. MR 89f:14052
Reference [15]
V.G. Drinfel’d, A new realization of Yangians and quantized affine algebras, Soviet Math. Dokl. 32 (1988), 212–216. MR 88j:17020
Reference [16]
G. Ellingsrud and S.A. Strømme, Towards the Chow ring of the Hilbert scheme of , J. reine angew. Math. 441 (1993), 33–44. MR 94i:14004
Reference [17]
E. Frenkel and E. Mukhin, Combinatorics of -characters of finite-dimensional representations of quantum affine algebras, preprint, math.QA/9911112.
Reference [18]
E. Frenkel and N. Reshetikhin, The -characters of representations of quantum affine algebras and deformations of -algebras, preprint, math.QA/9810055.
Reference [19]
W. Fulton, Intersection Theory, A Series of Modern Surveys in Math. 2, Springer-Verlag, 1984. MR 85k:14004
Reference [20]
V. Ginzburg, -modules, Springer’s representations and bivariant Chern classes, Adv. in Math. 61 (1986), 1–48.
Reference [21]
V. Ginzburg and E. Vasserot, Langlands reciprocity for affine quantum groups of type , International Math. Research Notices (1993) No.3, 67–85. MR 94j:17011
Reference [22]
V. Ginzburg, M. Kapranov and E. Vasserot, Langlands reciprocity for algebraic surfaces, Math. Res. Letters 2 (1995), 147–160. MR 96f:11086
Reference [23]
I. Grojnowski, Affinizing quantum algebras: From -modules to -theory, preprint, 1994.
Reference [24]
—, Instantons and affine algebras I: the Hilbert scheme and vertex operators, Math. Res. Letters 3 (1996), 275–291. MR 97f:14041
Reference [25]
G. Hatayama, A. Kuniba, M. Okado, T. Takagi and Y. Yamada, Remarks on fermionic formula, preprint, math.QA/9812022.
Reference [26]
D. Kazhdan and G. Lusztig, Proof of the Deligne-Langlands conjecture for Hecke algebras, Invent. Math. 87 (1987), 153–215. MR 88d:11121
Reference [27]
A. King, Moduli of representations of finite dimensional algebras, Quarterly J. of Math. 45 (1994), 515–530. MR 96a:16009
Reference [28]
M. Kleber, Finite dimensional representations of quantum affine algebras, preprint, math.QA/9809087.
Reference [29]
P.B. Kronheimer and H. Nakajima, Yang-Mills instantons on ALE gravitational instantons, Math. Ann. 288 (1990), 263–307. MR 92e:58038
Reference [30]
G. Lusztig, Green polynomials and singularities of unipotent classes, Adv. in Math. 42 (1981), 169–178. MR 83c:20059
[31]
—, Equivariant -theory and representations of Hecke algebras, Proc. Amer. Math. Soc. 94 (1985), 337–342. MR 88f:22054b
Reference [32]
—, Cuspidal local systems and graded Hecke algebras. I, Publ. Math. IHES 67 (1988), 145–202. MR 90e:16049
Reference [33]
—, Canonical bases arising from quantized enveloping algebras, J. Amer. Math. Soc. 3 (1990), 447–498. MR 90m:17023
Reference [34]
—, Quivers, perverse sheaves, and quantized enveloping algebras, J. Amer. Math. Soc. 4 (1991), 365–421. MR 91m:17018
Reference [35]
—, Affine quivers and canonical bases, Publ. Math. IHES 76 (1992), 111–163. MR 94h:16021
Reference [36]
—, Introduction to quantum group, Progress in Math. 110, Birkhäuser, 1993. MR 94m:17016
Reference [37]
—, Cuspidal local systems and graded Hecke algebras. II, in “Representation of Groups”, CMS Conf. Proc., 16, AMS, 1995, 217–275. MR 96m:22038
Reference [38]
—, On quiver varieties, Adv. in Math. 136 (1998), 141–182. MR 2000c:16016
Reference [39]
—, Bases in equivariant -theory. II, Representation Theory 3 (1999), 281–353. MR 2000h:20085
Reference [40]
—, Quiver varieties and Weyl group actions, preprint.
Reference [41]
I.G. Macdonald, Symmetric functions and Hall polynomials (2nd ed.), Oxford Math. Monographs, Oxford Univ. Press, 1995. MR 96h:05207
Reference [42]
A. Maffei, Quiver varieties of type , preprint, math.AG/9812142.
Reference [43]
D. Mumford, J. Fogarty and F. Kirwan, Geometric invariant theory, Third Enlarged Edition, Springer-Verlag, 1994. MR 95m:14012
Reference [44]
H. Nakajima, Instantons on ALE spaces, quiver varieties, and Kac-Moody algebras, Duke Math. J. 76 (1994), 365–416. MR 95i:53051
Reference [45]
—, Quiver varieties and Kac-Moody algebras, Duke Math. J. 91 (1998), 515–560. MR 99b:17033
Reference [46]
—, Lectures on Hilbert schemes of points on surfaces, Univ. Lect. Ser. 18, AMS, 1999. CMP 2000:02
Reference [47]
C.M. Ringel, Hall algebras and quantum groups, Invent. Math. 101 (1990), 583–592. MR 91i:16024
Reference [48]
Y. Saito, Quantum toroidal algebras and their vertex representations, Publ. RIMS 34 (1998), 155–177. MR 99d:17022
Reference [49]
Y. Saito, K. Takemura, and D. Uglov, Toroidal actions on level modules of , Transform. Group 3 (1998), 75–102. MR 99b:17015
Reference [50]
R. Sjamaar and E. Lerman, Stratified symplectic spaces and reduction, Ann. of Math. 134 (1991), 375–422. MR 92g:58036
Reference [51]
R. Sjamaar, Holomorphic slices, sympletic reduction and multiplicities of representations, Ann. of Math. 141 (1995), 87–129. MR 96a:58098
Reference [52]
T. Tanisaki, Hodge modules, equivariant -theory and Hecke algebras, Publ. RIMS 23 (1987), 841–879. MR 89c:20070
Reference [53]
R. Thomason, Algebraic -theory of group scheme actions, in “Algebraic Topology and Algebraic -theory”, Ann. of Math. Studies 113 (1987), 539–563. MR 89c:18016
Reference [54]
—, Equivariant algebraic vs. topological -homology Atiyah-Segal style, Duke Math. J. 56 (1988), 589–636. MR 89f:14015
Reference [55]
—, Une formule de Lefschetz en -théorie équivariante algébrique, Duke Math. J. 68 (1992), 447–462. MR 93m:19007
Reference [56]
M. Varagnolo and E. Vasserot, Double-loop algebras and the Fock space, Invent. Math. 133 (1998), 133–159. MR 99g:17035
Reference [57]
—, On the -theory of the cyclic quiver variety, preprint, math.AG/9902091.
Reference [58]
E. Vasserot, Affine quantum groups and equivariant -theory, Transformation Groups 3 (1998), 269–299. MR 99j:19007

Article Information

MSC 2000
Primary: 17B37 (Quantum groups and related deformations)
Secondary: 14D21 (Applications of vector bundles and moduli spaces in mathematical physics), 14L30 (Group actions on varieties or schemes), 16G20 (Representations of quivers and partially ordered sets), 33D80 (Connections with quantum groups, Chevalley groups, -adic groups, Hecke algebras, and related topics)
Author Information
Hiraku Nakajima
Department of Mathematics, Kyoto University, Kyoto 606-8502, Japan
nakajima@kusm.kyoto-u.ac.jp
MathSciNet
Additional Notes

The author was supported by the Grant-in-aid for Scientific Research (No.11740011), the Ministry of Education, Japan, and National Science Foundation Grant #DMS 97-29992.

Journal Information
Journal of the American Mathematical Society, Volume 14, Issue 1, ISSN 1088-6834, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on , revised on , and published on .
Copyright Information
Copyright 2000 American Mathematical Society
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/S0894-0347-00-00353-2
  • MathSciNet Review: 1808477
  • Show rawAMSref \bib{1808477}{article}{ author={Nakajima, Hiraku}, title={Quiver varieties and finite dimensional representations of quantum affine algebras}, journal={J. Amer. Math. Soc.}, volume={14}, number={1}, date={2001-01}, pages={145-238}, issn={0894-0347}, review={1808477}, doi={10.1090/S0894-0347-00-00353-2}, }

Settings

Change font size
Resize article panel
Enable equation enrichment

Note. To explore an equation, focus it (e.g., by clicking on it) and use the arrow keys to navigate its structure. Screenreader users should be advised that enabling speech synthesis will lead to duplicate aural rendering.

For more information please visit the AMS MathViewer documentation.