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