The McKay correspondence as an equivalence of derived categories

By Tom Bridgeland, Alastair King, and Miles Reid

To Andrei Tyurin on his 60th birthday

Abstract

Let be a finite group of automorphisms of a nonsingular three-dimensional complex variety , whose canonical bundle is locally trivial as a -sheaf. We prove that the Hilbert scheme parametrising -clusters in is a crepant resolution of and that there is a derived equivalence (Fourier–Mukai transform) between coherent sheaves on and coherent -sheaves on . This identifies the K theory of with the equivariant K theory of , and thus generalises the classical McKay correspondence. Some higher-dimensional extensions are possible.

1. Introduction

The classical McKay correspondence relates representations of a finite subgroup to the cohomology of the well-known minimal resolution of the Kleinian singularity . Gonzalez-Sprinberg and Verdier Reference 10 interpreted the McKay correspondence as an isomorphism on K theory, observing that the representation ring of is equal to the -equivariant K theory of . More precisely, they identify a basis of the K theory of the resolution consisting of the classes of certain tautological sheaves associated to the irreducible representations of .

It is natural to ask what happens when is replaced by an arbitrary nonsingular quasiprojective complex variety of dimension and by a finite group of automorphisms of , with the property that the stabiliser subgroup of any point acts on the tangent space as a subgroup of . Thus the canonical bundle is locally trivial as a -sheaf, in the sense that every point of has a -invariant open neighbourhood on which there is a nonvanishing -invariant -form. This implies that the quotient variety has only Gorenstein singularities.

A natural generalisation of the McKay correspondence would then be an isomorphism between the -equivariant K theory of and the ordinary K theory of a crepant resolution of , that is, a resolution of singularities such that . In the classical McKay case, the minimal resolution is crepant, but in higher dimensions crepant resolutions do not necessarily exist and, even when they do, they are not usually unique. However, it is now known that crepant resolutions of Gorenstein quotient singularities do exist in dimension , through a case by case analysis of the local linear actions by Ito, Markushevich and Roan (see Roan Reference 21 and references given there). In dimension , even such quotient singularities only have crepant resolutions in rather special cases.

In this paper, we take the point of view that the appropriate way to formulate and prove the McKay correspondence on K theory is to lift it to an equivalence of derived categories. In itself, this is not a new observation and it turns out that it was actually known to Gonzalez-Sprinberg and Verdier (see also Reid Reference 20, Conjecture 4.1). Furthermore, if the resolution is constructed as a moduli space of -equivariant objects on , then the correspondence should be given by a Fourier-Mukai transform determined by the universal object. This is the natural analogue of the classical statement that the tautological sheaves are a basis of the K theory. Both points of view are taken by Kapranov and Vasserot Reference 15 in proving the derived category version of the classical two-dimensional McKay correspondence.

The new and remarkable feature is that, by using the derived category and Fourier-Mukai transforms and, in particular, techniques developed in Reference 6 and Reference 7, the process of proving the equivalence of derived categories—when it works—also yields a proof that the moduli space is a crepant resolution. More specifically, we will give a sufficient condition for a certain natural moduli space, namely Nakamura’s -Hilbert scheme, to be a crepant resolution for which the McKay correspondence holds as an equivalence of derived categories. This condition is automatically satisfied in dimensions 2 and 3. Thus we simultaneously prove the existence of one crepant resolution of in three dimensions, without a case by case analysis, and verify the McKay correspondence for this resolution. We do not prove the McKay correspondence for an arbitrary crepant resolution although our methods should easily adapt to more general moduli spaces of -sheaves on , which may provide different crepant resolutions to the one considered here.

The -Hilbert scheme was introduced by Nakamura as a good candidate for a crepant resolution of . It parametrises -clusters or ‘scheme theoretic -orbits’ on : recall that a cluster is a zero-dimensional subscheme, and a -cluster is a -invariant cluster whose global sections are isomorphic to the regular representation of . Clearly, a -cluster has length and a free -orbit is a -cluster. There is a Hilbert–Chow morphism

which, on closed points, sends a -cluster to the orbit supporting it. Note that is a projective morphism, is onto and is birational on one component.

When and is Abelian, Nakamura Reference 18 proved that is irreducible and is a crepant resolution of (compare also Reid Reference 20 and Craw and Reid Reference 8). He conjectured that the same result holds for an arbitrary finite subgroup . Ito and Nakajima Reference 12 observed that the construction of Gonzalez-Sprinberg and Verdier Reference 10 is the case of a natural correspondence between the equivariant K theory of and the ordinary K theory of . They proved that this correspondence is an isomorphism when and is Abelian by constructing an explicit resolution of the diagonal in Beilinson style. Our approach via Fourier–Mukai transforms leaves this resolution of the diagonal implicit (it appears as the object of in Section 6), and seems to give a more direct argument. Two of the main consequences of the results of this paper are that Nakamura’s conjecture is true and that the natural correspondence on K theory is an isomorphism for all finite subgroups of .

Since it is not known whether is irreducible or even connected in general, we actually take as our initial candidate for a resolution the irreducible component of containing the free -orbits, that is, the component mapping birationally to . The aim is to show that is a crepant resolution, and to construct an equivalence between the derived categories of coherent sheaves on and of coherent -sheaves on . A more detailed analysis of the equivalence shows that when has dimension 3.

We now describe the correspondence and our results in more detail. Let be a nonsingular quasiprojective complex variety of dimension and let be a finite group of automorphisms of such that is locally trivial as a -sheaf. Put and let be the irreducible component containing the free orbits, as described above. Write for the universal closed subscheme and and for its projections to and . There is a commutative diagram of schemes

in which and are birational, and are finite, and is flat. Let act trivially on and , so that all morphisms in the diagram are equivariant.

Define the functor

where a sheaf on is viewed as a -sheaf by giving it the trivial action. Note that is already exact, so we do not need to write . Our main result is the following.

Theorem 1.1.

Suppose that the fibre product

has dimension . Then is a crepant resolution of and is an equivalence of categories.

When the condition of the theorem always holds because the exceptional locus of has dimension . In this case we can also show that is irreducible, so we obtain

Theorem 1.2.

Suppose . Then is irreducible and is a crepant resolution of , and is an equivalence of categories.

The condition of Theorem 1.1 also holds whenever preserves a complex symplectic form on and is a crepant resolution of , because such a resolution is symplectic and hence semi-small (see Verbitsky Reference 24, Theorem 2.8 and compare Kaledin Reference 14).

Corollary 1.3.

Suppose is a complex symplectic variety and acts by symplectic automorphisms. Assume that is a crepant resolution of . Then is an equivalence of categories.

Note that the condition of Theorem 1.1 certainly fails in dimension whenever has an exceptional divisor over a point. This is to be expected since there are many examples of finite subgroups for which the quotient singularity has no crepant resolution and also examples where, although crepant resolutions do exist, is not one.

Conventions

We work throughout in the category of schemes over . A point of a scheme always means a closed point.

2. Category theory

This section contains some basic category theory, most of which is well known. The only nontrivial part is Section 2.6 where we state a condition for an exact functor between triangulated categories to be an equivalence.

2.1. Triangulated categories

A triangulated category is an additive category equipped with a shift automorphism and a collection of distinguished triangles

of morphisms of satisfying certain axioms (see Verdier Reference 25). We write for and

A triangulated category is trivial if every object is a zero object.

The principal example of a triangulated category is the derived category of an Abelian category . An object of is a bounded complex of objects of up to quasi-isomorphism, the shift functor moves a complex to the left by one place and a distinguished triangle is the mapping cone of a morphism of complexes. In this case, for objects , one has .

A functor between triangulated categories is exact if it commutes with the shift automorphisms and takes distinguished triangles of to distinguished triangles of . For example, derived functors between derived categories are exact.

2.2. Adjoint functors

Let and be functors. An adjunction for is a bifunctorial isomorphism

In this case, we say that is left adjoint to or that is right adjoint to . When it exists, a left or right adjoint to a given functor is unique up to isomorphism of functors. The adjoint of a composite functor is the composite of the adjoints. An adjunction determines and is determined by two natural transformations and that come from applying the adjunction to and respectively (see Mac Lane Reference 16, IV.1 for more details).

The basic adjunctions we use in this paper are described in Section 3.1 below.

2.3. Fully faithful functors and equivalences

A functor is fully faithful if for any pair of objects , of , the map

is an isomorphism. One should think of as an ‘injective’ functor. This is clearer when has a left adjoint (or a right adjoint ), in which case is fully faithful if and only if the natural transformation (or ) is an isomorphism.

A functor is an equivalence if there is an ‘inverse’ functor such that and . In this case is both a left and right adjoint to (see Mac Lane Reference 16, IV.4). In practice, we show that is an equivalence by writing down an adjoint (a priori, one-sided) and proving that it is an inverse. One simple example of this is the following.

Lemma 2.1.

Let and be triangulated categories and a fully faithful exact functor with a right adjoint . Then is an equivalence if and only if implies for all objects .

Proof.

By assumption is an isomorphism, so is an equivalence if and only if is an isomorphism. Thus the ‘only if’ part of the lemma is immediate, since .

For the ‘if’ part, take any object and embed the natural adjunction map in a triangle

If we apply to this triangle, then is an isomorphism, because is an isomorphism and (Reference 16, IV.1, Theorem 1). Hence and so by hypothesis. Thus is an isomorphism, as required.

One may understand this lemma in a broader context as follows. The triangle (Equation 1) shows that, when is fully faithful with right adjoint , there is a ‘semi-orthogonal’ decomposition , where

Since is fully faithful, the fact that for some object necessarily means that , so only zero objects are in both subcategories. The semi-orthogonality condition also requires that for all and , which is immediate from the adjunction. The lemma then has the very reasonable interpretation that if is trivial, then and is an equivalence. Note that if is a left adjoint for , then there is a similar semi-orthogonal decomposition on the other side and a corresponding version of the lemma. For more details on semi-orthogonal decompositions see Bondal Reference 4.

2.4. Spanning classes and orthogonal decomposition

A spanning class for a triangulated category is a subclass of the objects of such that for any object

and

The following easy lemma is Reference 6, Example 2.2.

Lemma 2.2.

The set of skyscraper sheaves on a nonsingular projective variety is a spanning class for .

A triangulated category is decomposable as an orthogonal direct sum of two full subcategories and if every object of is isomorphic to a direct sum with , and if

for any pair of objects and all integers . The category is indecomposable if for any such decomposition one of the two subcategories is trivial. For example, if is a scheme, is indecomposable precisely when is connected. For more details see Bridgeland Reference 6.

2.5. Serre functors

The properties of Serre duality on a nonsingular projective variety were abstracted by Bondal and Kapranov Reference 5 into the notion of a Serre functor on a triangulated category. Let be a triangulated category in which all the sets are finite dimensional vector spaces. A Serre functor for is an exact equivalence inducing bifunctorial isomorphisms

that satisfy a simple compatibility condition (see Reference 5). When a Serre functor exists, it is unique up to isomorphism of functors. We say that has trivial Serre functor if for some integer the shift functor is a Serre functor for .

The main example is the bounded derived category of coherent sheaves on a nonsingular projective variety , having the Serre functor

Thus has trivial Serre functor if and only if the canonical bundle of is trivial.

2.6. A criterion for equivalence

Let be an exact functor between triangulated categories with Serre functors and . Assume that has a left adjoint . Then also has a right adjoint .

Theorem 2.3.

With assumptions as above, suppose also that there is a spanning class for such that

is an isomorphism for all and all . Then is fully faithful.

Proof.

See Reference 6, Theorem 2.3.

Theorem 2.4.

Suppose further that is nontrivial, that is indecomposable and that for all . Then is an equivalence of categories.

Proof.

Consider an object . For any and we have isomorphisms

using Serre duality and the adjunctions for and . Since is a spanning class we can conclude that precisely when . Then the result follows from Reference 6, Theorem 3.3.

The proof of Theorem 3.3 in Reference 6 may be understood as follows. If , then the semi-orthogonal decomposition described at the end of Section 2.3 becomes an orthogonal decomposition. Hence must be trivial, because is indecomposable and , and hence , is nontrivial. Thus and is an equivalence.

3. Derived categories of sheaves

This section is concerned with various general properties of complexes of -modules on a scheme . Note that all our schemes are of finite type over . Given a scheme , define to be the (unbounded) derived category of the Abelian category of quasicoherent sheaves on . Also define to be the full subcategory of consisting of complexes with bounded and coherent cohomology.

3.1. Geometric adjunctions

Here we describe three standard adjunctions that arise in algebraic geometry and are used frequently in what follows. For the first example, let be a scheme and an object of finite homological dimension. Then the derived dual

also has finite homological dimension, and the functor is both left and right adjoint to the functor .

For the second example take a morphism of schemes . The functor

has the left adjoint

If is proper, then takes into . If has finite Tor dimension (for example if is flat, or is nonsingular), then takes into .

The third example is Grothendieck duality. Again take a morphism of schemes . The functor has a right adjoint

and moreover, if is proper and of finite Tor dimension, there is an isomorphism of functors

Neeman Reference 19 has recently given a completely formal proof of these statements in terms of the Brown representability theorem.

Let be a nonsingular projective variety of dimension and write for the projection to a point. In this case . The above statement of Grothendieck duality implies that the functor

is a Serre functor on .

3.2. Duality for quasiprojective schemes

In order to apply Grothendieck duality on quasiprojective schemes, we need to restrict attention to sheaves with compact support. The support of an object is the locus of where is not exact, that is, the union of the supports of the cohomology sheaves of . It is always a closed subset of .

Given a scheme , define the category to be the full subcategory of consisting of complexes whose support is proper. Note that when itself is proper, is just the usual derived category .

If is a quasiprojective variety and is some projective closure, then the functor embeds as a full triangulated subcategory of . By resolution of singularities, if is nonsingular we can assume that is too. Then the Serre functor on restricts to give a Serre functor on . Thus if is a nonsingular quasiprojective variety of dimension , the category has a Serre functor given by (Equation 3).

The argument used to prove Lemma 2.2 is easily generalised to give the statement that the set of skyscraper sheaves on a nonsingular quasiprojective variety is a spanning class for .

3.3. Crepant resolutions

Let be a variety and a resolution of singularities. Given a point define to be the full subcategory of consisting of objects whose support is contained in the fibre . We have the following categorical criterion for to be crepant.

Lemma 3.1.

Assume that has rational singularities, that is, . Suppose has trivial Serre functor for each . Then is Gorenstein and is a crepant resolution.

Proof.

The Serre functor on is the restriction of the Serre functor on . Hence, by Section 3.2, the condition implies that for each the restriction of the functor to the category is isomorphic to the identity. Since contains the structure sheaves of all fattened neighbourhoods of the fibre this implies that the restriction of to each formal fibre of is trivial. To get the result, we must show that is a line bundle and that . Since , this is achieved by the following lemma.

Lemma 3.2.

Assume that has rational singularities. Then a line bundle on is the pullback of some line bundle on if and only if the restriction of to each formal fibre of is trivial. Moreover, when this holds, .

Proof.

For each point , the formal fibre of over is the fibre product

The restriction of the pullback of a line bundle from to each of these schemes is trivial because a line bundle has trivial formal stalks at points.

For the converse suppose that the restriction of to each of these formal fibres is trivial. The theorem on formal functions shows that the completions of the stalks of the sheaves and at any point are isomorphic for each . Since has rational singularities it follows that for all , and is a line bundle on .

Since is torsion free, the natural adjunction map is injective, so there is a short exact sequence

By the projection formula and the fact that has rational singularities,

The fact that is the unit of the adjunction for implies that has a left inverse, and in particular is surjective. Applying to (Equation 4) we conclude that .

Using the theorem on formal functions again, we can deduce that

In particular, has no nonzero global sections. Tensoring (Equation 4) with gives a contradiction unless . Hence is an isomorphism and we are done.

4. -sheaves

Throughout this section is a finite group acting on a scheme (on the left) by automorphisms. As in the last section, all schemes are of finite type over . We list some results we need concerning the category of sheaves on equipped with a compatible action, or -sheaves’ for short. Since is finite, most of the proofs are trivial and are left to the reader. The main point is that natural constructions involving sheaves on are canonical, so commute with automorphisms of .

4.1. Sheaves and functors

A -sheaf on is a quasicoherent sheaf of -modules together with a lift of the action to . More precisely, for each , there is a lift satisfying and .

If and are -sheaves, then there is a (right) action of on given by and the spaces of -invariant maps give the morphisms in the Abelian categories and of -sheaves.

The category has enough injectives (Grothendieck Reference 9, Proposition 5.1.2) so we may take -equivariant injective resolutions. Since is finite, if is a quasiprojective scheme there is an ample invertible -sheaf on and so we may also take -equivariant locally free resolutions. The functors are the -invariant parts of and are the derived functors of . Thus if is nonsingular of dimension , so that has global dimension , then the category also has global dimension .

The local functors and are defined in the obvious way on , as are pullback and pushforward for any -equivariant morphism of schemes . Thus, for example, . Natural isomorphisms such as are canonical, that is, commute with isomorphisms of the base, and hence are -equivariant. Therefore they restrict to natural isomorphisms

In other words, and are also adjoint functors between the categories and .

Similarly, the natural isomorphisms implicit in the projection formula, flat base change, etc. are canonical and hence -equivariant.

It seems worthwhile to single out the following point:

Lemma 4.1.

Let and be -sheaves on . Then, as a representation of , we have a direct sum decomposition

over the irreducible representations .

Proof.

The result amounts to showing that

Let be projection to a point, with acting trivially on so that the map is equivariant. Then is just the category of -modules. Note that and , so that the adjunction between and gives

as required.

4.2. Trivial actions

If the group acts trivially on , then any -sheaf decomposes as a direct sum

over the irreducible representations of (where is the trivial representation). The sheaves are just ordinary sheaves on . Furthermore, for . Thus the category decomposes as a direct sum and each summand is equivalent to .

In particular, every -sheaf has a fixed part and the functor

is the left and right adjoint to the functor

that is, ‘let act trivially’. Both functors are exact.

4.3. Derived categories

The -equivariant derived category is defined to be the full subcategory of the (unbounded) derived category of consisting of complexes with bounded and coherent cohomology.

The usual derived functors , , and may be defined on the equivariant derived category, and, as for sheaves, the standard properties of adjunctions, projection formula and flat base change then hold because the implicit natural isomorphisms are sufficiently canonical.

One way to obtain an equivariant Grothendieck duality is to refer to Neeman’s results Reference 19. Let be an equivariant morphism of schemes. The only thing to check is that equivariant pushdown commutes with small coproducts. This is proved exactly as in Reference 19. Then the functor has a right adjoint , and (Equation 2) holds when is proper and of finite Tor dimension.

As in the nonequivariant case this implies that if is a nonsingular quasiprojective variety of dimension , the full subcategory consisting of objects with compact supports has a Serre functor

where is the canonical bundle of with its induced -structure.

4.4. Indecomposability

If acts trivially on , then the results of Section 4.2 show that decomposes as a direct sum of orthogonal subcategories indexed by the irreducible representations of . More generally it is easy to see that is decomposable unless acts faithfully. We need the converse of this statement.

Lemma 4.2.

Suppose a finite group acts faithfully on a quasiprojective variety . Then is indecomposable.

Proof.

Suppose that decomposes as an orthogonal direct sum of two subcategories and . Any indecomposable object of lies in either or and

Since the action of is faithful, the general orbit is free. Let be a free orbit. Then is indecomposable as a -sheaf. Suppose without loss of generality that lies in .

Let be an irreducible representation of . The sheaf is indecomposable in and there exists an equivariant map so also lies in . Any indecomposable -sheaf supported in dimension 0 has a section, so by Lemma 4.1 there is an equivariant map , and thus lies in .

Finally given an indecomposable -sheaf , take an orbit contained in and let be the inclusion. Then is supported in dimension 0 and there is an equivariant map , so also lies in . Now is orthogonal to all sheaves, hence is trivial.

5. The intersection theorem

Our proof that is nonsingular follows an idea developed in Bridgeland and Maciocia Reference 7 for moduli spaces over K3 fibrations, and uses the following famous and difficult result of commutative algebra:

Theorem 5.1 (Intersection theorem).

Let be a local -algebra of dimension . Suppose that

is a complex of finitely generated free -modules with each homology module an -module of finite length. Then . Moreover, if and , then

and is regular.

The basic idea is as follows. Serre’s criterion states that any finite length -module has homological dimension and that is regular precisely if there is a finite length -module which has homological dimension exactly . The intersection theorem gives corresponding statements for complexes of -modules with finite length homology. As a rough slogan, “regularity is a property of the derived category”. For the main part of the proof, see Roberts Reference 22, Reference 23; for the final clause, see Reference 7.

We may rephrase the intersection theorem using the language of support and homological dimension. If is a scheme and an object in , then it is easy to check Reference 7 that, for any point ,

The homological dimension of a nonzero object , written , is the smallest nonnegative integer such that is isomorphic in to a complex of locally free sheaves on of length (that is, having terms). If no such integer exists we put . One can prove Reference 7 that if is quasiprojective, and is a nonnegative integer, then if and only if there is an integer such that for all points

The two parts of Theorem 5.1 now become the following (cf. Reference 7).

Corollary 5.2.

Let be a scheme and a nonzero object of . Then

Corollary 5.3.

Let be a scheme, and fix a point of codimension . Suppose that there is an object of such that for all points , and any integer ,

Suppose also that . Then is nonsingular at and .

6. The projective case

The aim of this section is to prove Theorem 1.1 under the additional assumption that is projective. The quasiprojective case involves some further technical difficulties that we deal with in the next section. Take notation as in the Introduction. We break the proof up into seven steps.

Step 1. Let and denote the projections. The functor may be rewritten

Note that has finite homological dimension, because is flat over and is nonsingular. Hence the derived dual also has finite homological dimension and we may define another functor , by the formula

where .

Now is left adjoint to because of the three standard adjunctions described in Section 3.1. The functor is the left adjoint to . The functor has the (left and right) adjoint . Finally the functor has the left adjoint and

Step 2. The composite functor is given by

where and are the projections of onto its factors, and is some object of . This is just composition of correspondences (see Mukai Reference 17, Proposition 1.3).

If is the closed embedding, then . For any pair of points , one has so that

using the adjunctions for and . Our first objective is to show that is supported on the diagonal , or equivalently that the groups in (Equation 5) vanish unless . When this plays the rôle of assumption (4.8) of Ito and Nakajima Reference 12.

Step 3. Let be -clusters. Then

To see this note that is generated as an -module by any nonzero constant section. But, since the global sections form the regular representation of , the constant sections are precisely the -invariant sections. Hence any equivariant morphism maps a generator to a scalar multiple of a generator and so is determined by that scalar.

Let and be distinct points of . Serre duality, together with our assumption that is locally trivial as a -sheaf, implies that

so that

Hence restricted to has homological dimension .

Step 4. Now we apply the intersection theorem. If and are points of such that , then the corresponding clusters and are disjoint, so that the groups in (Equation 5) vanish. Thus the support of is contained in the subscheme . By assumption this has codimension so Corollary 5.2 implies that

that is, is supported on the diagonal.

Step 5. Fix a point , and put . We proved above that is supported at the point . We claim that . Note that Corollary 5.3 then implies that is nonsingular at and .

To prove the claim, note that there is a canonical map , so we obtain a triangle

for some object of . Using the adjoint pair , this gives a long exact sequence

The homomorphism is the Kodaira–Spencer map for the family of clusters (Bridgeland Reference 6, Lemma 4.4). This is injective because is a fine moduli space for -clusters on . It follows that

An easy spectral sequence argument (see Reference 6, Example 2.2) shows that . Taking cohomology sheaves of the above triangle gives , which proves the claim.

Step 6. We have now proved that is nonsingular, and that for any pair of points , the homomorphisms

are isomorphisms. By assumption, the action of on is such that is trivial as a -sheaf on an open neighbourhood of each orbit . This implies that

in , for each . Applying Theorem 2.4 shows that is an equivalence of categories.

Step 7. It remains to show that is crepant. Take a point . The equivalence restricts to give an equivalence between the full subcategories and consisting of objects supported on the fibre and the orbit respectively.

The category has trivial Serre functor because is trivial as a -sheaf on a neighbourhood of . Thus also has trivial Serre functor and Lemma 3.1 gives the result.

This completes the proof of Theorem 1.1 in the case that is projective.

7. The quasiprojective case

In this section we complete the proof of Theorem 1.1. Once again, take notation as in the Introduction. The problem with the argument of the last section is that when is not projective, Grothendieck duality in the form we need only applies to objects with compact support. To get round this we first take a projective closure of and define adjoint functors as before. Then we restrict to a functor

The argument of the last section carries through to show that is nonsingular and crepant and that is an equivalence. It remains for us to show that is also an equivalence.

Step 8. The functor has a right adjoint

As before, the composition is given by

where and are the projections of onto its factors, and is some object of .

Since is an equivalence, for any point , and it follows that is actually the pushforward of a line bundle on to the diagonal in . The functor is then just twisting by , and to show that is fully faithful we must show that is trivial.

There is a morphism of functors , which for any point gives a commutative diagram

where is nonzero. Since is an isomorphism on the subcategory , the maps are all isomorphisms, so the section is an isomorphism.

Step 9. The fact that is an equivalence follows from Lemma 2.1 once we show that

Suppose . Using the adjunction for ,

whenever for some object . In particular, this holds for any with compact support.

If is nonzero, let be an orbit of contained in the support of . Let denote the inclusion, a projective equivariant morphism of schemes. Then the adjunction morphism is nonzero, which gives a contradiction.

This completes the proof of Theorem 1.1.

8. Nakamura’s conjecture

Recall that in Theorem 1.1 we took the space to be an irreducible component of . Note that when is nonsingular and is an equivalence, is actually a connected component. This is simply because for any point , the bijection

identifies the tangent space of at with the tangent space of at . In this section we wish to go further and prove that when has dimension 3, is in fact connected.

Proof of Nakamura’s conjecture.

Suppose for contradiction that there exists a -cluster not contained among the . Since is an equivalence we can take an object such that . The argument of Section Equation 6, Step 3, shows that for any point

This implies that has homological dimension 1, or more precisely, that is quasi-isomorphic to a complex of locally free sheaves of the form

But is supported on some -orbit in , so is supported on a fibre of , and hence in codimension . It follows that the complex (Equation 6) is exact on the left, so . In particular in the Grothendieck group of .

Let be a point of the fibre that is the support of . By Lemma 8.1 below, in , so that in , since the equivalence gives an isomorphism of Grothendieck groups.

Let be a nonsingular projective variety with an open inclusion . The functor induces a map on K groups, so in . But this contradicts Riemann–Roch, because if is a sufficiently ample line bundle on , then and are both positive.

Lemma 8.1.

If and are two -clusters on supported on the same orbit, then the corresponding elements and in the Grothendieck group of are equal.

Proof.

We need to show that, as -sheaves, and have composition series with the same simple factors. Suppose that they are both supported on the -orbit and let be the stabiliser subgroup of in . The restriction functor is an equivalence of categories from finite length -sheaves supported on to finite length -sheaves supported at . The reverse equivalence is the induction functor . Since the restriction of a -cluster supported on is an -cluster supported at , it is sufficient to prove the result for -clusters supported at .

If are the irreducible representations of , then we claim that the simple -sheaves supported at are precisely

These sheaves are certainly simple, since they are simple as -modules. On the other hand, any -sheaf supported at has a nonzero ordinary sheaf morphism . By Lemma 4.1 there must be a nonzero -sheaf morphism , for some , and, if were simple, then this would have to be an isomorphism.

Thus a composition series as an -sheaf is also a composition series as a -module. Hence all -clusters supported at have the same composition factors as -sheaves, since as -modules they are all the regular representation of .

9. K theoretic consequences of equivalence

In this section we put and assume that the functor is an equivalence of categories. This is always the case when . The main point is that such an equivalence of derived categories immediately gives an isomorphism of the corresponding Grothendieck groups.

9.1. Restricting to the exceptional fibres

Let denote the full subcategory of consisting of objects supported at the origin of . Similarly, let denote the full subcategory of consisting of objects supported on the subscheme of .

The equivalence induces an equivalence

so we obtain a diagram

in which the vertical arrows are embeddings of categories.

Note that the Euler characteristic gives natural bilinear pairings between the top and bottom categories on either side; if and are objects of and respectively, then we can compute the sums

because the Hom spaces are finite dimensional (even over a quasiprojective variety) when has finite length cohomology sheaves. Similarly, we can compute the ordinary Euler character on the left. The fact that is an equivalence of categories commuting with the shift functors immediately gives

for any objects of and of .

9.2. Equivalence of K groups

Let , , and be the Grothendieck groups of the corresponding derived categories. The equivalences of categories from the last section immediately give isomorphisms of these groups. The following lemma is proved in the same way as in Gonzalez-Sprinberg and Verdier Reference 10, Proposition 1.4.

Lemma 9.1.

The maps that send a representation of to the -sheaves and on give ring isomorphisms of the representation ring with and respectively.

We obtain a diagram of groups

in which the horizontal maps are isomorphisms but the vertical maps are not. In fact, if is the representation induced by the inclusion , then the map is multiplication by

This formula is obtained by considering a Koszul resolution of on , as in Reference 10, Proposition 1.4. For example, in the case one has .

The bilinear forms of Section 9.1 descend to give pairings on the Grothendieck groups. These forms are nondegenerate because if are the irreducible representations of , then the corresponding bases

are dual with respect to the pairing . Applying gives dual bases

as in Ito and Nakajima Reference 12.

10. Topological K theory and physics

With notation as in the Introduction, suppose that is projective, and further that is nonsingular and is an equivalence. For example suppose that or .

10.1. K theory and the orbifold Euler number

Let denote the topological complex K theory of and the -equivariant topological K theory of . There are natural forgetful maps

Since and its inverse are defined as correspondences, we may define correspondences

compatible with the maps , using the functors , and (also written ) on topological K theory, which extend to equivariant K theory, as usual, because they are canonical. Note that the definition and compatibility of is nontrivial; see Reference 1 for more details. But now the fact that and are mutually inverse implies that and are mutually inverse, that is, we have a graded isomorphism

Atiyah and Segal Reference 2 observed that the physicists’ orbifold Euler number of is the Euler characteristic of , that is,

On the other hand, since the Chern character gives a graded isomorphism , the Euler characteristic of is just the ordinary Euler number of . Hence the isomorphism (Equation 7) on topological K theory provides a natural explanation for the physicists’ Euler number conjecture

This was verified in the case as a consequence of the original McKay correspondence (cf. Reference 2). It was proved in the case by Roan Reference 21 in the more general case of quasiprojective Gorenstein orbifolds, since the numerical statement reduces to the local linear case , .

10.2. An example: The Kummer surface

One of the first interesting cases of the isomorphism (Equation 7) is when is an Abelian surface (topologically, a 4-torus ), acting by the involution and is a K3 surface. In this case is a nonsingular Kummer surface, having 16 disjoint -curves coming from resolving the images in of the 16 -fixed points in . Write for this fixed point set.

On the Abelian surface there are 32 flat line -bundles, arising from a choice of 2 -actions on each of the 16 square roots of . Each such flat line -bundle is characterised by a map such that at a fixed point the group acts on the fibre with weight . Now the set naturally has the structure of an affine -space over and the maps that occur are precisely the affine linear maps, including the two constant maps corresponding to the two actions on .

On the other hand, on the K3 surface one may consider the lattice spanned by and the smallest primitive sublattice containing . The elements of give precisely the rational linear combinations of the divisors which are themselves divisors. It is easy to see that and it can also be shown that the image of in the quotient consists of precisely the affine linear maps on (see Barth, Peters and Van de Ven Reference 3, Chapter VIII, Proposition 5.5).

We claim that under the correspondence , the flat line -bundle is taken to the line bundle , where

To check the claim note that is taken to , and that, in the local linear McKay correspondence for , the irreducible representation of weight is taken to the line bundle , dual to the -curve resolving the singularity.

Acknowledgements

The first author would like to thank the ICTP, Trieste, and EPSRC for financial support whilst this paper was written.

Mathematical Fragments

Theorem 1.1.

Suppose that the fibre product

has dimension . Then is a crepant resolution of and is an equivalence of categories.

Lemma 2.1.

Let and be triangulated categories and a fully faithful exact functor with a right adjoint . Then is an equivalence if and only if implies for all objects .

Equation (1)
Lemma 2.2.

The set of skyscraper sheaves on a nonsingular projective variety is a spanning class for .

Theorem 2.4.

Suppose further that is nontrivial, that is indecomposable and that for all . Then is an equivalence of categories.

Equation (2)
Equation (3)
Lemma 3.1.

Assume that has rational singularities, that is, . Suppose has trivial Serre functor for each . Then is Gorenstein and is a crepant resolution.

Equation (4)
Lemma 4.1.

Let and be -sheaves on . Then, as a representation of , we have a direct sum decomposition

over the irreducible representations .

Theorem 5.1 (Intersection theorem).

Let be a local -algebra of dimension . Suppose that

is a complex of finitely generated free -modules with each homology module an -module of finite length. Then . Moreover, if and , then

and is regular.

Corollary 5.2.

Let be a scheme and a nonzero object of . Then

Corollary 5.3.

Let be a scheme, and fix a point of codimension . Suppose that there is an object of such that for all points , and any integer ,

Suppose also that . Then is nonsingular at and .

Equation (5)
Equation (6)
Lemma 8.1.

If and are two -clusters on supported on the same orbit, then the corresponding elements and in the Grothendieck group of are equal.

Equation (7)

References

Reference [1]
M.F. Atiyah and F. Hirzebruch, The Riemann–Roch theorem for analytic embeddings, Topology 1 (1962) 151–166. MR 26:5593
Reference [2]
M.F. Atiyah and G. Segal, On equivariant Euler characteristics, J. Geom. Phys. 6 (1989) 671–677. MR 92c:19005
Reference [3]
W. Barth, C. Peters and A. Van de Ven, Compact Complex Surfaces, Ergeb. Math. 4, Springer, 1984. MR 86c:32026
Reference [4]
A. Bondal, Representation of associative algebras and coherent sheaves, Math. USSR Izv. 34 (1990) 23-41. MR 90i:14017
Reference [5]
A. Bondal and M. Kapranov, Representable functors, Serre functors, and mutations, Math. USSR Izv. 35 (1990) 519–541. MR 91b:14013
Reference [6]
T. Bridgeland, Equivalences of triangulated categories and Fourier–Mukai transforms, Bull. London Math. Soc. 31 (1999) 25–34. MR 99k:18014
Reference [7]
T. Bridgeland and A. Maciocia, Fourier–Mukai transforms for K3 and elliptic fibrations, preprint, math.AG 9908022.
Reference [8]
A. Craw and M. Reid, How to calculate , preprint, math.AG 9909085, 29 pp.
Reference [9]
A. Grothendieck, Sur quelques points d’algèbre homologique, Tôhoku Math. J. 9 (1957) 119–221. MR 21:1328
Reference [10]
G. Gonzalez-Sprinberg and J.-L. Verdier, Construction géométrique de la correspondance de McKay, Ann. Sci. École Norm. Sup. 16 (1983) 409–449. MR 85k:14019
[11]
R. Hartshorne, Residues and Duality, Lect. Notes Math. 20, Springer (1966). MR 36:5145
Reference [12]
Y. Ito and H. Nakajima, McKay correspondence and Hilbert schemes in dimension three, preprint, math.AG 9803120; Topology 39 (2000) 1155–1191. CMP 2001:01
[13]
Y. Ito and M. Reid, The McKay correspondence for finite subgroups of , in Higher dimensional complex varieties (Trento, 1994), M. Andreatta et al., Eds., de Gruyter, 1996, pp. 221–240. MR 98i:14018
Reference [14]
D. Kaledin, The McKay correspondence for symplectic quotient singularities, preprint, math.AG 9907087.
Reference [15]
M. Kapranov and E. Vasserot, Kleinian singularities, derived categories and Hall algebras, preprint, math.AG 9812016; Math. Ann. 316 (2000) 565–576. CMP 2000:11
Reference [16]
S. Mac Lane, Categories for the working mathematician, Graduate Texts in Math. 5, Springer, 1971. MR 50:7275
Reference [17]
S. Mukai, Duality between and with its application to Picard sheaves, Nagoya Math. J. 81 (1981) 153–175. MR 82f:14036
Reference [18]
I. Nakamura, Hilbert schemes of Abelian group orbits, to appear in J. Alg. Geom.
Reference [19]
A. Neeman, The Grothendieck duality theorem via Bousfield’s techniques and Brown representability, J. Amer. Math. Soc. 9 (1996) 205–236. MR 96c:18006
Reference [20]
M. Reid, McKay correspondence, in Proc. of algebraic geometry symposium (Kinosaki, Nov 1996), T. Katsura (Ed.), 14–41, alg-geom 9702016.
Reference [21]
S.-S. Roan, Minimal resolutions of Gorenstein orbifolds in dimension 3, Topology 35 (1996) 489–508. MR 97c:14013
Reference [22]
P. Roberts, Intersection theorems, in Commutative algebra (Berkeley, 1987), MSRI Publ. 15, Springer, 1989, pp. 417–436. MR 90j:13024
Reference [23]
P. Roberts, Multiplicities and Chern classes in local algebra, CUP, 1998. MR 2001a:13029
Reference [24]
M. Verbitsky, Holomorphic symplectic geometry and orbifold singularities, preprint, math.AG 9903175; Asian J. Math. 4 (2000) 553–563. CMP 2001:05
Reference [25]
J.-L. Verdier, Des catégories dérivées des catégories abéliennes, Astérisque 239 (1996). MR 98c:18007

Article Information

MSC 2000
Primary: 14E15 (Global theory and resolution of singularities), 14J30 (-folds)
Secondary: 18E30 (Derived categories, triangulated categories), 19L47 (Equivariant -theory)
Keywords
  • Quotient singularities
  • McKay correspondence
  • derived categories
Author Information
Tom Bridgeland
Department of Mathematics and Statistics, University of Edinburgh, King’s Buildings, Mayfield Road, Edinburgh EH9 3JZ, United Kingdom
tab@maths.ed.ac.uk
ORCID
MathSciNet
Alastair King
Department of Mathematical Sciences, University of Bath, Bath BA2 7AY, United Kingdom
a.d.king@maths.bath.ac.uk
Miles Reid
Math Institute, University of Warwick, Coventry CV4 7AL, United Kingdom
miles@maths.warwick.ac.uk
Additional Notes

Earlier versions of this paper carried the additional title “Mukai implies McKay”.

Journal Information
Journal of the American Mathematical Society, Volume 14, Issue 3, 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 2001 American Mathematical Society
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/S0894-0347-01-00368-X
  • MathSciNet Review: 1824990
  • Show rawAMSref \bib{1824990}{article}{ author={Bridgeland, Tom}, author={King, Alastair}, author={Reid, Miles}, title={The McKay correspondence as an equivalence of derived categories}, journal={J. Amer. Math. Soc.}, volume={14}, number={3}, date={2001-07}, pages={535-554}, issn={0894-0347}, review={1824990}, doi={10.1090/S0894-0347-01-00368-X}, }

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.