Spherical twists, relations and the center of autoequivalence groups of K3 surfaces
Abstract.
Homological mirror symmetry predicts that there is a relation between autoequivalence groups of derived categories of coherent sheaves on Calabi–Yau varieties, and the symplectic mapping class groups of symplectic manifolds. In this paper, as an analogue of Dehn twists for closed oriented real surfaces, we study spherical twists for dg-enhanced triangulated categories.
We introduce the intersection number and relate it to group-theoretic properties of spherical twists. We show an inequality analogous to a fundamental inequality in the theory of mapping class groups about the behavior of the intersection number via iterations of Dehn twists. We also classify the subgroups generated by two spherical twists using the intersection number. In passing, we prove a structure theorem for finite dimensional dg-modules over the graded dual numbers and use this to describe the autoequivalence group.
As an application, we compute the center of autoequivalence groups of derived categories of K3 surfaces.
1. Introduction
Let be a smooth projective variety over a field and the bounded derived categories of coherent sheaves on . The autoequivalence group consisting of exact self-equivalences of is an interesting object in group theory. There are some attempts to compute , but this problem is rather difficult in general. The aim of this paper is to study group structures of by focusing on spherical twists. The details are explained in the following two subsections.
1.1. Spherical twists and intersection number
Homological mirror symmetry predicts that there is a relation between autoequivalence groups of derived categories of coherent sheaves on Calabi–Yau varieties, and the symplectic mapping class groups of symplectic manifolds. As an analogue of Dehn twists along Lagrangian spheres, Seidel–Thomas introduced (-)spherical objects and an autoequivalence called the spherical twist along a spherical object ([ST01]). In the following, we consider a more general triangulated category with a dg-enhancement, see Section 2 for detailed settings.
To clarify the analogy, we denote the sum of dimensions of all extension groups by i.e. for ,
which we call the intersection number of and in this paper. Referring to some well-known facts about Dehn twists for real surfaces (see Section 3), we prove the following three theorems. The intersection number is essential to understand the statements and their proofs.
The first result is the inequality describing the behavior of the intersection number via iterations of spherical twists.
Theorem 1.1 (Theorem 4.1 and cf. Theorem 3.1).
Let be a spherical object and objects. For any , we have
In the mapping class groups of real surfaces, a mapping class commutes with a Dehn twist along a simple closed curve if and only if it preserves the curve up to isotopy. In autoequivalence groups, this basic fact corresponds to the following.
Theorem 1.2 (Theorem 4.6 and cf. Theorem 3.2).
Let be spherical objects, an autoequivalence and . Then the following are equivalent:
-
.
-
for some , and .
As a corollary, we show a one-to-one correspondence between spherical objects and spherical twists (Corollary 4.8 (i)).
It is well-known that two Dehn twists whose intersection number is greater than one generate the (non-abelian) free group of rank 2. We prove the corresponding result for autoequivalence groups.
1.2. The center of autoequivalence groups of K3 surfaces
Let be a complex algebraic K3 surface and the subgroup of autoequivalences trivially acting on the transcendental lattice of . The autoequivalence groups are usually studied by using the action on the cohomology. In contrast to the automorphism groups of K3 surfaces, there are non-trivial autoequivalences trivially acting on the cohomology: squares of spherical twists for example. These cohomologically trivial autoequivalences are detected by the action on the space of stability conditions, which is Bridgeland’s approach in [Bri08]. Then we naturally consider the subgroups and which preserve the distinguished component of (see Definition 7.6).
The center (Definition 7.1) measures the commutativity of a given group. The triviality of the center of the mapping class group is proved via the equivalence between the commutativity with Dehn twists and the fixability of simple closed curves (Theorem 3.2). Similarly, as an application of Theorem 1.2, we compute the center groups and .
Theorem 1.4 (Theorem 8.1).
Let be a K3 surface of any Picard rank, the order of the finite cyclic group , and a generator of . Then we have the following
-
.
-
1.3. Related works
We give some comments on related works with Theorem 1.1 and Theorem 1.3. Historically, in the case of mapping class groups, the proof for the freeness (Theorem 3.3) of a subgroup generated by two Dehn twists was published by Ishida in 1996 ([Ish96]). Then ping-pong lemma and the inequality (Theorem 3.1) about the intersection number were key to their proof. In a similar manner, for the -graded Fukaya (-)categories of exact symplectic manifolds with contact type boundary, Keating proved the inequality ([Kea14, Proposition 7.4]) analogous to Theorem 3.1, and the freeness ([Kea14, Theorem 1.1]) for Dehn twists along Lagrangian spheres by ping-pong lemma. For algebraic triangulated categories, Volkov also proved similar results: the inequality ([Vol22, Lemma 3.3]) under technical assumptions and the freeness ([Vol22, Theorem 2.7(4)]). As an another approach to the freeness, in his thesis ([Kim18, Theorem 4.4]), Jongmyeong Kim proved that general spherical twists whose intersection number is greater than one respectively, generate the free group of rank under the formality assumption for some dg-algebra obtained by spherical objects. He translates and reformulates Humphries’ argument into a categorical setting. Anschütz also proved the freeness in a very special case: is a CY2-category and in his master thesis ([Ans13, Theorem 1.2]).
In this paper, We use different techniques to prove the inequality (Theorem 1.1) and the freeness (Theorem 1.3). In our approach expanded in Section 2 6, we do not need the extra assumptions to prove the inequality. The inequality itself is important for some applications. Actually, the inequality implies Theorem 1.2 and this theorem implies the one-to-one correspondence between the group and intersection number (Corollary 5.3, 5.4 and 5.5), and is used in the computations of the center of autoequivalence groups of K3 surfaces in Section 8. Moreover, we treat the power of spherical twists throughout this paper. This is important in future applications. The formulation of this paper clarifies the analogy between the autoequivalence groups and the mapping class groups. We also obtain a description of modules over the graded dual numbers (Proposition 6.7) and autoequivalences of its derived categories (Appendix A).
Acknowledgements. F.B. was supported by the European Research Council (ERC) under the European Union Horizon 2020 research and innovation programme (grant agreement No.725010). K.K. is supported by JSPS KAKENHI Grant Number 20K22310 and 21K13780.
Notation and Convention. Let be a field.
-
•
For a -linear triangulated category , an autoequivalence of is a -linear exact self-equivalence . The autoequivalence group is the group of (natural isomorphism classes) of autoequivalences of .
-
•
For a smooth projective variety over , the category is the derived category of bounded complexes of coherent sheaves on .
-
•
A K3 surfaces means a complex algebraic K3 surface i.e. a smooth projective surface over the complex number field such that and . Let be the automorphism group of .
2. Preliminaries on dg-categories
For the convenience of the reader, and to set up the notation for dg-categories and related notions, we now briefly recall the notions we will need. For a more detailed treatment of dg-categories and their derived categories, the reader is referred to [AL17, § 2.1].
2.1. Dg-categories and dg-modules
Let us fix a field . We will write for the category whose objects are couples where is a -graded vector space over and is a -linear endomorphism such that for any , and . The morphism is called the differential. We will often denote simply by , and we will call it a -dg-vector space.
Given two -dg-vector spaces and , morphisms between them are given by
(2.1) |
where is a -linear morphism such that for any . The graded -vector space (2.1) can be endowed with the differential
for . Above we introduced the notation , whose right hand side describes the components of with respect to the direct sum decomposition in (2.1).
The above discussion implies that is naturally enriched over itself. Even more is true: carries a monoidal structure for which the hom space (2.1) is the internal hom. The tensor product of and is defined as
with differential .
We can now introduce the notion of a dg-category: a dg-category over is a small category enriched111We require the composition maps to be closed, degree zero morphisms in . in . Given two dg-categories and , a dg-functor is a functor whose induced maps on morphism spaces preserve the degree and the differential.
Notice that to any dg-category we can attach a category, called its homotopy category, which is denoted by and whose objects are the same as those of , but for any we have .
We will be mainly interested in dg-modules over a fixed dg-category . An -dg-module is dg-functor . Given an -dg-module , we write for the image of via .
We write for the dg-category of dg-modules over , and given two -dg-modules and we write for the -dg-vector space of morphisms222The elements of are graded natural transformations, see [AL17, § 2.1.1]. of -dg-modules between and . Furthermore, we write for the derived category of dg-modules over , and for the subcategory of compact objects.
We write for the full subcategory of whose objects are given by h-projective dg-modules i.e. those modules such that
for any such that is an acyclic -dg-vector space for any .
As in the general theory of modules over rings, any dg-module is quasi-isomorphic333Quasi-isomorphisms of -dg-modules are defined fiberwise. Namely, a morphism of -dg-modules is a quasi-isomorphism if is so for any . to an h-projective -dg-module, and we can use h-projective dg-modules both to compute morphisms in the derived category and to derive tensor products of dg-modules, see [AL17, § 2.1.4] for the definition of the latter.
In the following, we write for the subcategory of those h-projective dg-modules that are compact in the derived category. Furthermore, we write .
2.2. Dg-cones
The category is a triangulated category with shift functor defined fiberwise i.e. given we have for any .
The operation of taking cones in can be strictified to define an operation in . Namely, given and a closed, degree zero morphism of -dg-modules, we define a dg-cone of as a dg-module endowed with morphisms of degree zero
such that
It is easy to see that the dg-cone is uniquely defined up to a closed, degree zero isomorphism in , and that for any the dg-module with diffential is a dg-cone of .
Remark 2.1.
Given and a closed degree zero morphism of -dg-modules, the dg-cone is isomorphic, as an element in , to the cone of . Moreover, we have the distinguished triangle
(2.2) |
2.3. Twisted complexes
A notion that will be fundamental for us is that of a (one-sided) twisted complex. For a more detailed treatment of twisted complexes, the reader is referred to444A remark is in order. In both the given references, only the notion of a twisted complex with a finite number of non-zero terms is considered. However, the theory still applies for infinite twisted complexes. [BK89] or [AL17].
A one-sided twisted complex over is a collection of -dg-modules , , and morphisms , , of degree such that
for any .
We adopt the following conventions:
-
the one-sided twisted complex given by the modules and the morphisms will be denoted by
-
if only for , and for , then we denote the one-sided twisted complex by
One-sided twisted complexes can be packaged into a dg-category. Given two twisted complexes and morphisms of degree between them are given by
(2.3) |
Given a morphism of twisted complexes, the differential of is defined as follows. Let us write for the component belonging to . Then, its differential is given by
(2.4) |
where is the differential on morphisms in .
We adopt the following convention to define a morphism of twisted complexes: we write and we mean that is the morphism of twisted complexes whose component in is given by . With this convention, if a component is not specified, it means that it is zero.
2.4. Convolution of twisted complexes
Given a one-sided twisted complex , we define its convolution as the -dg-module with differential
The operation of convolution defines a dg-functor from the category of twisted complexes to that of -dg-modules, see [AL17, § 3.2]. In particular, a closed degree zero morphism between twisted complexes induces a closed degree zero morphism between the respective convolutions.
In § 6 we will convolve various one-sided twisted complexes, and we will need to define elements belonging to such convolutions. For this reason, we introduce the following notation: if is a one-sided twisted complex and is its convolution, then we will write
for the element whose component in is given by , .
2.5. Spherical twists
From now on we assume that is proper i.e. that for any the -dg-vector space has finite dimensional total cohomology. Furthermore, we restrict our attention to those proper dg-categories such that has a Serre functor .
Definition 2.2.
For , an object is -spherical if it satisfies and
Let be a spherical object. Then the functor
is a spherical functor in the sense of [AL17, Section 5]. The twist functor associated to is called the spherical twist along .
Example 2.3.
For a spherical object and any object , the object fits into an exact triangle
(2.5) |
where and .
The following is well-known.
Lemma 2.4 (cf. [Huy06, Chapter 8]).
Let be a spherical object. We have the following.
-
.
-
for any autoequivalence .
We note that, if , any spherical twist is of infinite order by Lemma 2.4 (i).
3. Mapping class groups and Dehn twists
Dehn twists are fundamental elements in mapping class groups. In this section, we recall some well-known facts on Dehn twists, whose analogous statements are considered in the subsequent sections.
Let be a connected oriented closed surface of genus , and be the mapping class groups of . For an isotopy class of a simple closed curve in , the Dehn twist along is denoted by .
The following is a fundamental inequality about the behavior of the intersection number via iterations of Dehn twists.
Theorem 3.1 (cf. [FM12, Proposition 3.4]).
Let be isotopy classes of simple closed curves in . Then for any ,
(3.1) |
This inequality implies the following.
Theorem 3.2 (cf. [FM12, §3.3 and Fact 3.8]).
Let be isotopy classes of simple closed curves in , a mapping class and . Then the following are equivalent:
-
.
-
and .
4. Intersection number
We consider group-theoretic properties of spherical twists via the intersection number introduced in this section.
Throughout this section, we suppose and is a derived category of a proper dg-category such that has a Serre functor .
Definition 4.1.
For any two objects , the intersection number of and is defined by
where .
For a -spherical object , the condition implies for any .
Example 4.2.
Let be a K3 surface and distinct -curves. Then are 2-spherical for all and
where is the intersection number of and on .
The first main result is a complete analogue of the inequality (3.1) in Theorem 3.1 as follows. The proof is given in Section 6.
Theorem 4.3.
Let be a -spherical object and objects such that and . For any , we have
(4.1) |
We then consider the analogue of Theorem 3.2.
Proposition 4.4 ([Kim18, Lemma B.3] and [Kea14, Section 4.4]).
Let be -spherical objects. Then the following are equivalent:
-
There is no integer such that .
-
The composition map does not hit the identity for every .
-
The composition maps and vanish for all .
Two objects are called distinct if there is no integer such that (cf. Proposition 4.4(i)).
Lemma 4.5.
Let be distinct -spherical objects. Then there exists an object such that
Proof.
When (resp. ), it suffices to set (resp. ).
Let us consider the case of . By shifting, we may assume that for some . Following the arguments in the proof of [Kim18, Proposition 5.1], we define
and let be the cone of the natural evaluation map i.e. is an exact triangle in . Applying to this triangle, we see that is surjective, and are surjective for all . Moreover by Proposition 4.4 (ii), is zero. For , we have
and for ,
where we have (resp. ) if (resp. ). Direct computations give . It therefore suffices to set . ∎
The following is the analogue of Theorem 3.2.
Theorem 4.6.
Let be -spherical objects, an autoequivalence and . Then the following are equivalent:
-
.
-
for some , and .
Proof.
By Lemma 2.4 (ii), it is enough to show the case . The direction from (ii) to (i) follows from [ST01, Proposition 2.6] and .
We now consider the converse. Suppose that and are distinct. When , we have
by and , hence .
When , there exists such that . By , we may assume that . Then, by [Kim18, Proposition 5.1 and Remark 5.3] and , there exists an object satisfying and . The inequality (4.1) in Theorem 4.3 implies that
We thus have .
Finally, since each spherical twist is of infinite order, we have . ∎
We also obtain that powers of spherical twists are not shifts.
Proposition 4.7.
Suppose that has at least two distinct -spherical objects. Then each spherical twist satisfies for all and .
Proof.
The claim follows from the same argument of Theorem 4.6 and . ∎
By Theorem 4.6 and its proof, we can prove the following.
Corollary 4.8.
Let be -spherical objects.
-
and are distinct if and only if .
-
and are conjugate in for some if and only if for some .
5. Subgroups generated by two spherical twists
We consider the analogue of Theorem 3.3 and relate the intersection number to presentations of subgroups generated by two spherical twists.
Throughout this section, we suppose and is a derived category of a proper dg-category such that has a Serre functor .
Lemma 5.1 (Ping-pong lemma).
Let be a group acting on a set , and elements of G. Suppose that there are non-empty, disjoint subsets of with the property that, for each , we have for every nonzero integer . Then the subgroup generated by and is isomorphic to
Using the ping-pong lemma, we prove the main result in this section.
Theorem 5.2.
Let be distinct -spherical objects. If , then for any .
Proof.
To apply the ping-pong lemma, we define the subsets of the set of isomorphism classes of objects in as follows:
These are obviously disjoint, and non-empty by Lemma 4.5. By the ping-pong lemma, it suffices to check that and for each . We only show the former inclusion.
As a corollary, we reveal the relationship between the intersection number and presentations of subgroups generated by two spherical twists.
Corollary 5.3.
Let be distinct -spherical objects. Then the following are equivalent:
-
, where is the braid group on 3 strands.
-
for some
-
.
Proof.
The assertions are shown by Seidel–Thomas ([ST01, Proposition 2.13 and Theorem 2.17]) and Nordskova–Volkov ([NV19, Theorem 1]) for more general -spherical objects, see also [Huy06, Proposition 8.22]. The braid relation gives Theorem 4.6 then implies for some .
The inequality
follows from the inequality (4.1) and easy computations. Applying , we have
hence this inequality holds only in the case of or . When , we have by the braid relation and the commutative relation: , which contradicts the distinctness. Assume that . Then the subgroup is isomorphic to the rank 2 free group by Theorem 5.2, which contradicts the braid relation. We therefore have . ∎
Corollary 5.4.
Let be distinct spherical objects. Then the following are equivalent:
-
-
for some
-
.
Proof.
The assertions hold since Lemma 2.4 (ii) implies the commutative relation. Clearly, (i) implies (ii) by Theorem 4.6.
The inequality
follows from the inequality (4.1) and easy computations. Applying , we have
hence this inequality holds only in the case of or .
Corollary 5.5.
Let be distinct spherical objects. Then the following are equivalent:
-
-
.
6. Proof of Theorem 4.3
In this section, we prove Theorem 4.3.
Throughout this section, we consider the case of , and let be a proper dg-category such that has a Serre functor .
6.1. Good representatives
The following Proposition 6.1 is one of the technical steps towards the proof of Theorem 4.3 and it is an analogue of [Kea14, Lemma 3.1]. The difference is that in [Kea14] Keating can deform the category so that for a Lagrangian sphere she has , , on the nose. We cannot achieve this because we are working in the more strict formalism of dg-categories, but the following proposition will be enough for our purposes.
To simplify the notation, we will write for the dg-algebra of graded dual numbers with and zero differential.
The rationale behind Proposition 6.1 is that, given a -spherical object , as there exists a unique structure of graded algebra on
Namely, we have
This implies, as is intrinsically formal, see e.g. [KS22, Proposition 2.2], that the dg-endomorphism-algebra of an h-projective resolution of is quasi-isomorphic to . What we prove in Proposition 6.1 is that we can choose the h-projective resolution of so that it carries and -dg-module structure.
Proposition 6.1.
Let be a d-spherical object. Then, there exists with the following properties:
-
in
-
there exists such that
-
under the isomorphism of (i), corresponds to the canonical extension .
The proof of the above proposition is quite long and technical, for this reason we split it into various parts.
First, let us fix some notation. In the following, we write for the canonical extension of with itself, and we write for a fixed h-projective resolution of together with a fixed quasi-isomorphism
(6.1) |
We begin by proving the following
Lemma 6.2.
With the notation as above, the sequence of -dg-modules and morphisms
can be lifted to a one-sided infinite twisted complex in .
Proof.
The components of the sought twisted complexes will be given by for . Then, to prove the statement of the lemma it is enough to find morphisms
of degree , , such that the following diagram commutes for every
(6.2) |
and such that
(6.3) |
Indeed, then the sought twisted complex will be given by , .
In the following, when we say that induces the morphism via the quasi-isomorphism (6.1), we mean that the diagram (6.2) commutes.
We will prove the existence of the morphisms by induction on . First of all, notice that if we have all the maps for , then to define the maps with it is enough to define . Indeed, if we have , then we can set
(6.4) |
for , and using (6.3) for we have555In the above equations, we make use of the fact that for any morphism and any it holds that .
We now begin the inductive construction. For the case we use that by the definition of an h-projective resolution we have
and therefore we can define by lifting along the previous isomorphism. With this choice and the definition (6.4), we get that induces via the quasi-isomorphism (6.1), as we wanted.
Now assume that we defined for any . We claim that to construct it is enough to prove that is a closed morphism. Indeed, notice that has degree as a morphism , and therefore it corresponds to a morphism of degree . The cohomology of is concentrated in degree and because is -spherical, and therefore a closed element can be non-trivial in cohomology if and only if it has degree or . Now, as we are in the case , we have either
or , depending on whether or . In either case, the degree of is not or , and therefore if is closed it must be the differential of some morphism of degree that we can take to be our .
We now prove that is closed. We have
where in passing from the second to the third line we used that and satisfy (6.3) by the induction hypothesis. Our aim is to prove that the above terms sum to zero. Take a decomposition of in three steps: . To get the term we have two possibilities: either from or . In the first case we get the term , in the second case , and they cancel out.
Hence, given for , we can define , and the inductive step is complete. Thus, the proof of the lemma is complete. ∎
The next step in the proof of Proposition 6.1 is the following
Lemma 6.3.
Let us write for the convolution of the twisted complex (6.5). Then, there exists a closed morphism of degree such that the dg-cone of is -h-projective.
Proof.
We will define a morphism at the level of the twisted complex (6.5), and then convolve it to a morphism .
Following the convention introduced in § 2.3, we define the morphism of twisted complexes
where . By definition, has degree . We now prove that is a closed morphism. As per the definition given in § 2.3, the differential of is given by
To prove that is closed, we fix and we focus our attention to the components of that map to . These are given by
(6.6) |
Using the definition of given in (6.4), we see that (6.6) is equal to the shift by of
Here the last equality follows from the fact that, given any , the first term contributes with , while the second term contributes with . Using (6.4), one sees that these two terms cancel out.
Hence, is a closed morphism of twisted complexes of degree , and we define as its convolution.
Let us write for the dg-cone of as defined in § 2.2.
We now show that is -h-projective. It is enough to prove that is -h-projective, and this is what we show. The fact that is h-projective follows from the fact that it is the convolution of a one-sided twisted complex whose components are -h-projective, and they are non-zero only in negative degree. Indeed, these properties imply that has an exhaustive filtration by h-projective -dg-modules, see also [Sta18, Tag 09KK], and thus it is -h-projective. ∎
In the lemma below, we show that in Lemma 6.3 is an -h-projective resolution of .
Lemma 6.4.
The dg-cone of the morphism constructed in Lemma 6.3 is quasi-isomorphic to .
Proof.
To prove the lemma, we construct a closed, degree zero morphism such that we have a distinguished triangle
thus proving that in .
We define as the convolution of a closed, degree zero morphism of twisted complexes , where is the twisted complex with in position zero. The degree zero morphism is defined as
Let us show that is closed. We fix , and we focus our attention on the component of starting from . By definition, this is given by
where the last equality follows from the fact that the satisfy (6.3). Hence, is a closed morphism of degree zero, and upon convolution (and a shift by ) it induces a morphism .
To conclude the proof of the lemma, we construct morphisms , , and such that
(6.7) |
realise as the dg-cone of .
The morphisms and are defined as the inclusion of into and the projection onto , respectively. Such maps exists because at the level of underlying graded modules is the direct sum .
The morphism is defined as the convolution of the morphism defined by
for .
The fact that the morphisms in (6.7) realise as the dg-cone of is an easy check, and we leave it to the reader.
We can now complete the proof by noticing that by Remark 2.1 we have the distinguished triangle
and therefore we have the quasi-isomorphisms
where we wrote for the morphism , where the second morphism is the one given by the definition of a dg-cone. ∎
We are now in the position to prove Proposition 6.1.
Proof (Proof of Proposition 6.1).
Lemma 6.4 constructs for us an -h-projective resolution of . This will be the one whose existence is claimed in the statement of Proposition 6.1.
To conclude the proof of Proposition 6.1, we only have to construct the morphism . We define using the direct sum decomposition of . Namely, as a graded module is given by , and when we write we mean that and in this decomposition. Then, is given by
It is clear that is a closed, degree morphism such that .
We now prove that under the quasi-isomorphisms the morphism corresponds to the canonical extension of . By the definition of , this is equivalent to prove that under the quasi-isomorphism the morphism corresponds to . We prove the latter statement.
From now on, every time we have a spherical object we implicitly assume we replaced it with as in Proposition 6.1. Hence, every spherical object is really and there is such that .
Let be a -spherical object. The following proposition is an analogue of [Kea14, Proposition 6.3] (recall the convention § 2.3 (ii)).
Proposition 6.5.
For any the object , is the convolution of the twisted complex
(6.8) |
where .
Proof.
If we had , we could replicate the proof of [Kea14, Proposition 6.3]. We now explain how to reduce to this case. Notice that the inclusion of the subalgebra is a quasi-isomorphism. Therefore, in the twisted complex representing we can replace with everywhere. The new twisted complex is of the same form as the one of [Kea14, Proposition 6.3], so we can apply that proof and get the result. ∎
Let us write for the convolution of the twisted complex
(6.9) |
Then, we have
Proposition 6.6.
Let and . Then, is isomorphic to the convolution of the twisted complex
where .
Proof.
The claim follows from the definition of and [AL17, Lemma 3.4]. ∎
6.2. Graded dual numbers
Recall that we write for the dg-algebra of the graded dual numbers with . The aim of this subsection is to classify finite dimensional dg-modules over . We write for the triangulated subcategory of given by -dg-modules with finite dimensional total cohomology.
To prove Proposition 6.7 below, we will use Koszul duality. Namely, let us define as the dg-algebra with and . Then, Koszul duality says that we have an equivalence sending to . For a recent proof of this statement, see [KS22].
Having means that to prove a structure theorem for modules in it is enough to study the structure of modules in , and this is what we do.
Before moving on to the proposition, let us notice a useful consequence of the equivalence . By definition, we have for any , and the unique non-trivial extension is given by the convolution of the twisted complex
(6.10) |
which is quasi-isomorphic . As is an equivalence sending to , we have
and therefore there is a unique non-trivial extension of by itself of degree in for any . This extension is given by the convolution of the twisted complex
(6.11) |
As is an equivalence, we get that is isomorphic, up to a shift that can be computed to be , to for any .
We write , , for the convolution of (6.11), and . Similarly, we write for the convolution of (6.11) with replaced by , and . The following proposition gives the desired classification (cf. [Kea14, Proposition 5.3]).
Proposition 6.7.
Let be a finite dimensional right (resp. left) -dg-module. Then, is quasi-isomorphic to a finite direct sum of copies of shifts of ’s (resp. ’s). Moreover, is compact if and only if (resp. ) does not appear.
Remark 6.8.
The above statement pairs up with [KS22, Proposition 2.2] to show why is generated by the as a triangulated category (without additional idempotent completion).
Proof.
We prove the statement for right modules; this will suffice because is commutative.
Let us fix and such that . Notice that if we forget the grading, then is a PID. Hence, the cohomology of the dg-module splits as a finite direct sum
(6.12) |
for some .
Let us write for the convolution of (6.10). Then, notice that is a free resolution of . Hence, replacing with in (6.12), we can lift the isomorphism (6.12) to a quasi-isomorphism666In slightly more detail: if appears in (6.12), then it means that there exists such that and for some . Then, we define by sending to and , respectively. Here we employed the notation we introduced in § 2.4 for elements belonging to the convolution of a twisted complex.
(6.13) |
Applying to (6.13) and using the isomorphisms and , we obtain the sought decomposition
Finally, to prove the claim about the compactness of recall that is closed under taking direct summands and notice that of the ’s the only non-compact one is (its derived endomorphism algebra is infinite dimensional, see e.g. [KS22, Lemma 3.4]). ∎
6.3. Proof
We are now ready to prove Theorem 4.3. We only need one last definition:
Definition 6.9.
Given two dg-modules over the graded dual numbers and , we define as the convolution of the twisted complex
where .
Remark 6.10.
The construction is independent of the quasi-isomorphism classes of and . Indeed, this follows easily from the definition of and e.g. [AL21, Corollary 2.12].
We state Theorem 4.3 again:
Theorem 6.11.
Let be a spherical object and objects such that and . For any , we have
(6.14) |
Proof.
By replacing by , it suffices to prove the claim for . The case is proved by direct computations and using , so we can assume .
By the definition of , see (6.9), we have the following distinguished triangle
Applying to this distinguished triangle, as and intersection numbers are subadditive on distinguished triangles, we see that it is enough to prove
By Proposition 6.6 we know that
Moreover, by Remark 6.10 we know that we can replace and by quasi-isomorphic -dg-modules. As and , by Proposition 6.7 we know that and split as direct sums of shifts of and ’s, respectively ’s, for . Hence, to conclude it is enough to prove that
(6.15) |
for .
If , then is a direct sum of shifts of , and therefore (6.15) becomes , which is true.
If and , then the right hand side of (6.15) is equal to . Hence, we have to find two non-zero cohomology classes in . We claim that the sought classes as given by
where we employed the notation we introduced in § 2.4 for elements belonging to the convolution of a twisted complex.
It is obvious that is closed. For , it follows from the fact that and are closed in and , respectively, and the definition of the differential in :
We prove that gives a non-zero cohomology class, the proof for is analogous. The element is a non-zero cohomology class in . Hence, if is the differential of some element, it must be the differential of an element of the form with . The differential of such an element is given by (we can assume we always have on the left because the tensor products are -linear)
(6.16) |
For (6.16) to be equal to , we must have
However, by the definition of , the equation is not satisfied by any , and therefore is a non-zero cohomology class. Hence, we get , as we wanted.
If and , the situation is as in the previous point.
If and , the right hand side of (6.15) is equal to 4. The four elements which give rise to non-zero cohomology classes are given by
We have exhausted all the possible cases for (6.15), and therefore we have concluded the proof of the theorem. ∎
Remark 6.12.
As mentioned in §1.3, using different techniques, a particular case of the above theorem was proved by Volkov in [Vol22, Lemma 3.3].
7. Preliminaries on K3 surfaces
In this section, we prepare basic properties of the autoequivalence groups of derived categories of K3 surfaces for the computations of the center groups in Section 8.
Let be a K3 surface.
7.1. Hodge structures on Mukai lattices
The integral cohomology group of has the lattice structure given by the Mukai pairing
for . The lattice called the Mukai lattice of is an even unimodular lattice of signature . The Mukai lattice has a weight two Hodge structure given by and
This Hodge structure contains the ordinary Hodge structure on as a primitive sub-Hodge structure. The algebraic part of denoted by , is equal to and has signature .
For an object , the Mukai vector of is given by
By the Riemann–Roch formula, we have the isomorphism satisfying for any objects , where is the numerical Grothendieck group of and is the Euler pairing on it.
7.2. Groups
A Hodge isometry of is an isomorphism of the Hodge structure preserving the Mukai pairing. The group of Hodge isometries is denoted by . Let be the transcendental lattice of which is the transcendental part of the Hodge structures and . Restricting the Hodge structure with the Mukai pairing to sub-Hodge structures and , we similarly define the groups of Hodge isometries and , respectively. Then is a finite cyclic group, and faithfully acts on by a root of unity.
Using the action of on , the following two groups are defined
These two subgroups are normal. The group is of finite index, and its element is called a symplectic automorphism. The natural group homomorphisms
are injective, thus is also a finite cyclic group.
For any autoequivalence , we define the cohomological Fourier–Mukai transform associated to by
which is a Hodge isometry of , thus induces the action
The two subgroups and of are defined by
These two subgroups are normal, and clearly . The group is of finite index, and its element is called a Calabi–Yau autoequivalence. We note that is Calabi–Yau if and only if it respects the Serre duality pairings
induced by a choice of holomorphic volume forms in , see [BB17, Appendix A].
We recall the definition of centralizer and center groups.
Definition 7.1.
Let be a group, and a subgroup of .
-
The centralizer group of in is defined by
-
The center group of is defined by
It is easy to see , but these are not equal in general.
7.3. Stability conditions on
We review the space of Bridgeland stability conditions on the derived categories of K3 surfaces and the action on it of the autoequivalence group.
Let be a K3 surface and fix a norm on .
Definition 7.2 ([Bri07, Definition 5.1]).
A (numerical) stability condition on consists of a group homomorphism called central charge and a family of full additive subcategory of called slicing, such that
-
For , we have for some .
-
For all , we have .
-
For and , we have .
-
For each , there is a collection of exact triangles called Harder–Narasimhan filtration of :
(7.1) with and .
-
(support property) There exists a constant such that for all , we have
(7.2)
For any interval , define to be the extension-closed subcategory of generated by the subcategories for . Then is the heart of a bounded t-structure on , hence an abelian category. The full subcategory is also shown to be abelian. A non-zero object is called -semistable of phase , and especially a simple object in is called -stable. Taking the Harder–Narasimhan filtration (7.1) of , we define and . The object is called -semistable factor of . Define to be the set of numerical stability conditions on .
We prepare some terminologies on the stability on the heart of a -structure on .
Definition 7.3.
Let be the heart of a bounded -structure on . A stability function on is a group homomorphism such that for all , the complex number lies in the semiclosed upper half plane .
Given a stability function on , the phase of an object is defined to be . An object is -semistable (resp. -stable) if for all subobjects , we have (resp. ). We say that a stability function satisfies the Harder–Narasimhan property if each object admits a filtration (called Harder–Narasimhan filtration of ) such that is -semistable for with , and the support property if there exists a constant such that for all -semistable objects , we have .
The following proposition shows the relationship between stability conditions and stability functions on the heart of a bounded -structure.
Proposition 7.4 ([Bri07, Proposition 5.3]).
To give a stability condition on is equivalent to giving the heart of a bounded t-structure on , and a stability function on with the Harder–Narasimhan property and the support property.
For the proof, we construct the slicing , from the pair , by
and extend for all by . Conversely, for a stability condition , the heart is given by . We also denote stability conditions by .
We recall that the Mukai vector and the Mukai pairing on , then the central charge of a numerical stability condition takes the form for some . Bridgeland constructed a family of stability conditions on as follows: Let
For , we set . The category is the heart of a t-structure obtained by tilting the standard t-structure with respect to the torsion pair on given by
where for a torsion free sheaf , (resp. ) is the maximal (resp. minimal) slope of -semistable factors of . Then is a stability condition on ([Bri08, Lemma 6.2 and Proposition 11.2]).
Let be a non-zero object of and be a stability condition on . The mass of is defined by
where are -semistable factors of . The following generalized metric (i.e. with values in ) on is defined by Bridgeland ([Bri07, Proposition 8.1]):
This generalized metric induces the topology on . Then the generalized metric takes a finite value on each connected component of , thus is a metric space in the strict sense.
Theorem 7.5 ([Bri07, Theorem 7.1]).
The map
(7.3) |
is a local homeomorphism, where is equipped with the natural linear topology.
Therefore the space (and each connected component ) naturally admits a structure of finite dimensional complex manifolds.
There is a left action of on given by
(7.4) |
This action of is isometric with respect to .
Let be the connected component of containing the set of geometric stability conditions i.e. one for which all structure sheaves of points are stable of the same phase. It is easy to check that the above stability condition for each is geometric.
Definition 7.6.
The group is defined as the subgroup of which preserves the connected component .
Proposition 7.7 ([Har12, proof of Proposition 7.9]).
The following autoequivalences preserve the distinguished component i.e. are elements in :
-
•
shift for
-
•
line bundle tensor for
-
•
pullback for
-
•
The composition
where is a -dimensional fine compact moduli space of Gieseker-stable torsion free sheaves on with the universal family such that .
-
•
spherical twist along a Gieseker-stable spherical bundle (e.g. ).
-
•
spherical twist along for any -curve on
We additionally define the following groups
Define the open subset consisting of vectors whose real and imaginary parts span a positive definite 2-plane in . This subset has two connected components, distinguished by the orientation induced on this 2-plane; let be the component containing vectors of the form for an ample class . Consider the root system
consisting of -classes in , and the corresponding hyperplane complement
Theorem 7.8 ([Bri08, Theorem 1.1]).
The map
is a normal covering, and the group is identified with the group of deck transformations of .
The following is a Bridgeland conjecture on the space of stability conditions on and the action on it of .
Conjecture 7.9 ([Bri08, Conjecture 1.2]).
Let be a K3 surface. Then
-
is simply-connected.
-
any autoequivalence of preserves the distinguished component , equivalently we have
This conjecture clearly implies an isomorphism
More strongly, Bayer–Bridgeland proved the contractibility of in [BB17].
In Section 8, we consider the center groups , , and the centralizer group .
8. The center of autoequivalence groups of K3 surfaces
Let be a K3 surface of any Picard rank.
8.1. Main results
We compute the center groups of and and the centralizer groups .
Theorem 8.1.
Let be a K3 surface, the order of , and a generator of . Then we have the following
-
.
-
Proof.
-
We note that and , see Proposition 7.7. Fix any . By Theorem 4.6, the relation implies for some . For any line bundle on , we have by . By similar arguments as in the proof of [Huy12, Lemma A.2], is in , hence . It remains to show that . Fix any and set
Since the transcendental part of the Hodge structure is equal to , is of finite index, so that . By [Huy12, Lemma A.3], acts on trivially, hence also does. We thus have since is isomorphic to the group of deck transformations of the normal cover , see Theorem 7.8. We therefore have , thus .
By , similar arguments give
The other inclusion follows from
-
By (i), we have
Each element of is of the form for some and . Then is equal to if and only if (a) is odd and , or (b) is even and i.e. . By the faithfulness of the action of on , the case (a) is realized only when is even and . We therefore have
which completes the proof.
∎
When Conjecture 7.9 (i) and (ii) are true, the group is naturally isomorphic to the orbifold fundamental group of the stringy Kähler moduli space of ([BB17, Section 7] and [Huy16, Conjecture 3.14]), so this quotient group is important in the context of mirror symmetry.
Corollary 8.2.
Let be a K3 surface. Then we have
Proof.
Let be a natural group homomorphism induced by the quotient. By Theorem 8.1(ii), whether is even or not. We note that if and only if any satisfies for some .
8.2. Examples
We here collect several examples of the order of the finite cyclic group , which determines the center groups by Theorem 8.1 and Corollary 8.2.
Example 8.3.
Let be a K3 surface of odd Picard rank. Then is isomorphic to (cf. [Huy16, Cor 3.3.5]). Therefore (resp. ) if and only if (resp. ).
Example 8.4.
Example 8.5.
Let be a K3 surface of Picard rank 2, with infinite automorphism group. Then by [GLP10, Corollary 1], is isomorphic to or . Since is finite and , one has .
Example 8.6.
Example 8.7.
Let be a K3 surface with an unimodular or 2-elementary transcendental lattice. Then there exists an involution satisfying and , hence is even.
Example 8.8.
Let be a K3 surface with , where is the Euler function. Then is in the set
and is uniquely determined by due to Kondo ([Kon92, Main Theorem]), Vorontsov ([Vor83]), Machida–Oguiso ([MO98, Theorem 3]) and Oguiso–Zhang ([OZ00, Theorem 2]). Especially, is even (resp. odd) if and only if is unimodular (resp. non-unimodular).
Appendix A Fully faithful functors and the autoequivalence group for graded dual numbers
Using Proposition 6.7, we can describe the autoequivalence groups and for the dg-algebra of the graded dual numbers .
Recall that, for a morphism of dg-algebras , the functor is defined as tensor product with seen as an bimodule. For let be the morphism of dg-algebras defined by sending to .
The following theorem is a generalisation of [AM15, Corollay 5.2, 5.11] to the graded setting.
Theorem A.1.
Every fully faithful endofunctor of is an autoequivalence, and for every there exists and such that .
Proof.
If is fully faithful, then it must induce an isomorphism of graded algebras . As is compact, by Proposition 6.7 we know that it decomposes as a direct sum of shifts of ’s for . It is easy to check that such a direct sum has endomorphism algebra equal to if and only if there is just one summand and it is for some . Hence, . Now notice that the only maps of graded algebras (which coincide with the only dg-algebra maps from to itself) are the ’s. Hence, is an endofunctor of that acts as the identity of . Given Proposition 6.7, this implies that , and therefore , which is an autoequivalence, and the first claim follows.
For the second claim notice that if then it induces an autoequivalence of , and we can repeat the argument above to conclude that for some . ∎
References
- [AL17] R. Anno and T. Logvinenko. Spherical DG-functors. J. Eur. Math. Soc. (JEMS), 19(9):2577–2656, 2017.
- [AL21] R. Anno and T. Logvinenko. Bar category of modules and homotopy adjunction for tensor functors. Int. Math. Res. Not. IMRN, (2):1353–1462, 2021.
- [AM15] F. Amodeo and R. Moschetti. Fourier–Mukai functors and perfect complexes on dual numbers. J. Algebra, 437:133–160, 2015.
- [Ans13] Johannes Anschütz. Autoequivalences of derived categories of local K3 surfaces. master thesis, University of Bonn, 2013.
- [BB17] A. Bayer and T. Bridgeland. Derived automorphism groups of K3 surfaces of Picard rank 1. Duke Math. J., 166(1):75–124, 2017.
- [BK89] A. I. Bondal and M. M. Kapranov. Representable functors, serre functors, and reconstructions. Izv. Akad. Nauk SSSR Ser. Mat., 53(6):1183–1205, 1989.
- [Bri07] T. Bridgeland. Stability conditions on triangulated categories. Ann. of Math., 166:317–345, 2007.
- [Bri08] T. Bridgeland. Stability conditions on K3 surfaces. Duke Math. J., 141(2):241–291, 2008.
- [DHKK14] G. Dimitrov, F. Haiden, L. Katzarkov, and M. Kontsevich. Dynamical systems and categories. Contemporary Mathematics, 621:133–170, 2014.
- [FM12] B. Farb and D. Margalit. A primer on mapping class groups. volume 49 of Princeton Mathematical Series. Princeton University Press, Princeton, 2012.
- [GLP10] F. Galluzzi, G. Lombardo, and C. Peters. Automorphisms of indefinite binary quadratic forms and K3-surfaces with Picard number 2. Rend. Sem. Mat. Univ. Politec. Torino, 68(1):57–77, 2010.
- [Har12] H. Hartmann. Cusps of the Kähler moduli space and stability conditions on K3 surfaces. Math. Ann., 354(1):1–42, 2012.
- [Huy06] D. Huybrechts. Fourier–Mukai transforms in algebraic geometry. Oxford Mathematical Monographs, The Clarendon Press. Oxford University Press, Oxford, 2006.
- [Huy12] D. Huybrechts. Stability conditions via spherical objects. Math. Z., 271(3-4):1253–1270, 2012.
- [Huy16] D. Huybrechts. Lectures on K3 surfaces. Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 2016.
- [Ish96] A. Ishida. The structure of subgroup of mapping class groups generated by two Dehn twists. Proc. Japan Acad. Ser. A Math. Sci., 72(10):240–241, 1996.
- [Kea14] A. Keating. Dehn twists and free subgroups of symplectic mapping class groups. J. Topology, 7(2):436–474, 2014.
- [Kim18] J. Kim. A freeness criterion for spherical twists. PhD thesis, Nagoya University, 2018.
- [Kon92] S. Kondo. Automorphisms of algebraic K3 surfaces which act trivially on picard groups. J. Math. Soc. Japan., 44:75–98, 1992.
- [KS22] A. Kuznetsov and E. Shinder. Categorical absorptions of singularities and degenerations. arXiv:2207.06477, 2022.
- [KST20] K. Kikuta, Y. Shiraishi, and A. Takahashi. A note on entropy of auto-equivalences: lower bound and the case of orbifold projective lines. Nagoya Math. J., 238:86–103, 2020.
- [MO98] N. Machida and K. Oguiso. On K3 surfaces admitting finite non-symplectic group actions. J. Math. Sci. Univ. Tokyo, 5(2):273–297, 1998.
- [NV19] A. Nordskova and Y. Volkov. Faithful actions of braid groups by twists along ade-configurations of spherical objects. arXiv.1910.02401, 2019.
- [OZ00] K. Oguiso and D.-Q. Zhang. On Vorontsov’s theorem on K3 surfaces with non-symplectic group actions. Proc. Amer. Math. Soc., 128(6):1571–1580, 2000.
- [SI77] T. Shioda and H. Inose. On singular K3 surfaces. In Complex analysis and algebraic geometry, 119–136. Iwanami Shoten, Tokyo, 1977.
- [ST01] P. Seidel and R. Thomas. Braid group actions on derived categories of coherent sheaves. Duke Math. J., 108(1):37–108, 2001.
- [Sta18] The Stacks Project Authors. Stacks Project, https://stacks.math.columbia.edu, 2018.
- [Vin83] È. Vinberg. The two most algebraic K3 surfaces. Math. Ann., 265(1):1–21, 1983.
- [Vol22] Y. Volkov. Groups generated by two twists along spherical sequences. Mathematical Proceedings of the Cambridge Philosophical Society, online:1–25, 2022.
- [Vor83] S. P. Vorontsov. Automorphisms of even lattices that arise in connection with automorphisms of algebraic K3 surfaces. Vestnik Mosk. Univ. Math., 38:19–21, 1983.