This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Spherical twists, relations and the center of autoequivalence groups of K3 surfaces

Federico Barbacovi Department of Mathematics, University College London, 25 Gordon Street, London, WC1H 0AY, UK federico.barbacovi.18@ucl.ac.uk  and  Kohei Kikuta Department of Mathematics, Graduate School of Science, Osaka University, Toyonaka Osaka, 560-0043, Japan kikuta@math.sci.osaka-u.ac.jp
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 XX be a smooth projective variety over a field KK and 𝒟b(X){\mathcal{D}}^{b}(X) the bounded derived categories of coherent sheaves on XX. The autoequivalence group Aut(𝒟b(X)){\rm Aut}({\mathcal{D}}^{b}(X)) consisting of exact self-equivalences of 𝒟b(X){\mathcal{D}}^{b}(X) is an interesting object in group theory. There are some attempts to compute Aut(𝒟b(X)){\rm Aut}({\mathcal{D}}^{b}(X)), but this problem is rather difficult in general. The aim of this paper is to study group structures of Aut(𝒟b(X)){\rm Aut}({\mathcal{D}}^{b}(X)) 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 (dd-)spherical objects and an autoequivalence TEAut(𝒟b(X))T_{E}\in{\rm Aut}({\mathcal{D}}^{b}(X)) called the spherical twist along a spherical object E𝒟b(X)E\in{\mathcal{D}}^{b}(X) ([ST01]). In the following, we consider a more general triangulated category 𝒟{\mathcal{D}} 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(M,N)i(M,N) i.e. for M,N𝒟M,N\in{\mathcal{D}},

i(M,N):=pdimKHom𝒟(M,N[p]),i(M,N):=\sum_{p\in{\mathbb{Z}}}\dim_{K}{\rm Hom}_{\mathcal{D}}(M,N[p]),

which we call the intersection number of MM and NN 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 E𝒟E\in{\mathcal{D}} be a spherical object and M,N𝒟M,N\in{\mathcal{D}} objects. For any k\{0}k\in{\mathbb{Z}}\backslash\{0\}, we have

i(E,M)i(E,N)i(TEkM,N)+i(M,N).i(E,M)i(E,N)\leq i(T^{k}_{E}M,N)+i(M,N).

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 E1,E2𝒟E_{1},E_{2}\in{\mathcal{D}} be spherical objects, ΦAut(𝒟)\Phi\in{\rm Aut}({\mathcal{D}}) an autoequivalence and k1,k2\{0}k_{1},k_{2}\in{\mathbb{Z}}\backslash\{0\}. Then the following are equivalent:

  1. (i)(\rm{i})

    ΦTE1k1=TE2k2Φ\Phi\circ T_{E_{1}}^{k_{1}}=T_{E_{2}}^{k_{2}}\circ\Phi.

  2. (ii)(\rm{ii})

    Φ(E1)=E2[l]\Phi(E_{1})=E_{2}[l] for some ll\in{\mathbb{Z}}, and k1=k2k_{1}=k_{2}.

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.

Theorem 1.3 (Theorem 5.2 and cf. Theorem 3.3).

Let E1,E2𝒟E_{1},E_{2}\in{\mathcal{D}} be spherical objects satisfying E1≄E2[l]E_{1}\not\simeq E_{2}[l] for any ll\in{\mathbb{Z}}. If i(E1,E2)2i(E_{1},E_{2})\geq 2, then TE1k1,TE2k2\langle T_{E_{1}}^{k_{1}},T_{E_{2}}^{k_{2}}\rangle is isomorphic to the free group F2F_{2} of rank 22 for any k1,k2\{0}k_{1},k_{2}\in{\mathbb{Z}}\backslash\{0\}.

We use the ping-pong lemma for the proof. This approach is the same as the case of mapping class groups. As a corollary, we reveal the relationship between the intersection number and presentations of subgroups generated by two spherical twists (Corollary 5.3, 5.4 and 5.5).

1.2. The center of autoequivalence groups of K3 surfaces

Let XX be a complex algebraic K3 surface and AutCY(𝒟b(X)){\rm Aut}_{{\rm CY}}({\mathcal{D}}^{b}(X)) the subgroup of autoequivalences trivially acting on the transcendental lattice of XX. 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 Stab(X){\rm Stab}(X) of stability conditions, which is Bridgeland’s approach in [Bri08]. Then we naturally consider the subgroups Aut(𝒟b(X)){\rm Aut}^{\dagger}({\mathcal{D}}^{b}(X)) and AutCY(𝒟b(X)){\rm Aut}^{\dagger}_{{\rm CY}}({\mathcal{D}}^{b}(X)) which preserve the distinguished component of Stab(X){\rm Stab}(X) (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 Z(Aut(𝒟b(X)))Z({\rm Aut}^{\dagger}({\mathcal{D}}^{b}(X))) and Z(AutCY(𝒟b(X)))Z({\rm Aut}^{\dagger}_{{\rm CY}}({\mathcal{D}}^{b}(X))).

Theorem 1.4 (Theorem 8.1).

Let XX be a K3 surface of any Picard rank, mXm_{X} the order of the finite cyclic group Autt(X):={fAut(X)|H2(f)|NS(X)=idNS(X)}{\rm Aut}_{t}(X):=\left\{f\in{\rm Aut}(X)~{}\middle|~{}H^{2}(f)|_{{\rm NS}(X)}=\mathrm{id}_{{\rm NS}(X)}\right\}, and ftf_{t} a generator of Autt(X){\rm Aut}_{t}(X). Then we have the following

  1. (i)(\rm{i})

    Z(Aut(𝒟b(X)))=Autt(X)×[1](/mX)×Z({\rm Aut}^{\dagger}({\mathcal{D}}^{b}(X)))={\rm Aut}_{t}(X)\times{\mathbb{Z}}[1]\simeq({\mathbb{Z}}/m_{X})\times{\mathbb{Z}}.

  2. (ii)(\rm{ii})
    Z(AutCY(𝒟b(X)))\displaystyle Z({\rm Aut}^{\dagger}_{{\rm CY}}({\mathcal{D}}^{b}(X))) =\displaystyle= {(ft)mX/2[1]if mX is even[2]if mX is odd\displaystyle\begin{cases}\langle~{}(f_{t}^{*})^{m_{X}/2}\circ[1]~{}\rangle&\mbox{if }m_{X}\mbox{ is even}\\ {\mathbb{Z}}[2]&\mbox{if }m_{X}\mbox{ is odd}\end{cases}
    \displaystyle\simeq .\displaystyle{\mathbb{Z}}.

We also compute the center of the quotient AutCY(𝒟b(X))/[2]{\rm Aut}^{\dagger}_{{\rm CY}}({\mathcal{D}}^{b}(X))/{\mathbb{Z}}[2] (Corollary 8.2), which is closely related to the orbifold fundamental group of the stringy Kähler moduli space of XX. These results reveal that the number mXm_{X} determines the group structure of the center groups, so we explain some examples of mXm_{X} in §8.2.

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 /2{\mathbb{Z}}/2-graded Fukaya (AA_{\infty}-)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 nn spherical twists whose intersection number is greater than one respectively, generate the free group of rank nn 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: 𝒟{\mathcal{D}} is a CY2-category and i(E1,E2)=2i(E_{1},E_{2})=2 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 \sim 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 TE1,TE2\langle T_{E_{1}},T_{E_{2}}\rangle and intersection number i(E1,E2)i(E_{1},E_{2}) (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 KK be a field.

  • For a KK-linear triangulated category 𝒟{\mathcal{D}}, an autoequivalence of 𝒟{\mathcal{D}} is a KK-linear exact self-equivalence 𝒟𝒟{\mathcal{D}}\to{\mathcal{D}}. The autoequivalence group Aut(𝒟){\rm Aut}({\mathcal{D}}) is the group of (natural isomorphism classes) of autoequivalences of 𝒟{\mathcal{D}}.

  • For a smooth projective variety XX over KK, the category 𝒟b(X):=𝒟b(Coh(X)){\mathcal{D}}^{b}(X):={\mathcal{D}}^{b}({\rm Coh}(X)) is the derived category of bounded complexes of coherent sheaves on XX.

  • A K3 surfaces XX means a complex algebraic K3 surface i.e. a smooth projective surface over the complex number field {\mathbb{C}} such that ωX𝒪X\omega_{X}\simeq\mathcal{O}_{X} and H1(X,𝒪X)=0H^{1}(X,\mathcal{O}_{X})=0. Let Aut(X){\rm Aut}(X) be the automorphism group of XX.

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 KK. We will write Mod-K\textbf{Mod}\text{-}K for the category whose objects are couples (V,dV)(V,d_{V}) where V=nVnV=\oplus_{n\in\mathbb{Z}}V_{n} is a \mathbb{Z}-graded vector space over KK and dV:VVd_{V}\colon V\rightarrow V is a KK-linear endomorphism such that dV(Vn)Vn+1d_{V}(V_{n})\subset V_{n+1} for any nn\in\mathbb{Z}, and dV2=0d_{V}^{2}=0. The morphism dVd_{V} is called the differential. We will often denote (V,dV)(V,d_{V}) simply by VV, and we will call it a KK-dg-vector space.

Given two KK-dg-vector spaces VV and WW, morphisms between them are given by

HomMod-K(V,W)=nHomMod-Kn(V,W){\rm Hom}_{\textbf{Mod}\text{-}K}(V,W)=\bigoplus_{n\in\mathbb{Z}}{\rm Hom}^{n}_{\textbf{Mod}\text{-}K}(V,W) (2.1)

where fnHomMod-n(V,W)f_{n}\in{\rm Hom}^{n}_{\textbf{Mod}\text{-}}(V,W) is a KK-linear morphism f:VWf\colon V\rightarrow W such that fn(Vm)Wm+nf_{n}(V_{m})\subset W_{m+n} for any mm\in\mathbb{Z}. The graded KK-vector space (2.1) can be endowed with the differential

df=d({fn})={dWfn(1)nfndV}df=d(\{f_{n}\})=\{d_{W}\circ f_{n}-(-1)^{n}f_{n}\circ d_{V}\}

for fHomMod-K(V,W)f\in{\rm Hom}_{\textbf{Mod}\text{-}K}(V,W). Above we introduced the notation f={fn}f=\{f_{n}\}, whose right hand side describes the components of ff with respect to the direct sum decomposition in (2.1).

The above discussion implies that Mod-K\textbf{Mod}\text{-}K is naturally enriched over itself. Even more is true: Mod-K\textbf{Mod}\text{-}K carries a monoidal structure for which the hom space (2.1) is the internal hom. The tensor product of VV and WW is defined as

VKW=n(i+j=nVnKWm)V\otimes_{K}W=\bigoplus_{n\in\mathbb{Z}}\left(\bigoplus_{i+j=n}V_{n}\otimes_{K}W_{m}\right)

with differential dVKW=dVid+iddWd_{V\otimes_{K}W}=d_{V}\otimes\mathrm{id}+\mathrm{id}\otimes d_{W}.

We can now introduce the notion of a dg-category: a dg-category over KK is a small category 𝒜{\mathcal{A}} enriched111We require the composition maps to be closed, degree zero morphisms in Mod-K\textbf{Mod}\text{-}K. in Mod-K\textbf{Mod}\text{-}K. Given two dg-categories 𝒜{\mathcal{A}} and {\mathcal{B}}, a dg-functor Φ:𝒜\Phi\colon{\mathcal{A}}\rightarrow{\mathcal{B}} is a functor whose induced maps on morphism spaces preserve the degree and the differential.

Notice that to any dg-category 𝒜{\mathcal{A}} we can attach a category, called its homotopy category, which is denoted by H0(𝒜)H^{0}({\mathcal{A}}) and whose objects are the same as those of 𝒜{\mathcal{A}}, but for any a1,a2𝒜a_{1},a_{2}\in{\mathcal{A}} we have HomH0(𝒜)(a1,a2):=H0(Hom𝒜(a1,a2)){\rm Hom}_{H^{0}({\mathcal{A}})}(a_{1},a_{2}):=H^{0}({\rm Hom}_{{\mathcal{A}}}(a_{1},a_{2})).

We will be mainly interested in dg-modules over a fixed dg-category 𝒜{\mathcal{A}}. An 𝒜{\mathcal{A}}-dg-module is dg-functor 𝒜opMod-K{\mathcal{A}}^{\mathrm{op}}\rightarrow\textbf{Mod}\text{-}K. Given an 𝒜{\mathcal{A}}-dg-module MM, we write MaM_{a} for the image of a𝒜a\in{\mathcal{A}} via MM.

We write Mod-𝒜\textbf{Mod}\text{-}{\mathcal{A}} for the dg-category of dg-modules over 𝒜{\mathcal{A}}, and given two 𝒜{\mathcal{A}}-dg-modules MM and NN we write Hom𝒜(M,N)\mathrm{Hom}_{{\mathcal{A}}}(M,N) for the KK-dg-vector space of morphisms222The elements of Hom𝒜(M,N)\mathrm{Hom}_{{\mathcal{A}}}(M,N) are graded natural transformations, see [AL17, § 2.1.1]. of 𝒜{\mathcal{A}}-dg-modules between MM and NN. Furthermore, we write 𝒟(𝒜){\mathcal{D}}({\mathcal{A}}) for the derived category of dg-modules over 𝒜{\mathcal{A}}, and 𝒟(𝒜)c𝒟(𝒜){\mathcal{D}}({\mathcal{A}})^{c}\subset{\mathcal{D}}({\mathcal{A}}) for the subcategory of compact objects.

We write 𝒫(𝒜){\mathcal{P}}({\mathcal{A}}) for the full subcategory of Mod-𝒜\textbf{Mod}\text{-}{\mathcal{A}} whose objects are given by h-projective dg-modules i.e. those modules PMod-𝒜P\in\textbf{Mod}\text{-}{\mathcal{A}} such that

HomH0(Mod-𝒜)(P,S)=0\mathrm{Hom}_{H^{0}(\textbf{Mod}\text{-}{\mathcal{A}})}(P,S)=0

for any SMod-𝒜S\in\textbf{Mod}\text{-}{\mathcal{A}} such that SaS_{a} is an acyclic KK-dg-vector space for any a𝒜a\in{\mathcal{A}}.

As in the general theory of modules over rings, any dg-module is quasi-isomorphic333Quasi-isomorphisms of 𝒜{\mathcal{A}}-dg-modules are defined fiberwise. Namely, a morphism f:MNf\colon M\rightarrow N of 𝒜{\mathcal{A}}-dg-modules is a quasi-isomorphism if fa:MaNaf_{a}\colon M_{a}\rightarrow N_{a} is so for any a𝒜a\in{\mathcal{A}}. to an h-projective 𝒜{\mathcal{A}}-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 𝒫(𝒜)c{\mathcal{P}}({\mathcal{A}})^{c} for the subcategory of those h-projective dg-modules that are compact in the derived category. Furthermore, we write =K\otimes=\otimes_{K}.

2.2. Dg-cones

The category 𝒟(𝒜){\mathcal{D}}({\mathcal{A}}) is a triangulated category with shift functor defined fiberwise i.e. given M𝒟(𝒜)M\in{\mathcal{D}}({\mathcal{A}}) we have (M[1])a=Ma[1](M[1])_{a}=M_{a}[1] for any a𝒜a\in{\mathcal{A}}.

The operation of taking cones in 𝒟(𝒜){\mathcal{D}}({\mathcal{A}}) can be strictified to define an operation in Mod-𝒜\textbf{Mod}\text{-}{\mathcal{A}}. Namely, given M,NMod-𝒜M,N\in\textbf{Mod}\text{-}{\mathcal{A}} and f:MNf\colon M\rightarrow N a closed, degree zero morphism of 𝒜{\mathcal{A}}-dg-modules, we define a dg-cone of ff as a dg-module CMod-𝒜C\in\textbf{Mod}\text{-}{\mathcal{A}} endowed with morphisms of degree zero

M[1]𝑖C𝑝M[1]N𝑗C𝑞N\begin{array}[]{lcr}M[1]\xrightarrow{i}C\xrightarrow{p}M[1]&&N\xrightarrow{j}C\xrightarrow{q}N\end{array}

such that

pi=idM[1]qj=idNip+jq=idCdp=dj=0di=jfdq=fp\begin{array}[]{cccccc}pi=\mathrm{id}_{M[1]}&qj=\mathrm{id}_{N}&ip+jq=\mathrm{id}_{C}&dp=dj=0&di=jf&dq=-fp\end{array}

It is easy to see that the dg-cone is uniquely defined up to a closed, degree zero isomorphism in Mod-𝒜\textbf{Mod}\text{-}{\mathcal{A}}, and that for any ff the dg-module C(f):=M[1]NC(f):=M[1]\oplus N with diffential d(m,n)=(dm,dnf(n))d(m,n)=(dm,dn-f(n)) is a dg-cone of ff.

Remark 2.1.

Given M,NMod-𝒜M,N\in\textbf{Mod}\text{-}{\mathcal{A}} and f:MNf\colon M\rightarrow N a closed degree zero morphism of 𝒜{\mathcal{A}}-dg-modules, the dg-cone C(f)C(f) is isomorphic, as an element in 𝒟(𝒜){\mathcal{D}}({\mathcal{A}}), to the cone of ff. Moreover, we have the distinguished triangle

M𝑓N𝑗C(f)𝑝M[1].M\xrightarrow{f}N\xrightarrow{j}C(f)\xrightarrow{p}M[1]. (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 𝒜{\mathcal{A}} is a collection of 𝒜{\mathcal{A}}-dg-modules EiE_{i}, ii\in\mathbb{Z}, and morphisms αij:EiEj\alpha_{ij}\colon E_{i}\rightarrow E_{j}, i<ji<j, of degree ij+1i-j+1 such that

(1)jdαij+i<k<jαkjαik=0(-1)^{j}d\alpha_{ij}+\sum_{i<k<j}\alpha_{kj}\circ\alpha_{ik}=0

for any i<ji<j.

We adopt the following conventions:

  1. (i)(\rm{i})

    the one-sided twisted complex given by the modules EiE_{i} and the morphisms αij\alpha_{ij} will be denoted by (Ei,αij)(E_{i},\alpha_{ij})

  2. (ii)(\rm{ii})

    if Ei0E_{i}\neq 0 only for i=n,,0i=-n,\dots,0, and αij=0\alpha_{ij}=0 for ji>1j-i>1, then we denote the one-sided twisted complex (Ei,αij)(E_{i},\alpha_{ij}) by

    Enαn(n+1)En+1α(n+1)(n+2)α10E0E_{-n}\xrightarrow{\alpha_{-n(-n+1)}}E_{-n+1}\xrightarrow{\alpha_{(-n+1)(-n+2)}}\dots\xrightarrow{\alpha_{-10}}E_{0}

One-sided twisted complexes can be packaged into a dg-category. Given two twisted complexes (Ei,αij)(E_{i},\alpha_{ij}) and (Fi,βij)(F_{i},\beta_{ij}) morphisms of degree pp between them are given by

p=q+lkHom𝒜q(Ek,Fl).\bigoplus_{p=q+l-k}{\rm Hom}_{{\mathcal{A}}}^{q}(E_{k},F_{l}). (2.3)

Given γ:(Ei,αij)(Fi,βij)\gamma\colon(E_{i},\alpha_{ij})\rightarrow(F_{i},\beta_{ij}) a morphism of twisted complexes, the differential of γ\gamma is defined as follows. Let us write γqkl:EkFl\gamma_{q}^{kl}\colon E_{k}\rightarrow F_{l} for the component γ\gamma belonging to Hom𝒜q(Ek,Fl){\rm Hom}_{{\mathcal{A}}}^{q}(E_{k},F_{l}). Then, its differential is given by

d(γqkl)=(1)ld𝒜(γqkl)+l<mβlmγqkl(1)q+lkm<kγqklαmkd(\gamma_{q}^{kl})=(-1)^{l}d_{{\mathcal{A}}}(\gamma_{q}^{kl})+\sum_{l<m}\beta_{lm}\circ\gamma_{q}^{kl}-(-1)^{q+l-k}\sum_{m<k}\gamma_{q}^{kl}\circ\alpha_{mk} (2.4)

where d𝒜d_{{\mathcal{A}}} is the differential on morphisms in 𝒜{\mathcal{A}}.

We adopt the following convention to define a morphism of twisted complexes: we write γ={γqkl:EkFl}\gamma=\{\gamma_{q}^{kl}\colon E_{k}\rightarrow F_{l}\} and we mean that γ\gamma is the morphism of twisted complexes whose component in Hom𝒜q(Ek,Fl){\rm Hom}_{{\mathcal{A}}}^{q}(E_{k},F_{l}) is given by γqkl\gamma_{q}^{kl}. 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 (Ei,αij)(E_{i},\alpha_{ij}), we define its convolution as the 𝒜{\mathcal{A}}-dg-module iEi[i]\oplus_{i\in\mathbb{Z}}E_{i}[-i] with differential

idEi[i]+i<jαij.\sum_{i\in\mathbb{Z}}d_{E_{i}[-i]}+\sum_{i<j}\alpha_{ij}.

The operation of convolution defines a dg-functor from the category of twisted complexes to that of 𝒜{\mathcal{A}}-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 (Ei,αij)(E_{i},\alpha_{ij}) is a one-sided twisted complex and iEi[i]\oplus_{i\in\mathbb{Z}}E_{i}[-i] is its convolution, then we will write

(,e3,e2,e1,e0,)(\dots,e_{-3},e_{-2},e_{-1},e_{0},\dots)

for the element eiEi[i]e\in\oplus_{i\in\mathbb{Z}}E_{i}[-i] whose component in Ei[i]E_{i}[-i] is given by eie_{i}, ii\in\mathbb{Z}.

2.5. Spherical twists

From now on we assume that 𝒜{\mathcal{A}} is proper i.e. that for any a1,a2𝒜a_{1},a_{2}\in{\mathcal{A}} the KK-dg-vector space Hom𝒜(a1,a2)\mathrm{Hom}_{{\mathcal{A}}}(a_{1},a_{2}) has finite dimensional total cohomology. Furthermore, we restrict our attention to those proper dg-categories such that 𝒟(𝒜)c{\mathcal{D}}({\mathcal{A}})^{c} has a Serre functor 𝒮{\mathcal{S}}.

Definition 2.2.

For dd\in\mathbb{Z}, an object E𝒟(𝒜)cE\in{\mathcal{D}}({\mathcal{A}})^{c} is dd-spherical if it satisfies 𝒮(E)E[d]\mathcal{S}(E)\simeq E[d] and

nHom𝒟(𝒜)(E,E[n])[n]kk[d].\bigoplus_{n\in\mathbb{Z}}\mathrm{Hom}_{{\mathcal{D}}({\mathcal{A}})}(E,E[n])[-n]\simeq k\oplus k[-d].

Let E𝒟(𝒜)cE\in{\mathcal{D}}({\mathcal{A}})^{c} be a spherical object. Then the functor

K𝕃E:𝒟(K)𝒟(𝒜)-\otimes^{{\mathbb{L}}}_{K}E:~{}{\mathcal{D}}(K)\to{\mathcal{D}}({\mathcal{A}})

is a spherical functor in the sense of [AL17, Section 5]. The twist functor TEAut(𝒟(𝒜))T_{E}\in{\rm Aut}({\mathcal{D}}({\mathcal{A}})) associated to K𝕃E-\otimes^{{\mathbb{L}}}_{K}E is called the spherical twist along EE.

Example 2.3.

When 𝒟(𝒜)c=𝒟b(X){\mathcal{D}}({\mathcal{A}})^{c}={\mathcal{D}}^{b}(X), the first condition in Definition 2.2 is clearly equivalent to EωXEE\otimes\omega_{X}\simeq E. For a spherical object E𝒟b(X)E\in{\mathcal{D}}^{b}(X), the spherical twist TEAut(𝒟b(X))T_{E}\in{\rm Aut}({\mathcal{D}}^{b}(X)) along EE is isomorphic to a Fourier–Mukai transform Φ𝒫E\Phi_{{\mathcal{P}}_{E}} whose kernel is the cone of the composition of the restriction to the diagonal Δ\Delta with the trace

𝒫E:=C(EE(EE)|Δtr𝒪Δ)𝒟b(X×X),{\mathcal{P}}_{E}:=C\left(E^{\vee}\boxtimes E\rightarrow(E^{\vee}\boxtimes E)|_{\Delta}\xrightarrow{tr}\mathcal{O}_{\Delta}\right)\in{\mathcal{D}}^{b}(X\times X),

see [Huy06, Section 8.1] for details.

For a spherical object E𝒟(𝒜)cE\in{\mathcal{D}}({\mathcal{A}})^{c} and any object M𝒟(𝒜)M\in{\mathcal{D}}({\mathcal{A}}), the object TE(M)T_{E}(M) fits into an exact triangle

Hom(E,M)KEevMTE(M),{\rm Hom}^{\bullet}(E,M)\otimes_{K}E\xrightarrow{ev}M\rightarrow T_{E}(M), (2.5)

where Hom(M,N):=pHomp(M,N)[p]{\rm Hom}^{\bullet}(M,N):=\bigoplus_{p\in{\mathbb{Z}}}{\rm Hom}^{p}(M,N)[-p] and Homp(M,N):=Hom𝒟(𝒜)(M,N[p])\mathrm{Hom}^{p}(M,N):=\mathrm{Hom}_{{\mathcal{D}}({\mathcal{A}})}(M,N[p]).

The following is well-known.

Lemma 2.4 (cf. [Huy06, Chapter 8]).

Let E𝒟(𝒜)cE\in{\mathcal{D}}({\mathcal{A}})^{c} be a spherical object. We have the following.

  1. (i)(\rm{i})

    TE(E)=E[1d]T_{E}(E)=E[1-d].

  2. (ii)(\rm{ii})

    ΦTEΦ1=TΦ(E)\Phi\circ T_{E}\circ\Phi^{-1}=T_{\Phi(E)} for any autoequivalence ΦAut(𝒟(𝒜)c)\Phi\in{\rm Aut}({\mathcal{D}}({\mathcal{A}})^{c}).

We note that, if d1d\neq 1, 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 Σg\Sigma_{g} be a connected oriented closed surface of genus g2g\geq 2, and MCG(Σg){\rm MCG}(\Sigma_{g}) be the mapping class groups of Σg\Sigma_{g}. For an isotopy class of a simple closed curve in Σg\Sigma_{g}, the Dehn twist along aa is denoted by TaMCG(Σg)T_{a}\in{\rm MCG}(\Sigma_{g}).

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 a,b,ca,b,c be isotopy classes of simple closed curves in Σg\Sigma_{g}. Then for any k\{0}k\in{\mathbb{Z}}\backslash\{0\},

i(a,b)i(a,c)i(Tak(c),b)+i(b,c).i(a,b)i(a,c)\leq i(T_{a}^{k}(c),b)+i(b,c). (3.1)

This inequality implies the following.

Theorem 3.2 (cf. [FM12, §3.3 and Fact 3.8]).

Let a1,a2a_{1},a_{2} be isotopy classes of simple closed curves in Σg\Sigma_{g}, ϕMCG(Σg)\phi\in{\rm MCG}(\Sigma_{g}) a mapping class and k1,k2\{0}k_{1},k_{2}\in{\mathbb{Z}}\backslash\{0\}. Then the following are equivalent:

  1. (i)(\rm{i})

    ϕTa1k1=Ta2k2ϕ\phi\circ T_{a_{1}}^{k_{1}}=T_{a_{2}}^{k_{2}}\circ\phi.

  2. (ii)(\rm{ii})

    ϕ(a1)=a2\phi(a_{1})=a_{2} and k1=k2k_{1}=k_{2}.

Theorem 3.3 ([Ish96, Theorem 1.2] and cf. [FM12, Theorem 3.14]).

Let a1,a2a_{1},a_{2} be two isotopy classes of simple closed curves in Σg\Sigma_{g}. If the intersection number i(a1,a2)2i(a_{1},a_{2})\geq 2, then Ta1k1,Ta2k2F2\langle T_{a_{1}}^{k_{1}},T_{a_{2}}^{k_{2}}\rangle\simeq F_{2} for any k1,k2\{0}k_{1},k_{2}\in{\mathbb{Z}}\backslash\{0\}, where F2F_{2} is the free group of rank 22.

4. Intersection number

We consider group-theoretic properties of spherical twists via the intersection number introduced in this section.

Throughout this section, we suppose d>1d>1 and 𝒟:=𝒟(𝒜){\mathcal{D}}:={\mathcal{D}}({\mathcal{A}}) is a derived category of a proper dg-category 𝒜{\mathcal{A}} such that 𝒟c:=𝒟(𝒜)c{\mathcal{D}}^{c}:={\mathcal{D}}({\mathcal{A}})^{c} has a Serre functor 𝒮{\mathcal{S}}.

Definition 4.1.

For any two objects M,N𝒟M,N\in{\mathcal{D}}, the intersection number i(M,N)i(M,N) of MM and NN is defined by

i(M,N):=phomp(M,N)0{+},i(M,N):=\sum_{p\in{\mathbb{Z}}}{\rm hom}^{p}(M,N)\in{\mathbb{Z}}_{\geq 0}\sqcup\{+\infty\},

where homp(M,N):=dimKHomp(M,N){\rm hom}^{p}(M,N):=\dim_{K}{\rm Hom}^{p}(M,N).

For a dd-spherical object E𝒟cE\in{\mathcal{D}}^{c}, the condition 𝒮(E)E[d]\mathcal{S}(E)\simeq E[d] implies i(E,M)=i(M,E)i(E,M)=i(M,E) for any M𝒟M\in{\mathcal{D}}.

Example 4.2.

Let XX be a K3 surface and C1,C2C_{1},C_{2} distinct (2)(-2)-curves. Then 𝒪C1(j1),𝒪C2(j2)𝒟b(X)\mathcal{O}_{C_{1}}(j_{1}),\mathcal{O}_{C_{2}}(j_{2})\in{\mathcal{D}}^{b}(X) are 2-spherical for all j1,j2j_{1},j_{2}\in{\mathbb{Z}} and

i(𝒪C1(j1),𝒪C2(j2))=hom1(𝒪C1(j1),𝒪C2(j2))=C1.C2,i(\mathcal{O}_{C_{1}}(j_{1}),\mathcal{O}_{C_{2}}(j_{2}))={\rm hom}^{1}(\mathcal{O}_{C_{1}}(j_{1}),\mathcal{O}_{C_{2}}(j_{2}))=C_{1}.C_{2},

where C1.C2C_{1}.C_{2} is the intersection number of C1C_{1} and C2C_{2} on H2(X,)H^{2}(X,{\mathbb{Z}}).

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 E𝒟cE\in{\mathcal{D}}^{c} be a dd-spherical object and M,N𝒟M,N\in{\mathcal{D}} objects such that i(E,M)i(E,M) and i(E,N)<i(E,N)<\infty. For any k\{0}k\in{\mathbb{Z}}\backslash\{0\}, we have

i(E,M)i(E,N)i(TEkM,N)+i(M,N).i(E,M)i(E,N)\leq i(T^{k}_{E}M,N)+i(M,N). (4.1)

We then consider the analogue of Theorem 3.2.

Proposition 4.4 ([Kim18, Lemma B.3] and [Kea14, Section 4.4]).

Let E1,E2𝒟cE_{1},E_{2}\in{\mathcal{D}}^{c} be dd-spherical objects. Then the following are equivalent:

  1. (i)(\rm{i})

    There is no integer ll\in{\mathbb{Z}} such that E1E2[l]E_{1}\simeq E_{2}[l].

  2. (ii)(\rm{ii})

    The composition map Hom(Ei,Ej)Hom(Ej,Ei)Hom(Ei,Ei){\rm Hom}^{\bullet}(E_{i},E_{j})\otimes{\rm Hom}^{\bullet}(E_{j},E_{i})\to{\rm Hom}^{\bullet}(E_{i},E_{i}) does not hit the identity idEi\mathrm{id}_{E_{i}} for every iji\neq j.

  3. (iii)(\rm{iii})

    The composition maps Homd(Ei,Ei)[d]Hom(Ei,Ej)Hom+d(Ei,Ej){\rm Hom}^{d}(E_{i},E_{i})[-d]\otimes{\rm Hom}^{\bullet}(E_{i},E_{j})\to{\rm Hom}^{\bullet+d}(E_{i},E_{j}) and Hom(Ei,Ej)Homd(Ej,Ej)[d]Hom+d(Ei,Ej){\rm Hom}^{\bullet}(E_{i},E_{j})\otimes{\rm Hom}^{d}(E_{j},E_{j})[-d]\to{\rm Hom}^{\bullet+d}(E_{i},E_{j}) vanish for all iji\neq j.

Two objects M,N𝒟M,N\in{\mathcal{D}} are called distinct if there is no integer ll\in{\mathbb{Z}} such that MN[l]M\simeq N[l] (cf. Proposition 4.4(i)).

Lemma 4.5.

Let E1,E2𝒟cE_{1},E_{2}\in{\mathcal{D}}^{c} be distinct dd-spherical objects. Then there exists an object S𝒟cS\in{\mathcal{D}}^{c} such that

i(E1,S)>i(E2,S).i(E_{1},S)>i(E_{2},S).
Proof.

When i(E1,E2)1i(E_{1},E_{2})\leq 1 (resp. i(E1,E2)3i(E_{1},E_{2})\geq 3), it suffices to set S:=E1S:=E_{1} (resp. S:=E2S:=E_{2}).

Let us consider the case of i(E1,E2)=2i(E_{1},E_{2})=2. By shifting, we may assume that i(E1,E2)=hom0(E1,E2)+homp(E1,E2)i(E_{1},E_{2})={\rm hom}^{0}(E_{1},E_{2})+{\rm hom}^{p}(E_{1},E_{2}) for some p0p\geq 0. Following the arguments in the proof of [Kim18, Proposition 5.1], we define

Z:=Homd(E1,E1)[d]E1Hom(E2,E1)E2𝒟c,Z:={\rm Hom}^{d}(E_{1},E_{1})[-d]\otimes E_{1}\oplus{\rm Hom}^{\bullet}(E_{2},E_{1})\otimes E_{2}\in{\mathcal{D}}^{c},

and let E1E^{\prime}_{1} be the cone of the natural evaluation map i.e. ZE1E1Z\to E_{1}\to E^{\prime}_{1} is an exact triangle in 𝒟c{\mathcal{D}}^{c}. Applying Hom(Ej,)(j=1,2){\rm Hom}(E_{j},-)~{}(j=1,2) to this triangle, we see that Homd(E1,Z)Homd(E1,E1){\rm Hom}^{d}(E_{1},Z)\to{\rm Hom}^{d}(E_{1},E_{1}) is surjective, and Homi(E2,Z)Homi(E2,E1){\rm Hom}^{i}(E_{2},Z)\to{\rm Hom}^{i}(E_{2},E_{1}) are surjective for all ii\in{\mathbb{Z}}. Moreover by Proposition 4.4 (ii), Hom0(E1,Z)Hom0(E1,E1){\rm Hom}^{0}(E_{1},Z)\to{\rm Hom}^{0}(E_{1},E_{1}) is zero. For p=0p=0, we have

homi(E1,Z)={5i=d1i=2d0otherwise homi(E2,Z)={2i=d4i=2d0otherwise,{\rm hom}^{i}(E_{1},Z)=\begin{cases}5&i=d\\ 1&i=2d\\ 0&otherwise\end{cases}~{}~{}\text{ }~{}~{}{\rm hom}^{i}(E_{2},Z)=\begin{cases}2&i=d\\ 4&i=2d\\ 0&otherwise\end{cases},

and for p>0p>0,

homi(E1,Z)={1i=dp3i=d1i=d+p1i=2d0otherwise homi(E2,Z)={1i=dp1i=d2i=2dp2i=2d0otherwise,{\rm hom}^{i}(E_{1},Z)=\begin{cases}1&i=d-p\\ 3&i=d\\ 1&i=d+p\\ 1&i=2d\\ 0&otherwise\end{cases}~{}~{}\text{ }~{}~{}{\rm hom}^{i}(E_{2},Z)=\begin{cases}1&i=d-p\\ 1&i=d\\ 2&i=2d-p\\ 2&i=2d\\ 0&otherwise\end{cases},

where we have homd+p(E1,Z)=2{\rm hom}^{d+p}(E_{1},Z)=2 (resp. homd(E2,Z)=3{\rm hom}^{d}(E_{2},Z)=3) if d+p=2dd+p=2d (resp. d=2dpd=2d-p). Direct computations give i(E1,E1)=6>4=i(E2,E1)i(E_{1},E^{\prime}_{1})=6>4=i(E_{2},E^{\prime}_{1}). It therefore suffices to set S:=E1S:=E^{\prime}_{1}. ∎

The following is the analogue of Theorem 3.2.

Theorem 4.6.

Let E1,E2𝒟cE_{1},E_{2}\in{\mathcal{D}}^{c} be dd-spherical objects, ΦAut(𝒟c)\Phi\in{\rm Aut}({\mathcal{D}}^{c}) an autoequivalence and k1,k2\{0}k_{1},k_{2}\in{\mathbb{Z}}\backslash\{0\}. Then the following are equivalent:

  1. (i)(\rm{i})

    ΦTE1k1=TE2k2Φ\Phi\circ T_{E_{1}}^{k_{1}}=T_{E_{2}}^{k_{2}}\circ\Phi.

  2. (ii)(\rm{ii})

    Φ(E1)=E2[l]\Phi(E_{1})=E_{2}[l] for some ll\in{\mathbb{Z}}, and k1=k2k_{1}=k_{2}.

Proof.

By Lemma 2.4 (ii), it is enough to show the case Φ=id𝒟b(X)\Phi={\rm id}_{{\mathcal{D}}^{b}(X)}. The direction from (ii) to (i) follows from [ST01, Proposition 2.6] and TE[1]=TET_{E[1]}=T_{E}.

We now consider the converse. Suppose that E1E_{1} and E2E_{2} are distinct. When i(E1,E2)=0i(E_{1},E_{2})=0, we have

TE1k1(E1)=E1[k1(1d)]E1=TE2k2(E1)T_{E_{1}}^{k_{1}}(E_{1})=E_{1}[k_{1}(1-d)]\neq E_{1}=T_{E_{2}}^{k_{2}}(E_{1})

by d>1d>1 and k10k_{1}\neq 0, hence TE1k1TE2k2T_{E_{1}}^{k_{1}}\neq T_{E_{2}}^{k_{2}}.

When i(E1,E2)=1i(E_{1},E_{2})=1, there exists p0p_{0}\in{\mathbb{Z}} such that i(E1,E2)=homp0(E1,E2)=1i(E_{1},E_{2})={\rm hom}^{p_{0}}(E_{1},E_{2})=1. By TE[1]=TET_{E[1]}=T_{E}, we may assume that p0=1p_{0}=1. Then, by [Kim18, Proposition 5.1 and Remark 5.3] and d>1d>1, there exists an object S𝒟S\in{\mathcal{D}} satisfying i(E1,S)=1i(E_{1},S)=1 and i(E2,S)=0i(E_{2},S)=0. The inequality (4.1) in Theorem 4.3 implies that

i(TE1k1E2,S)\displaystyle i(T_{E_{1}}^{k_{1}}E_{2},S) \displaystyle\geq i(E1,E2)i(E1,S)i(E2,S)\displaystyle i(E_{1},E_{2})i(E_{1},S)-i(E_{2},S)
=\displaystyle= i(E1,E2)\displaystyle i(E_{1},E_{2})
>\displaystyle> 0=i(TE2k2E2,S).\displaystyle 0=i(T_{E_{2}}^{k_{2}}E_{2},S).

We thus have TE1k1TE2k2T_{E_{1}}^{k_{1}}\neq T_{E_{2}}^{k_{2}}.

When i(E1,E2)2i(E_{1},E_{2})\geq 2, there exists S𝒟cS\in{\mathcal{D}}^{c} satisfying i(E1,S)>i(E2,S)i(E_{1},S)>i(E_{2},S) by Lemma 4.5. The inequality (4.1) implies that

i(TE1k1E2,S)\displaystyle i(T_{E_{1}}^{k_{1}}E_{2},S) \displaystyle\geq i(E1,E2)i(E1,S)i(E2,S)\displaystyle i(E_{1},E_{2})i(E_{1},S)-i(E_{2},S)
>\displaystyle> i(E1,E2)i(E2,S)i(E2,S)=(i(E1,E2)1)i(E2,S)\displaystyle i(E_{1},E_{2})i(E_{2},S)-i(E_{2},S)=(i(E_{1},E_{2})-1)i(E_{2},S)
\displaystyle\geq i(E2,S)=i(TE2k2E2,S).\displaystyle i(E_{2},S)=i(T_{E_{2}}^{k_{2}}E_{2},S).

Hence we have TE1k1TE2k2T_{E_{1}}^{k_{1}}\neq T_{E_{2}}^{k_{2}}.

Finally, since each spherical twist is of infinite order, we have k1=k2k_{1}=k_{2}. ∎

We also obtain that powers of spherical twists are not shifts.

Proposition 4.7.

Suppose that 𝒟c{\mathcal{D}}^{c} has at least two distinct dd-spherical objects. Then each spherical twist TEAut(𝒟c)T_{E}\in{\rm Aut}({\mathcal{D}}^{c}) satisfies TEk1[k2]T_{E}^{k_{1}}\neq[k_{2}] for all k1\{0}k_{1}\in{\mathbb{Z}}\backslash\{0\} and k2k_{2}\in{\mathbb{Z}}.

Proof.

The claim follows from the same argument of Theorem 4.6 and i(E,M)=i(E[k2],M)i(E,M)=i(E[k_{2}],M). ∎

By Theorem 4.6 and its proof, we can prove the following.

Corollary 4.8.

Let E1,E2𝒟cE_{1},E_{2}\in{\mathcal{D}}^{c} be dd-spherical objects.

  1. (i)(\rm{i})

    E1E_{1} and E2E_{2} are distinct if and only if TE1TE2T_{E_{1}}\neq T_{E_{2}}.

  2. (ii)(\rm{ii})

    TE1k1T_{E_{1}}^{k_{1}} and TE2k2T_{E_{2}}^{k_{2}} are conjugate in Aut(𝒟c){\rm Aut}({\mathcal{D}}^{c}) for some k1,k2\{0}k_{1},k_{2}\in{\mathbb{Z}}\backslash\{0\} if and only if Φ(E1)=E2\Phi(E_{1})=E_{2} for some ΦAut(𝒟c)\Phi\in{\rm Aut}({\mathcal{D}}^{c}).

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 d>1d>1 and 𝒟:=𝒟(𝒜){\mathcal{D}}:={\mathcal{D}}({\mathcal{A}}) is a derived category of a proper dg-category 𝒜{\mathcal{A}} such that 𝒟c:=𝒟(𝒜)c{\mathcal{D}}^{c}:={\mathcal{D}}({\mathcal{A}})^{c} has a Serre functor 𝒮{\mathcal{S}}.

Lemma 5.1 (Ping-pong lemma).

Let GG be a group acting on a set WW, and g1,g2g_{1},g_{2} elements of G. Suppose that there are non-empty, disjoint subsets W1,W2W_{1},W_{2} of WW with the property that, for each i,j(ij)i,j(i\neq j), we have gil(Wj)Wig_{i}^{l}(W_{j})\subset W_{i} for every nonzero integer ll. Then the subgroup g1,g2\langle g_{1},g_{2}\rangle generated by g1g_{1} and g2g_{2} is isomorphic to F2F_{2}

Using the ping-pong lemma, we prove the main result in this section.

Theorem 5.2.

Let E1,E2𝒟cE_{1},E_{2}\in{\mathcal{D}}^{c} be distinct dd-spherical objects. If i(E1,E2)2i(E_{1},E_{2})\geq 2, then TE1k1,TE2k2F2\langle T_{E_{1}}^{k_{1}},T_{E_{2}}^{k_{2}}\rangle\simeq F_{2} for any k1,k2\{0}k_{1},k_{2}\in{\mathbb{Z}}\backslash\{0\}.

Proof.

To apply the ping-pong lemma, we define the subsets W1,W2W_{1},W_{2} of the set of isomorphism classes of objects in 𝒟c{\mathcal{D}}^{c} as follows:

W1\displaystyle W_{1} :=\displaystyle:= {[S1]|S1𝒟c such that i(S1,E2)>i(S1,E1)}\displaystyle\{[S_{1}]~{}|~{}S_{1}\in{\mathcal{D}}^{c}\text{ such that }i(S_{1},E_{2})>i(S_{1},E_{1})\}
W2\displaystyle W_{2} :=\displaystyle:= {[S2]|S2𝒟c such that i(S2,E1)>i(S2,E2)}.\displaystyle\{[S_{2}]~{}|~{}S_{2}\in{\mathcal{D}}^{c}\text{ such that }i(S_{2},E_{1})>i(S_{2},E_{2})\}.

These are obviously disjoint, and non-empty by Lemma 4.5. By the ping-pong lemma, it suffices to check that TE1k1l(W2)W1T_{E_{1}}^{k_{1}l}(W_{2})\subset W_{1} and TE2k2l(W1)W2T_{E_{2}}^{k_{2}l}(W_{1})\subset W_{2} for each l\{0}l\in{\mathbb{Z}}\backslash\{0\}. We only show the former inclusion.

For each SW2S\in W_{2} and l\{0}l\in{\mathbb{Z}}\backslash\{0\}, the inequality (4.1) in Theorem 4.3 gives

i(TE1k1l(S),E2)\displaystyle i(T_{E_{1}}^{k_{1}l}(S),E_{2}) \displaystyle\geq i(E1,S)i(E1,E2)i(S,E2)\displaystyle i(E_{1},S)i(E_{1},E_{2})-i(S,E_{2})
\displaystyle\geq 2i(E1,S)i(S,E2)\displaystyle 2i(E_{1},S)-i(S,E_{2})
>\displaystyle> 2i(E1,S)i(S,E1)\displaystyle 2i(E_{1},S)-i(S,E_{1})
=\displaystyle= i(E1,S)=i(E1,TE1k1l(S))=i(TE1k1l(S),E1).\displaystyle i(E_{1},S)=i(E_{1},T_{E_{1}}^{k_{1}l}(S))=i(T_{E_{1}}^{k_{1}l}(S),E_{1}).

Thus TE1k1l(S)W1T_{E_{1}}^{k_{1}l}(S)\in W_{1} as desired. ∎

As a corollary, we reveal the relationship between the intersection number and presentations of subgroups generated by two spherical twists.

Corollary 5.3.

Let E1,E2𝒟cE_{1},E_{2}\in{\mathcal{D}}^{c} be distinct dd-spherical objects. Then the following are equivalent:

  1. (i)(\rm{i})

    TE1,TE2B3\langle T_{E_{1}},T_{E_{2}}\rangle\simeq B_{3}, where B3B_{3} is the braid group on 3 strands.

  2. (ii)(\rm{ii})

    TE1TE2(E1)=E2[l]T_{E_{1}}T_{E_{2}}(E_{1})=E_{2}[l] for some ll\in{\mathbb{Z}}

  3. (iii)(\rm{iii})

    i(E1,E2)=1i(E_{1},E_{2})=1.

Proof.

The assertions (iii)(ii)(i){\rm(iii)}\Rightarrow{\rm(ii)}\Rightarrow{\rm(i)} are shown by Seidel–Thomas ([ST01, Proposition 2.13 and Theorem 2.17]) and Nordskova–Volkov ([NV19, Theorem 1]) for more general dd-spherical objects, see also [Huy06, Proposition 8.22]. The braid relation gives (TE1TE2)TE1(TE1TE2)1=TE2.(T_{E_{1}}T_{E_{2}})T_{E_{1}}(T_{E_{1}}T_{E_{2}})^{-1}=T_{E_{2}}. Theorem 4.6 then implies TE1TE2(E1)=E2[l]T_{E_{1}}T_{E_{2}}(E_{1})=E_{2}[l] for some ll\in{\mathbb{Z}}.

The inequality

|i(TE1E2,E2)i(E1,E2)2|i(E2,E2)=2.\left|i(T_{E_{1}}E_{2},E_{2})-i(E_{1},E_{2})^{2}\right|\leq i(E_{2},E_{2})=2.

follows from the inequality (4.1) and easy computations. Applying TE1TE2(E1)=E2[l]T_{E_{1}}T_{E_{2}}(E_{1})=E_{2}[l], we have

i(TE1E2,E2)=i(E2,TE11E2)=i(E2,TE21TE11E2)=i(E2,E1),i(T_{E_{1}}E_{2},E_{2})=i(E_{2},T_{E_{1}}^{-1}E_{2})=i(E_{2},T_{E_{2}}^{-1}T_{E_{1}}^{-1}E_{2})=i(E_{2},E_{1}),

hence this inequality holds only in the case of i(E1,E2)=0,1i(E_{1},E_{2})=0,1 or 22. When i(E1,E2)=0i(E_{1},E_{2})=0, we have TE1=TE2T_{E_{1}}=T_{E_{2}} by the braid relation and the commutative relation: TE1TE2=TE2TE1T_{E_{1}}\circ T_{E_{2}}=T_{E_{2}}\circ T_{E_{1}}, which contradicts the distinctness. Assume that i(E1,E2)=2i(E_{1},E_{2})=2. Then the subgroup TE1,TE2\langle T_{E_{1}},T_{E_{2}}\rangle is isomorphic to the rank 2 free group by Theorem 5.2, which contradicts the braid relation. We therefore have i(E1,E2)=1i(E_{1},E_{2})=1. ∎

Corollary 5.4.

Let E1,E2𝒟cE_{1},E_{2}\in{\mathcal{D}}^{c} be distinct spherical objects. Then the following are equivalent:

  1. (i)(\rm{i})

    TE1,TE2×\langle T_{E_{1}},T_{E_{2}}\rangle\simeq{\mathbb{Z}}\times{\mathbb{Z}}

  2. (ii)(\rm{ii})

    TE1(E2)=E2[l]T_{E_{1}}(E_{2})=E_{2}[l] for some ll\in{\mathbb{Z}}

  3. (iii)(\rm{iii})

    i(E1,E2)=0i(E_{1},E_{2})=0.

Proof.

The assertions (iii)(ii)(i){\rm(iii)}\Rightarrow{\rm(ii)}\Rightarrow{\rm(i)} hold since Lemma 2.4 (ii) implies the commutative relation. Clearly, (i) implies (ii) by Theorem 4.6.

The inequality

|i(TE1E2,E2)i(E1,E2)2|i(E2,E2)=2.\left|i(T_{E_{1}}E_{2},E_{2})-i(E_{1},E_{2})^{2}\right|\leq i(E_{2},E_{2})=2.

follows from the inequality (4.1) and easy computations. Applying TE1(E2)=E2[l]T_{E_{1}}(E_{2})=E_{2}[l], we have

i(TE1E2,E2)=i(E2[l],E2)=2,i(T_{E_{1}}E_{2},E_{2})=i(E_{2}[l],E_{2})=2,

hence this inequality holds only in the case of i(E1,E2)=0,1i(E_{1},E_{2})=0,1 or 22.

When i(E1,E2)=1i(E_{1},E_{2})=1, we have TE1=TE2T_{E_{1}}=T_{E_{2}} by the braid relation (Corollary 5.3) and the commutative relation, which contradicts the distinctness. Assume that i(E1,E2)=2i(E_{1},E_{2})=2. Then the subgroup TE1,TE2\langle T_{E_{1}},T_{E_{2}}\rangle is isomorphic to the rank 2 free group by Theorem 5.2, which contradicts the commutative relation. We therefore have i(E1,E2)=0i(E_{1},E_{2})=0. ∎

Corollary 5.5.

Let E1,E2𝒟cE_{1},E_{2}\in{\mathcal{D}}^{c} be distinct spherical objects. Then the following are equivalent:

  1. (i)(\rm{i})

    TE1,TE2F2\langle T_{E_{1}},T_{E_{2}}\rangle\simeq F_{2}

  2. (ii)(\rm{ii})

    i(E1,E2)2i(E_{1},E_{2})\geq 2.

Proof.

The assertion (ii)(i){\rm(ii)}\Rightarrow{\rm(i)} follows from Theorem 5.2. The converse is given by Corollary 5.3 and Corollary 5.4. ∎

6. Proof of Theorem 4.3

In this section, we prove Theorem 4.3.

Throughout this section, we consider the case of d0d\neq 0, and let 𝒜{\mathcal{A}} be a proper dg-category such that 𝒟(𝒜)c{\mathcal{D}}({\mathcal{A}})^{c} has a Serre functor 𝒮{\mathcal{S}}.

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 LL she has Hom(L,L)=K[ϵ]/ϵ2,deg(ϵ)=d{\rm Hom}(L,L)=K[\epsilon]/\epsilon^{2},\ \mathrm{deg}(\epsilon)=d, d(ϵ)=0d(\epsilon)=0, 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 Ad:=K[ϵ]/ϵ2A_{d}:=K[\epsilon]/\epsilon^{2} for the dg-algebra of graded dual numbers with deg(ϵ)=d\deg(\epsilon)=d and zero differential.

The rationale behind Proposition 6.1 is that, given a dd-spherical object EE, as d0d\neq 0 there exists a unique structure of graded algebra on

nHom𝒟(𝒜)(E,E[n])[n].\bigoplus_{n\in\mathbb{Z}}\mathrm{Hom}_{{\mathcal{D}}({\mathcal{A}})}(E,E[n])[-n].

Namely, we have

nHom𝒟(𝒜)(E,E[n])[n]Ad.\bigoplus_{n\in\mathbb{Z}}\mathrm{Hom}_{{\mathcal{D}}({\mathcal{A}})}(E,E[n])[-n]\simeq A_{d}.

This implies, as AdA_{d} is intrinsically formal, see e.g. [KS22, Proposition 2.2], that the dg-endomorphism-algebra of an h-projective resolution of EE is quasi-isomorphic to AdA_{d}. What we prove in Proposition 6.1 is that we can choose the h-projective resolution of EE so that it carries and AdA_{d}-dg-module structure.

Proposition 6.1.

Let E𝒟(𝒜)cE\in{\mathcal{D}}({\mathcal{A}})^{c} be a d-spherical object. Then, there exists F𝒫(𝒜)cF\in{\mathcal{P}}({\mathcal{A}})^{c} with the following properties:

  1. (i)(\rm{i})

    FEF\simeq E in 𝒟(𝒜){\mathcal{D}}({\mathcal{A}})

  2. (ii)(\rm{ii})

    there exists αϵHom𝒫(𝒜)d(F,F)\alpha_{\epsilon}\in{\rm Hom}^{d}_{{\mathcal{P}}({\mathcal{A}})}(F,F) such that αϵ2=0\alpha_{\epsilon}^{2}=0

  3. (iii)(\rm{iii})

    under the isomorphism of (i), αϵ\alpha_{\epsilon} corresponds to the canonical extension E[d]EE[-d]\to E.

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 ε:E[d]E\varepsilon\colon E[-d]\rightarrow E for the canonical extension of EE with itself, and we write E𝒫(𝒜)E^{\prime}\in{\mathcal{P}}({\mathcal{A}}) for a fixed h-projective resolution of EE together with a fixed quasi-isomorphism

EE.E^{\prime}\rightarrow E. (6.1)

We begin by proving the following

Lemma 6.2.

With the notation as above, the sequence of 𝒜{\mathcal{A}}-dg-modules and morphisms

𝜀E[2d](1)d+1εE[d]𝜀E\dots\xrightarrow{\varepsilon}E[-2d]\xrightarrow{(-1)^{d+1}\varepsilon}E[-d]\xrightarrow{\varepsilon}E

can be lifted to a one-sided infinite twisted complex in 𝒫(𝒜){\mathcal{P}}({\mathcal{A}}).

Proof.

The components of the sought twisted complexes will be given by E[id]E^{\prime}[id] for i0i\leq 0. Then, to prove the statement of the lemma it is enough to find morphisms

αij:E[id]E[jd]\alpha_{ij}\colon E^{\prime}[id]\rightarrow E^{\prime}[jd]

of degree ij+1i-j+1, i<ji<j, such that the following diagram commutes for every i1i\leq-1

E[id]{E^{\prime}[id]}E[(i+1)d]{E^{\prime}[(i+1)d]}E[id]{E[id]}E[(i+1)d]{E[(i+1)d]}αi(i+1)\scriptstyle{\alpha_{i(i+1)}}(6.1)\scriptstyle{\eqref{eqn:quasi-iso-E}}(6.1)\scriptstyle{\eqref{eqn:quasi-iso-E}}(1)(i+1)(d+1)ε\scriptstyle{(-1)^{(i+1)(d+1)}\varepsilon} (6.2)

and such that

(1)jdαij+i<k<jαkjαik=0.(-1)^{j}d\alpha_{ij}+\sum_{i<k<j}\alpha_{kj}\circ\alpha_{ik}=0. (6.3)

Indeed, then the sought twisted complex will be given by (E[id],αij)(E^{\prime}[id],\alpha_{ij}), i0i\leq 0.

In the following, when we say that αi(i+1)\alpha_{i(i+1)} induces the morphism (1)(i+1)(d+1)ε(-1)^{(i+1)(d+1)}\varepsilon via the quasi-isomorphism (6.1), we mean that the diagram (6.2) commutes.

We will prove the existence of the morphisms αij\alpha_{ij} by induction on jij-i. First of all, notice that if we have all the maps αij\alpha_{ij} for ji<nj-i<n, then to define the maps αij\alpha_{ij} with ji=nj-i=n it is enough to define αn0\alpha_{-n0}. Indeed, if we have αn0\alpha_{-n0}, then we can set

αij:=(1)j(d+1)αn0[jd]\alpha_{ij}:=(-1)^{j(d+1)}\alpha_{-n0}[jd] (6.4)

for ji=nj-i=n, and using (6.3) for αn0\alpha_{-n0} we have555In the above equations, we make use of the fact that for any morphism ff and any mm\in\mathbb{Z} it holds that d(f[m])=(1)m(df)[m]d(f[m])=(-1)^{m}(df)[m].

(1)jd(αij)\displaystyle(-1)^{j}d(\alpha_{ij}) =d(αn0)[jd]\displaystyle\,=d(\alpha_{-n0})[jd]
=(n<k<0αk0αnk)[jd]\displaystyle\,=-\left(\sum_{-n<k<0}\alpha_{k0}\circ\alpha_{-nk}\right)[jd]
=(i<k+j<j((1)j(d+1)α(k+j)j)((1)j(d+1)αi(k+j)))\displaystyle\,=-\left(\sum_{i<k+j<j}((-1)^{j(d+1)}\alpha_{(k+j)j})\circ((-1)^{j(d+1)}\alpha_{i(k+j)})\right)
=i<r<jαrjαir.\displaystyle\,=-\sum_{i<r<j}\alpha_{rj}\circ\alpha_{ir}.

We now begin the inductive construction. For the case ji=1j-i=1 we use that by the definition of an h-projective resolution we have

H0(Hom𝒜(E,E[d]))Hom𝒟(𝒜)(E,E[d]),H^{0}({\rm Hom}_{{\mathcal{A}}}(E^{\prime},E^{\prime}[d]))\simeq{\rm Hom}_{{\mathcal{D}}({\mathcal{A}})}(E,E[d]),

and therefore we can define α10\alpha_{-10} by lifting ε\varepsilon along the previous isomorphism. With this choice and the definition (6.4), we get that αi(i+1)\alpha_{i(i+1)} induces (1)(i+1)(d+1)ε(-1)^{(i+1)(d+1)}\varepsilon via the quasi-isomorphism (6.1), as we wanted.

Now assume that we defined αij\alpha_{ij} for any ji<nj-i<n. We claim that to construct αn0\alpha_{-n0} it is enough to prove that n<k<0αk0αnk\sum_{-n<k<0}\alpha_{k0}\circ\alpha_{-nk} is a closed morphism. Indeed, notice that n<k<0αk0αnk\sum_{-n<k<0}\alpha_{k0}\circ\alpha_{-nk} has degree 2n2-n as a morphism E[nd]EE^{\prime}[-nd]\rightarrow E^{\prime}, and therefore it corresponds to a morphism EEE^{\prime}\rightarrow E^{\prime} of degree nd+2nnd+2-n. The cohomology of Hom𝒜(E,E){\rm Hom}_{{\mathcal{A}}}(E^{\prime},E^{\prime}) is concentrated in degree 0 and dd because EE is dd-spherical, and therefore a closed element can be non-trivial in cohomology if and only if it has degree 0 or dd. Now, as we are in the case n2n\geq 2, we have either

nd+2n2(d1)+2=2dnd+2-n\geq 2(d-1)+2=2d

or nd+2n2dnd+2-n\leq 2d, depending on whether d>0d>0 or d<0d<0. In either case, the degree of n<k<0αk0αnk\sum_{-n<k<0}\alpha_{k0}\circ\alpha_{-nk} is not 0 or dd, and therefore if n<k<0αk0αnk\sum_{-n<k<0}\alpha_{k0}\circ\alpha_{-nk} is closed it must be the differential of some morphism E[nd]EE^{\prime}[-nd]\rightarrow E^{\prime} of degree 1n1-n that we can take to be our αn0\alpha_{-n0}.

We now prove that n<k<0αk0αnk\sum_{-n<k<0}\alpha_{k0}\circ\alpha_{-nk} is closed. We have

d(n<k<0αk0αnk)\displaystyle\,d\left(\sum_{-n<k<0}\alpha_{k0}\circ\alpha_{-nk}\right)
=\displaystyle= n<k<0d(αk0)αnk+(1)k+1αk0d(αnk)\displaystyle\,\sum_{-n<k<0}d(\alpha_{k0})\circ\alpha_{-nk}+(-1)^{k+1}\alpha_{k0}\circ d(\alpha_{-nk})
=\displaystyle= n<k<0k<r<0α0rαkrαnk+n<k<0(1)2(k+1)n<s<kαk0αskαns\displaystyle\,-\sum_{-n<k<0}\sum_{k<r<0}\alpha_{0r}\circ\alpha_{kr}\circ\alpha_{-nk}+\sum_{-n<k<0}(-1)^{2(k+1)}\sum_{-n<s<k}\alpha_{k0}\circ\alpha_{sk}\circ\alpha_{-ns}

where in passing from the second to the third line we used that αk0\alpha_{k0} and αnk\alpha_{-nk} satisfy (6.3) by the induction hypothesis. Our aim is to prove that the above terms sum to zero. Take a decomposition of n0-n\to 0 in three steps: n<s<r<0-n<s<r<0. To get the term αr0αsrαns\alpha_{r0}\circ\alpha_{sr}\circ\alpha_{-ns} we have two possibilities: either from d(αs0)αnsd(\alpha_{s0})\circ\alpha_{-ns} or (1)s+1αr0d(αns)(-1)^{s+1}\alpha_{r0}\circ d(\alpha_{-ns}). In the first case we get the term αr0αsrαns-\alpha_{r0}\circ\alpha_{sr}\circ\alpha_{-ns}, in the second case (1)2(s+1)αr0αsrαns(-1)^{2(s+1)}\alpha_{r0}\circ\alpha_{sr}\circ\alpha_{-ns}, and they cancel out.

Hence, given αij\alpha_{ij} for ji<nj-i<n, we can define αn0\alpha_{-n0}, and the inductive step is complete. Thus, the proof of the lemma is complete. ∎

In the following, we write

(E[id],αij)(E^{\prime}[id],\alpha_{ij}) (6.5)

i0i\leq 0, for the twisted complex constructed in Lemma 6.2.

The next step in the proof of Proposition 6.1 is the following

Lemma 6.3.

Let us write E′′E^{\prime\prime} for the convolution of the twisted complex (6.5). Then, there exists a closed morphism p:E′′E′′p\colon E^{\prime\prime}\rightarrow E^{\prime\prime} of degree 1d1-d such that the dg-cone of p[1]p[-1] is 𝒜{\mathcal{A}}-h-projective.

Proof.

We will define a morphism π\pi at the level of the twisted complex (6.5), and then convolve it to a morphism p:E′′E′′p\colon E^{\prime\prime}\rightarrow E^{\prime\prime}.

Following the convention introduced in § 2.3, we define the morphism of twisted complexes

π={πdj(1j):=id:E[jd]E[(1j)d]}\pi=\left\{\pi_{-d}^{-j(1-j)}:={\rm id}\colon E^{\prime}[-jd]\rightarrow E^{\prime}[(1-j)d]\right\}

where j1j\geq 1. By definition, π\pi has degree 1d1-d. We now prove that is a closed morphism. As per the definition given in § 2.3, the differential of πdk(1k)\pi_{-d}^{-k(1-k)} is given by

1k<mα(1k)mπdk(1k)(1)1dm<kπdk(1k)αm(k).\sum_{1-k<m}\alpha_{(1-k)m}\circ\pi_{-d}^{-k(1-k)}-(-1)^{1-d}\sum_{m<-k}\pi_{-d}^{-k(1-k)}\circ\alpha_{m(-k)}.

To prove that π\pi is closed, we fix j0j\geq 0 and we focus our attention to the components of π\pi that map to E[jd]E^{\prime}[-jd]. These are given by

1+j<kα(1k)jπdk(1k)(1)1dm<1jπd(1j)jαm(1j).\sum_{1+j<k}\alpha_{(1-k)-j}\circ\pi_{-d}^{-k(1-k)}-(-1)^{1-d}\sum_{m<-1-j}\pi_{-d}^{(-1-j)-j}\circ\alpha_{m(-1-j)}. (6.6)

Using the definition of αij\alpha_{ij} given in (6.4), we see that (6.6) is equal to the shift by jd-jd of

(1k+j<0(1)j(d+1)α(1k+j)0πd(jk)(1k+j)(1)(j+1)(d+1)m+1+j<0πd10α(m+j)1)\displaystyle\,\left(\sum_{1-k+j<0}(-1)^{j(d+1)}\alpha_{(1-k+j)0}\circ\pi_{-d}^{(j-k)(1-k+j)}-(-1)^{(j+1)(d+1)}\sum_{m+1+j<0}\pi_{-d}^{-10}\circ\alpha_{(m+j)-1}\right)
=\displaystyle= (1)j(d+1)(n<0αn0πd(n1)n(1)(d+1)n<1πd10αn(1))=0.\displaystyle\,(-1)^{j(d+1)}\left(\sum_{n<0}\alpha_{n0}\circ\pi_{-d}^{(n-1)n}-(-1)^{(d+1)}\sum_{n<-1}\pi_{-d}^{-10}\circ\alpha_{n(-1)}\right)=0.

Here the last equality follows from the fact that, given any n<0n<0, the first term contributes with αn0πd(n1)n\alpha_{n0}\circ\pi_{-d}^{(n-1)n}, while the second term contributes with (1)d+1πd10α(n1)1(-1)^{d+1}\pi_{-d}^{-10}\circ\alpha_{(n-1)-1}. Using (6.4), one sees that these two terms cancel out.

Hence, π:(6.5)(6.5)\pi\colon\eqref{eqn:twisted-complex-lift}\rightarrow\eqref{eqn:twisted-complex-lift} is a closed morphism of twisted complexes of degree 1d1-d, and we define p:E′′E′′[1d]p\colon E^{\prime\prime}\rightarrow E^{\prime\prime}[1-d] as its convolution.

Let us write F:=C(p[1])F:=C(p[-1]) for the dg-cone of p[1]:E′′[1]E′′[d]p[-1]\colon E^{\prime\prime}[-1]\rightarrow E^{\prime\prime}[-d] as defined in § 2.2.

We now show that FF is 𝒜{\mathcal{A}}-h-projective. It is enough to prove that E′′E^{\prime\prime} is 𝒜{\mathcal{A}}-h-projective, and this is what we show. The fact that E′′E^{\prime\prime} is h-projective follows from the fact that it is the convolution of a one-sided twisted complex whose components are 𝒜{\mathcal{A}}-h-projective, and they are non-zero only in negative degree. Indeed, these properties imply that E′′E^{\prime\prime} has an exhaustive filtration by h-projective 𝒜{\mathcal{A}}-dg-modules, see also [Sta18, Tag 09KK], and thus it is 𝒜{\mathcal{A}}-h-projective. ∎

In the lemma below, we show that FF in Lemma 6.3 is an 𝒜{\mathcal{A}}-h-projective resolution of EE.

Lemma 6.4.

The dg-cone FF of the morphism p[1]:E′′[1]E′′[d]p[-1]\colon E^{\prime\prime}[-1]\rightarrow E^{\prime\prime}[-d] constructed in Lemma 6.3 is quasi-isomorphic to EE.

Proof.

To prove the lemma, we construct a closed, degree zero morphism f:E′′[d]Ef\colon E^{\prime\prime}[-d]\rightarrow E^{\prime} such that we have a distinguished triangle

E′′[d]𝑓EE′′𝑝E′′[1d]E^{\prime\prime}[-d]\xrightarrow{f}E^{\prime}\rightarrow E^{\prime\prime}\xrightarrow{p}E^{\prime\prime}[1-d]

thus proving that F=C(p[1])EEF=C(p[-1])\simeq E^{\prime}\simeq E in 𝒟(𝒜){\mathcal{D}}({\mathcal{A}}).

We define ff as the convolution of a closed, degree zero morphism of twisted complexes φ:(6.5)(E[d],0)\varphi\colon\eqref{eqn:twisted-complex-lift}\rightarrow(E^{\prime}[d],0), where (E[d],0)(E^{\prime}[d],0) is the twisted complex with E[d]E^{\prime}[d] in position zero. The degree zero morphism φ\varphi is defined as

φ={φnn0:=α(n1)0[d]:E[nd]E[d]}n0.\begin{array}[]{lcr}\varphi=\{\varphi_{n}^{n0}:=\alpha_{(n-1)0}[d]\colon E^{\prime}[-nd]\rightarrow E^{\prime}[d]\}&&n\geq 0.\end{array}

Let us show that φ\varphi is closed. We fix j0j\leq 0, and we focus our attention on the component of dφd\varphi starting from E[jd]E^{\prime}[jd]. By definition, this is given by

d(φj0)j<k0φ0k0αjk\displaystyle d(\varphi^{j0})-\sum_{j<k\leq 0}\varphi_{0}^{k0}\circ\alpha_{jk}
=\displaystyle= (1)dd(α(j1)0)[d]j<k0α(k1)0[d]αjk\displaystyle(-1)^{d}d(\alpha_{(j-1)0})[d]-\sum_{j<k\leq 0}\alpha_{(k-1)0}[d]\circ\alpha_{jk}
=\displaystyle= (1)dd(α(j1)0)[d](1)d+1j1<k1<0(α(k1)0α(j1)(k1))[d]\displaystyle(-1)^{d}d(\alpha_{(j-1)0})[d]-(-1)^{d+1}\sum_{j-1<k-1<0}(\alpha_{(k-1)0}\circ\alpha_{(j-1)(k-1)})[d]
=\displaystyle= 0\displaystyle 0

where the last equality follows from the fact that the αij\alpha_{ij} satisfy (6.3). Hence, φ\varphi is a closed morphism of degree zero, and upon convolution (and a shift by d-d) it induces a morphism f:E′′[d]Ef\colon E^{\prime\prime}[-d]\rightarrow E^{\prime}.

To conclude the proof of the lemma, we construct morphisms ii, jj, and qq such that

E′′[1d]𝑖E′′𝑝E′′[1d]andE𝑗E′′𝑞E\begin{array}[]{lcr}E^{\prime\prime}[1-d]\xrightarrow{i}E^{\prime\prime}\xrightarrow{p}E^{\prime\prime}[1-d]&\mathrm{and}&E^{\prime}\xrightarrow{j}E^{\prime\prime}\xrightarrow{q}E^{\prime}\end{array} (6.7)

realise E′′E^{\prime\prime} as the dg-cone of ff.

The morphisms jj and qq are defined as the inclusion of EE^{\prime} into E′′E^{\prime\prime} and the projection E′′E^{\prime\prime} onto EE^{\prime}, respectively. Such maps exists because at the level of underlying graded modules E′′E^{\prime\prime} is the direct sum i0E[id]\oplus_{i\leq 0}E^{\prime}[id].

The morphism ii is defined as the convolution of the morphism ι:(6.5)(6.5)\iota\colon\eqref{eqn:twisted-complex-lift}\rightarrow\eqref{eqn:twisted-complex-lift} defined by

ι={ιd(1j)j:=id:E[(1j)d]E[jd]}\iota=\left\{\iota_{d}^{(1-j)-j}:=\mathrm{id}\colon E^{\prime}[(1-j)d]\rightarrow E^{\prime}[-jd]\right\}

for j1j\geq 1.

The fact that the morphisms in (6.7) realise E′′E^{\prime\prime} as the dg-cone of ff 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

E′′[d]𝑓E𝑗E′′𝑝E′′[1d]E^{\prime\prime}[-d]\xrightarrow{f}E^{\prime}\xrightarrow{j}E^{\prime\prime}\xrightarrow{p}E^{\prime\prime}[1-d]

and therefore we have the quasi-isomorphisms

F=C(p[1])𝑗E(6.1)EF=C(p[-1])\xleftarrow{j}E^{\prime}\xrightarrow{\eqref{eqn:quasi-iso-E}}E

where we wrote j:EFj\colon E^{\prime}\rightarrow F for the morphism E𝑗E′′C(p[1])E^{\prime}\xrightarrow{j}E^{\prime\prime}\rightarrow C(p[-1]), 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 𝒜{\mathcal{A}}-h-projective resolution FF of EE. This FF 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 αϵ\alpha_{\epsilon}. We define αϵ:FF\alpha_{\epsilon}\colon F\rightarrow F using the direct sum decomposition of FF. Namely, as a graded module FF is given by E′′E′′[d]E^{\prime\prime}\oplus E^{\prime\prime}[-d], and when we write (a,b)F(a,b)\in F we mean that aE′′a\in E^{\prime\prime} and bE′′[d]b\in E^{\prime\prime}[-d] in this decomposition. Then, αϵ\alpha_{\epsilon} is given by

F=C(p[1])(a,b)αϵ(a,b):=(0,a)C(p[1])=F.F=C(p[-1])\ni(a,b)\mapsto\alpha_{\epsilon}(a,b):=(0,a)\in C(p[-1])=F.

It is clear that αϵ\alpha_{\epsilon} is a closed, degree dd morphism such that αϵ2=0\alpha_{\epsilon}^{2}=0.

We now prove that under the quasi-isomorphisms F𝑗E(6.1)EF\xleftarrow{j}E^{\prime}\xrightarrow{\eqref{eqn:quasi-iso-E}}E the morphism αϵ\alpha_{\epsilon} corresponds to the canonical extension ε\varepsilon of EE. By the definition of α10\alpha_{-10}, this is equivalent to prove that under the quasi-isomorphism E𝑗FE^{\prime}\xrightarrow{j}F the morphism αϵ\alpha_{\epsilon} corresponds to α10\alpha_{-10}. We prove the latter statement.

Take eEe\in E^{\prime} a closed element. Then (recall the notation for elements belonging to the convolution of a one-sided twisted complex that we introduced in § 2.4)

αϵ(j(e))=(0,(,0,0,e))=((,0,0,α10(e)),0)+dF((,0,e,0),0).\alpha_{\epsilon}(j(e))=(0,(\dots,0,0,e))=((\dots,0,0,\alpha_{-10}(e)),0)+d_{F}((\dots,0,e,0),0).

Therefore, we get

αϵ(j(e))=j(α10(e))+dF((,0,e,0),0),\alpha_{\epsilon}(j(e))=j(\alpha_{-10}(e))+d_{F}((\dots,0,e,0),0),

which means that αϵj=jα10\alpha_{\epsilon}\circ j=j\circ\alpha_{-10} in 𝒟(𝒜){\mathcal{D}}({\mathcal{A}}), as we wanted. Thus, the proof of Proposition 6.1 is complete. ∎

From now on, every time we have a spherical object EE we implicitly assume we replaced it with FF as in Proposition 6.1. Hence, every spherical object EE is really E𝒫(𝒜)cE\in{\mathcal{P}}({\mathcal{A}})^{c} and there is αϵHom𝒜d(E,E)\alpha_{\epsilon}\in{\rm Hom}^{d}_{{\mathcal{A}}}(E,E) such that αϵ2=0\alpha_{\epsilon}^{2}=0.

Let E𝒟(𝒜)cE\in{\mathcal{D}}({\mathcal{A}})^{c} be a dd-spherical object. The following proposition is an analogue of [Kea14, Proposition 6.3] (recall the convention § 2.3 (ii)).

Proposition 6.5.

For any N𝒟(𝒜)N\in{\mathcal{D}}({\mathcal{A}}) the object TEnN,n0T^{n}_{E}N,\ n\geq 0, is the convolution of the twisted complex

RHom𝒜(E,N)E[nd]dnd2RHom𝒜(E,N)EevN,{\rm RHom}_{\mathcal{A}}(E,N)\otimes E[-nd]\xrightarrow{d_{n}}\cdots\xrightarrow{d_{2}}{\rm RHom}_{\mathcal{A}}(E,N)\otimes E\xrightarrow{ev}N, (6.8)

where dk(fe):=fαϵe+(1)kfαϵed_{k}(f\otimes e):=f\alpha_{\epsilon}\otimes e+(-1)^{k}f\otimes\alpha_{\epsilon}e.

Proof.

If we had RHom𝒜(E,E)=Hom𝒜(E,E)Ad{\rm RHom}_{\mathcal{A}}(E,E)={\rm Hom}_{\mathcal{A}}(E,E)\simeq A_{d}, 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 K[αϵ]Hom𝒜(E,E)K[\alpha_{\epsilon}]\subset{\rm Hom}_{\mathcal{A}}(E,E) is a quasi-isomorphism. Therefore, in the twisted complex representing TEnNT^{n}_{E}N we can replace Hom𝒜(E,E){\rm Hom}_{\mathcal{A}}(E,E) with K[αϵ]K[\alpha_{\epsilon}] 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 Cn(E,N)C_{n}(E,N) for the convolution of the twisted complex

RHom𝒜(E,N)E[nd]dnd2RHom𝒜(E,N)E.{\rm RHom}_{\mathcal{A}}(E,N)\otimes E[-nd]\xrightarrow{d_{n}}\cdots\xrightarrow{d_{2}}{\rm RHom}_{\mathcal{A}}(E,N)\otimes E. (6.9)

Then, we have

Proposition 6.6.

Let M,N𝒟(𝒜)M,N\in{\mathcal{D}}({\mathcal{A}}) and n0n\geq 0. Then, RHom𝒜(M,Cn(E,N)){\rm RHom}_{\mathcal{A}}(M,C_{n}(E,N)) is isomorphic to the convolution of the twisted complex

RHom𝒜(E,N)RHom𝒜(M,E)[nd]dnd2RHom𝒜(E,N)RHom𝒜(M,E),{\rm RHom}_{\mathcal{A}}(E,N)\otimes{\rm RHom}_{\mathcal{A}}(M,E)[-nd]\xrightarrow{d_{n}}\cdots\xrightarrow{d_{2}}{\rm RHom}_{\mathcal{A}}(E,N)\otimes{\rm RHom}_{\mathcal{A}}(M,E),

where dk(fg):=fαϵg+(1)kfαϵgd_{k}(f\otimes g):=f\alpha_{\epsilon}\otimes g+(-1)^{k}f\otimes\alpha_{\epsilon}g.

Proof.

The claim follows from the definition of Cn(E,N)C_{n}(E,N) and [AL17, Lemma 3.4]. ∎

6.2. Graded dual numbers

Recall that we write Ad=K[ϵ]/ϵ2A_{d}=K[\epsilon]/\epsilon^{2} for the dg-algebra of the graded dual numbers with deg(ϵ)=d\mathrm{deg}(\epsilon)=d. The aim of this subsection is to classify finite dimensional dg-modules over AdA_{d}. We write 𝒟b(Ad){\mathcal{D}}^{b}(A_{d}) for the triangulated subcategory of 𝒟(Ad){\mathcal{D}}(A_{d}) given by AdA_{d}-dg-modules with finite dimensional total cohomology.

To prove Proposition 6.7 below, we will use Koszul duality. Namely, let us define K[q]K[q] as the dg-algebra with deg(q)=1d\deg(q)=1-d and d(q)=0d(q)=0. Then, Koszul duality says that we have an equivalence Φ:𝒟(K[q])c𝒟b(Ad)\Phi\colon{\mathcal{D}}(K[q])^{c}\xrightarrow{\simeq}{\mathcal{D}}^{b}(A_{d}) sending K[q]K[q] to KK. For a recent proof of this statement, see [KS22].

Having Φ\Phi means that to prove a structure theorem for modules in 𝒟b(Ad){\mathcal{D}}^{b}(A_{d}) it is enough to study the structure of modules in 𝒟(K[q])c{\mathcal{D}}(K[q])^{c}, and this is what we do.

Before moving on to the proposition, let us notice a useful consequence of the equivalence 𝒟(K[q])c𝒟b(Ad){\mathcal{D}}(K[q])^{c}\simeq{\mathcal{D}}^{b}(A_{d}). By definition, we have Hom𝒟(K[q])c(K[q][m(1d)],K[q])K{\rm Hom}_{{\mathcal{D}}(K[q])^{c}}(K[q][-m(1-d)],K[q])\simeq K for any m1m\geq 1, and the unique non-trivial extension is given by the convolution of the twisted complex

K[q][m(1d)]qmK[q],K[q][-m(1-d)]\xrightarrow{q^{m}}K[q], (6.10)

which is quasi-isomorphic K[q]/qmK[q]/q^{m}. As Φ\Phi is an equivalence sending K[q]K[q] to KK, we have

KHom𝒟(K[q])c(K[q][m(1d)],K[q])Hom𝒟b(Ad)(K[m(1d)],K),K\simeq{\rm Hom}_{{\mathcal{D}}(K[q])^{c}}(K[q][-m(1-d)],K[q])\simeq{\rm Hom}_{{\mathcal{D}}^{b}(A_{d})}(K[-m(1-d)],K),

and therefore there is a unique non-trivial extension of KK by itself of degree m(1d)m(1-d) in 𝒟b(Ad){\mathcal{D}}^{b}(A_{d}) for any m1m\geq 1. This extension is given by the convolution of the twisted complex

Ad[(1m)d]ϵAd[(2m)d]ϵϵAd[d]ϵAd,A_{d}[(1-m)d]\xrightarrow{\epsilon}A_{d}[(2-m)d]\xrightarrow{\epsilon}\dots\xrightarrow{\epsilon}A_{d}[-d]\xrightarrow{\epsilon}A_{d}, (6.11)

As Φ\Phi is an equivalence, we get that Φ((6.10))\Phi(\eqref{eqn:extension-K[q]}) is isomorphic, up to a shift that can be computed to be m(1d)1m(1-d)-1, to (6.11)\eqref{eqn:extension-A-d-finite} for any m1m\geq 1.

We write BnB_{n}, n1n\geq 1, for the convolution of (6.11), and B0=K𝒟b(Ad)B_{0}=K\in{\mathcal{D}}^{b}(A_{d}). Similarly, we write CnC_{n} for the convolution of (6.11) with AdA_{d} replaced by AdopA_{d}^{op}, and C0=K𝒟b(Adop)C_{0}=K\in{\mathcal{D}}^{b}(A_{d}^{op}). The following proposition gives the desired classification (cf. [Kea14, Proposition 5.3]).

Proposition 6.7.

Let MM be a finite dimensional right (resp. left) AdA_{d}-dg-module. Then, MM is quasi-isomorphic to a finite direct sum of copies of shifts of BnB_{n}’s (resp. CnC_{n}’s). Moreover, MM is compact if and only if B0B_{0} (resp. C0C_{0}) does not appear.

Remark 6.8.

The above statement pairs up with [KS22, Proposition 2.2] to show why 𝒟(Ad)c(𝒟b(Ad)){\mathcal{D}}(A_{d})^{c}(\subsetneq{\mathcal{D}}^{b}(A_{d})) is generated by the AdA_{d} as a triangulated category (without additional idempotent completion).

Proof.

We prove the statement for right modules; this will suffice because AdA_{d} is commutative.

Let us fix M𝒟b(Ad)M\in{\mathcal{D}}^{b}(A_{d}) and M𝒟(K[q])cM^{\prime}\in{\mathcal{D}}(K[q])^{c} such that Φ(M)M\Phi(M^{\prime})\simeq M. Notice that if we forget the grading, then K[q]K[q] is a PID. Hence, the cohomology of the dg-module MM^{\prime} splits as a finite direct sum

H(M)K[q][s1]K[q][st]K[q]/qn1[m1]K[q]/qnr[mr]H^{\bullet}(M^{\prime})\simeq K[q][s_{1}]\oplus\cdots\oplus K[q][s_{t}]\oplus K[q]/q^{n_{1}}[m_{1}]\oplus\cdots\oplus K[q]/q^{n_{r}}[m_{r}] (6.12)

for some nr0,mi,sin_{r}\geq 0,~{}m_{i},s_{i}\in{\mathbb{Z}}.

Let us write FmF_{m} for the convolution of (6.10). Then, notice that FmF_{m} is a free resolution of K[q]/qmK[q]/q^{m}. Hence, replacing K[q]/qniK[q]/q^{n_{i}} with FniF_{n_{i}} in (6.12), we can lift the isomorphism (6.12) to a quasi-isomorphism666In slightly more detail: if K[q]/qniK[q]/q^{n_{i}} appears in (6.12), then it means that there exists mMm\in M^{\prime} such that d(m)=0d(m)=0 and qnim=d(m)q^{n_{i}}m=d(m^{\prime}) for some mMm^{\prime}\in M. Then, we define FniMF_{n_{i}}\rightarrow M^{\prime} by sending (1,0),(0,1)Fni(1,0),(0,1)\in F_{n_{i}} to mm and mm^{\prime}, respectively. Here we employed the notation we introduced in § 2.4 for elements belonging to the convolution of a twisted complex.

MK[q][s1]K[q][st]Fn1[m1]Fnr[mr]M^{\prime}\simeq K[q][s_{1}]\oplus\cdots\oplus K[q][s_{t}]\oplus F_{n_{1}}[m_{1}]\oplus\cdots\oplus F_{n_{r}}[m_{r}] (6.13)

Applying Φ\Phi to (6.13) and using the isomorphisms Φ((6.10))(6.11)[m(1d)+1]\Phi(\eqref{eqn:extension-K[q]})\simeq\eqref{eqn:extension-A-d-finite}[-m(1-d)+1] and Φ(K[q])B0\Phi(K[q])\simeq B_{0}, we obtain the sought decomposition

MΦ(M)i=1tB0[si]j=1rBnj[mjnj(1d)+1].M\simeq\Phi(M^{\prime})\simeq\bigoplus^{t}_{i=1}B_{0}[s_{i}]\oplus\bigoplus^{r}_{j=1}B_{n_{j}}[m_{j}-n_{j}(1-d)+1].

Finally, to prove the claim about the compactness of MM recall that 𝒟(Ad)c{\mathcal{D}}(A_{d})^{c} is closed under taking direct summands and notice that of the BnB_{n}’s the only non-compact one is B0B_{0} (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 M,NM,N two dg-modules over the graded dual numbers AdA_{d} and n0n\geq 0, we define (MAdN)n(M\otimes_{A_{d}}N)_{n} as the convolution of the twisted complex

MN[nd]dnd3MN[d]d2MN,M\otimes N[-nd]\xrightarrow{d_{n}}\cdots\xrightarrow{d_{3}}M\otimes N[-d]\xrightarrow{d_{2}}M\otimes N,

where dk(mn):=mϵn+(1)kmϵnd_{k}(m\otimes n):=m\epsilon\otimes n+(-1)^{k}m\otimes\epsilon n.

Remark 6.10.

The construction (MAdN)n(M\otimes_{A_{d}}N)_{n} is independent of the quasi-isomorphism classes of MM and NN. Indeed, this follows easily from the definition of (MAdN)n(M\otimes_{A_{d}}N)_{n} and e.g. [AL21, Corollary 2.12].

We state Theorem 4.3 again:

Theorem 6.11.

Let E𝒟(𝒜)cE\in{\mathcal{D}}({\mathcal{A}})^{c} be a spherical object and M,N𝒟(𝒜)M,N\in{\mathcal{D}}({\mathcal{A}}) objects such that i(E,M)i(E,M) and i(E,N)<i(E,N)<\infty. For any k\{0}k\in{\mathbb{Z}}\backslash\{0\}, we have

i(E,M)i(E,N)i(TEkM,N)+i(M,N).i(E,M)i(E,N)\leq i(T^{k}_{E}M,N)+i(M,N). (6.14)
Proof.

By replacing MM by TEkMT_{E}^{-k}M, it suffices to prove the claim for k<0k<0. The case k=1k=-1 is proved by direct computations and using i(Hom(E,TE1M)E,N)=i(E,M)i(E,N)i({\rm Hom}^{\bullet}(E,T_{E}^{-1}M)\otimes E,N)=i(E,M)i(E,N), so we can assume k>1-k>1.

By the definition of Ck(E,N)C_{-k}(E,N), see (6.9), we have the following distinguished triangle

Ck(E,N)evNTEkN.C_{-k}(E,N)\xrightarrow{ev}N\rightarrow T_{E}^{-k}N.

Applying RHom𝒜(M,){\rm RHom}_{\mathcal{A}}(M,-) to this distinguished triangle, as i(M,TEkN)=i(TEkM,N)i(M,T^{-k}_{E}N)=i(T^{k}_{E}M,N) and intersection numbers are subadditive on distinguished triangles, we see that it is enough to prove

dimKRHom𝒜(M,Ck(E,N))i(M,E)i(E,N)(=i(E,M)i(E,N)).\dim_{K}{\rm RHom}_{\mathcal{A}}(M,C_{-k}(E,N))\geq i(M,E)i(E,N)(=i(E,M)i(E,N)).

By Proposition 6.6 we know that

RHom𝒜(M,Ck(E,N))(RHom𝒜(E,N)AdRHom𝒜(M,E))k.{\rm RHom}_{\mathcal{A}}(M,C_{-k}(E,N))\simeq\left({\rm RHom}_{\mathcal{A}}(E,N)\otimes_{A_{d}}{\rm RHom}_{\mathcal{A}}(M,E)\right)_{-k}.

Moreover, by Remark 6.10 we know that we can replace RHom𝒜(E,N){\rm RHom}_{\mathcal{A}}(E,N) and RHom𝒜(M,E){\rm RHom}_{\mathcal{A}}(M,E) by quasi-isomorphic AdA_{d}-dg-modules. As i(E,M)i(E,M) and i(E,N)<i(E,N)<\infty, by Proposition 6.7 we know that RHom𝒜(E,N){\rm RHom}_{\mathcal{A}}(E,N) and RHom𝒜(M,E){\rm RHom}_{\mathcal{A}}(M,E) split as direct sums of shifts of KK and BrB_{r}’s, respectively CpC_{p}’s, for r,p1r,p\geq 1. Hence, to conclude it is enough to prove that

dimKH((T1AT2)k)dimKH(T1)dimKH(T2)\dim_{K}H^{\bullet}((T_{1}\otimes_{A}T_{2})_{-k})\geq\dim_{K}H^{\bullet}(T_{1})\cdot\dim_{K}H^{\bullet}(T_{2}) (6.15)

for T1{K,Br},T2{K,Cp}T_{1}\in\{K,B_{r}\},~{}T_{2}\in\{K,C_{p}\}.

If T1=T2=KT_{1}=T_{2}=K, then (KAdK)k(K\otimes_{A_{d}}K)_{-k} is a direct sum of shifts of KK, and therefore (6.15) becomes k1-k\geq 1, which is true.

If T1=KT_{1}=K and T2=CpT_{2}=C_{p}, then the right hand side of (6.15) is equal to dimKH(Cp)=2\dim_{K}H^{\bullet}(C_{p})=2. Hence, we have to find two non-zero cohomology classes in (KACp)k(K\otimes_{A}C_{p})_{-k}. We claim that the sought classes as given by

(0,0,,0,11)and(1(ϵ,0,0,,0),0,,0))\begin{array}[]{lcr}(0,0,\dots,0,1\otimes 1)&\mathrm{and}&(1\otimes(\epsilon,0,0,\dots,0),0,\dots,0))\end{array}

where we employed the notation we introduced in § 2.4 for elements belonging to the convolution of a twisted complex.

It is obvious that (0,0,,0,11)(0,0,\dots,0,1\otimes 1) is closed. For (1(ϵ,0,0,,0),0,,0)(1\otimes(\epsilon,0,0,\dots,0),0,\dots,0), it follows from the fact that 11 and (ϵ,0,0,,0)(\epsilon,0,0,\dots,0) are closed in KK and CpC_{p}, respectively, and the definition of the differential in (KACp)k(K\otimes_{A}C_{p})_{-k}:

d((1(ϵ,0,0,,0),0,,0)))\displaystyle d((1\otimes(\epsilon,0,0,\dots,0),0,\dots,0)))
=\displaystyle= (d(1(ϵ,0,0,,0)),1ϵ(ϵ,0,0,,0)+(1)k1(ϵ2,0,0,,0),0,,0)\displaystyle(d(1\otimes(\epsilon,0,0,\dots,0)),1\cdot\epsilon\otimes(\epsilon,0,0,\dots,0)+(-1)^{k}1\otimes(\epsilon^{2},0,0,\dots,0),0,\dots,0)
=\displaystyle= (0,,0).\displaystyle(0,\dots,0).

We prove that (0,0,,0,11)(0,0,\dots,0,1\otimes 1) gives a non-zero cohomology class, the proof for (1(ϵ,0,0,,0),0,,0)(1\otimes(\epsilon,0,0,\dots,0),0,\dots,0) is analogous. The element 111\otimes 1 is a non-zero cohomology class in KCpK\otimes C_{p}. Hence, if (0,0,,0,11)(0,0,\dots,0,1\otimes 1) is the differential of some element, it must be the differential of an element of the form (0,,0,a1b1,a0b0)(0,\dots,0,a_{-1}\otimes b_{-1},a_{0}\otimes b_{0}) with a1b10a_{-1}\otimes b_{-1}\neq 0. The differential of such an element is given by (we can assume we always have 11 on the left because the tensor products are KK-linear)

d((0,,0,1b1,1b0))=(0,,0,1d(b1),1d(b0)1ϵb1).d((0,\dots,0,1\otimes b_{-1},1\otimes b_{0}))=(0,\dots,0,1\otimes d(b_{-1}),1\otimes d(b_{0})-1\otimes\epsilon b_{-1}). (6.16)

For (6.16) to be equal to (0,0,,0,11)(0,0,\dots,0,1\otimes 1), we must have

1d(b0)1ϵb1=11d(b0)ϵb1=1.1\otimes d(b_{0})-1\otimes\epsilon b_{-1}=1\otimes 1\iff d(b_{0})-\epsilon b_{-1}=1.

However, by the definition of CpC_{p}, the equation d(b0)ϵb1=1d(b_{0})-\epsilon b_{-1}=1 is not satisfied by any b0,b1Cpb_{0},b_{-1}\in C_{p}, and therefore 111\otimes 1 is a non-zero cohomology class. Hence, we get dimKH((KACp)k)2\dim_{K}H^{\bullet}((K\otimes_{A}C_{p})_{-k})\geq 2, as we wanted.

If T1=BrT_{1}=B_{r} and T2=KT_{2}=K, the situation is as in the previous point.

If T1=BrT_{1}=B_{r} and T2=CpT_{2}=C_{p}, 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

(0,,0,11)((0,,0,ϵ)(0,,0,ϵ),0,,0)((0,,0,ϵ)(0,,0,1),0,,0)+(1)k+1((0,,0,1)(0,,0,ϵ),0,,0)((ϵ,0,,0)(ϵ,0,,0),0,,0)\begin{array}[]{c}(0,\dots,0,1\otimes 1)\\ ((0,\dots,0,\epsilon)\otimes(0,\dots,0,\epsilon),0,\dots,0)\\ ((0,\dots,0,\epsilon)\otimes(0,\dots,0,1),0,\dots,0)+(-1)^{k+1}((0,\dots,0,1)\otimes(0,\dots,0,\epsilon),0,\dots,0)\\ ((\epsilon,0,\dots,0)\otimes(\epsilon,0,\dots,0),0,\dots,0)\end{array}

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 XX be a K3 surface.

7.1. Hodge structures on Mukai lattices

The integral cohomology group H(X,)H^{*}(X,{\mathbb{Z}}) of XX has the lattice structure given by the Mukai pairing

((r1,c1,s1),(r2,c2,s2)):=c1c2r1s2r2s1((r_{1},c_{1},s_{1}),(r_{2},c_{2},s_{2})):=c_{1}\cdot c_{2}-r_{1}s_{2}-r_{2}s_{1}

for (r1,c1,s1),(r2,c2,s2)H(X,)(r_{1},c_{1},s_{1}),(r_{2},c_{2},s_{2})\in H^{*}(X,{\mathbb{Z}}). The lattice H(X,)H^{*}(X,{\mathbb{Z}}) called the Mukai lattice of XX is an even unimodular lattice of signature (4,20)(4,20). The Mukai lattice has a weight two Hodge structure H~(X,)\widetilde{H}(X,{\mathbb{Z}}) given by H~(X,)=p+q=2H~p,q(X)\widetilde{H}(X,{\mathbb{Z}})\otimes_{\mathbb{Z}}{\mathbb{C}}=\bigoplus_{p+q=2}\widetilde{H}^{p,q}(X) and

H~2,0(X):=H2,0(X),H~1,1(X):=p=02Hp,p(X),H~0,2(X):=H0,2(X).\widetilde{H}^{2,0}(X):=H^{2,0}(X),~{}\widetilde{H}^{1,1}(X):=\bigoplus_{p=0}^{2}H^{p,p}(X),~{}\widetilde{H}^{0,2}(X):=H^{0,2}(X).

This Hodge structure contains the ordinary Hodge structure on H2(X,)H^{2}(X,{\mathbb{Z}}) as a primitive sub-Hodge structure. The algebraic part of H~(X,)\widetilde{H}(X,{\mathbb{Z}}) denoted by (X){\mathbb{N}}(X), is equal to H0(X,)NS(X)H4(X,)H^{0}(X,{\mathbb{Z}})\oplus\mathrm{NS}(X)\oplus H^{4}(X,{\mathbb{Z}}) and has signature (2,ρ(X))(2,\rho(X)).

For an object E𝒟b(X)E\in{\mathcal{D}}^{b}(X), the Mukai vector v(E)H2(X,)v(E)\in H^{2*}(X,{\mathbb{Q}}) of EE is given by

v(E):=ch(E)tdX=(rk(E),c1(E),χ(E)rk(E)).v(E):=\mathrm{ch}(E)\sqrt{\mathrm{td}_{X}}=(\mathrm{rk}(E),c_{1}(E),\chi(E)-\mathrm{rk}(E)).

By the Riemann–Roch formula, we have the isomorphism v:𝒩(𝒟b(X))(X)v:{\mathcal{N}}({\mathcal{D}}^{b}(X))\xrightarrow{\sim}{\mathbb{N}}(X) satisfying (v(E),v(F))=χ(E,F)(v(E),v(F))=-\chi(E,F) for any objects E,F𝒟b(X)E,F\in{\mathcal{D}}^{b}(X), where 𝒩(𝒟b(X)){\mathcal{N}}({\mathcal{D}}^{b}(X)) is the numerical Grothendieck group of 𝒟b(X){\mathcal{D}}^{b}(X) and χ\chi is the Euler pairing on it.

7.2. Groups

A Hodge isometry φ:H~(X,)H~(X,)\varphi:\widetilde{H}(X,{\mathbb{Z}})\to\widetilde{H}(X,{\mathbb{Z}}) of H~(X,)\widetilde{H}(X,{\mathbb{Z}}) is an isomorphism of the Hodge structure preserving the Mukai pairing. The group of Hodge isometries is denoted by O(H~(X,))\mathrm{O}(\widetilde{H}(X,{\mathbb{Z}})). Let T(X)T(X) be the transcendental lattice of XX which is the transcendental part of the Hodge structures H~(X,)\widetilde{H}(X,{\mathbb{Z}}) and H2(X,)H^{2}(X,{\mathbb{Z}}). Restricting the Hodge structure H~(X,)\widetilde{H}(X,{\mathbb{Z}}) with the Mukai pairing to sub-Hodge structures H2(X,),NS(X)H^{2}(X,{\mathbb{Z}}),{\rm NS}(X) and T(X)T(X), we similarly define the groups of Hodge isometries O(H2(X,)),O(NS(X))\mathrm{O}(H^{2}(X,{\mathbb{Z}})),\mathrm{O}({\rm NS}(X)) and O(T(X))\mathrm{O}(T(X)), respectively. Then O(T(X))\mathrm{O}(T(X)) is a finite cyclic group, and faithfully acts on H2,0(X)H^{2,0}(X)\simeq{\mathbb{C}} by a root of unity.

Using the action of Aut(X){\rm Aut}(X) on H2(X,)H^{2}(X,{\mathbb{Z}}), the following two groups are defined

Auts(X)\displaystyle{\rm Aut}_{s}(X) :=\displaystyle:= {fAut(X)|H2(f)|H2,0(X)=idH2,0(X)}\displaystyle\left\{f\in{\rm Aut}(X)~{}\middle|~{}H^{2}(f)|_{H^{2,0}(X)}=\mathrm{id}_{H^{2,0}(X)}\right\}
=\displaystyle= {fAut(X)|H2(f)|T(X)=idT(X)}\displaystyle\left\{f\in{\rm Aut}(X)~{}\middle|~{}H^{2}(f)|_{T(X)}=\mathrm{id}_{T(X)}\right\}
Autt(X)\displaystyle{\rm Aut}_{t}(X) :=\displaystyle:= {fAut(X)|H2(f)|NS(X)=idNS(X)}.\displaystyle\left\{f\in{\rm Aut}(X)~{}\middle|~{}H^{2}(f)|_{{\rm NS}(X)}=\mathrm{id}_{{\rm NS}(X)}\right\}.

These two subgroups are normal. The group Auts(X)Aut(X){\rm Aut}_{s}(X)\subset{\rm Aut}(X) is of finite index, and its element is called a symplectic automorphism. The natural group homomorphisms

Auts(X)O(NS(X)) and Autt(X)O(T(X)).{\rm Aut}_{s}(X)\rightarrow\mathrm{O}(\mathrm{NS}(X))~{}~{}~{}~{}~{}\text{ and }~{}~{}~{}~{}~{}{\rm Aut}_{t}(X)\rightarrow\mathrm{O}(T(X)).

are injective, thus Autt(X){\rm Aut}_{t}(X) is also a finite cyclic group.

For any autoequivalence ΦAut(𝒟b(X))\Phi_{\mathcal{E}}\in{\rm Aut}({\mathcal{D}}^{b}(X)), we define the cohomological Fourier–Mukai transform ΦH:H(X,)H(X,)\Phi^{H}_{{\mathcal{E}}}:H^{*}(X,{\mathbb{Z}})\xrightarrow{\sim}H^{*}(X,{\mathbb{Z}}) associated to Φ\Phi_{\mathcal{E}} by

ΦH(v):=p(q(v)v()),\Phi^{H}_{\mathcal{E}}(v):=p_{*}(q^{*}(v)\cdot v({\mathcal{E}})),

which is a Hodge isometry of H~(X,)\widetilde{H}(X,{\mathbb{Z}}), thus induces the action

Aut(𝒟b(X))O(H~(X,)).{\rm Aut}({\mathcal{D}}^{b}(X))\rightarrow\mathrm{O}(\widetilde{H}(X,{\mathbb{Z}})).

The two subgroups Aut0(𝒟b(X)){\rm Aut}_{0}({\mathcal{D}}^{b}(X)) and AutCY(𝒟b(X)){\rm Aut}_{{\rm CY}}({\mathcal{D}}^{b}(X)) of Aut(𝒟b(X)){\rm Aut}({\mathcal{D}}^{b}(X)) are defined by

Aut0(𝒟b(X))\displaystyle{\rm Aut}_{0}({\mathcal{D}}^{b}(X)) :=\displaystyle:= {ΦAut(𝒟b(X))|ΦH=idH(X,)}\displaystyle\left\{\Phi\in{\rm Aut}({\mathcal{D}}^{b}(X))~{}\middle|~{}\Phi^{H}=\mathrm{id}_{H^{*}(X,{\mathbb{Z}})}\right\}
AutCY(𝒟b(X))\displaystyle{\rm Aut}_{{\rm CY}}({\mathcal{D}}^{b}(X)) :=\displaystyle:= {ΦAut(𝒟b(X))|ΦH|H2,0(X)=idH2,0(X)}\displaystyle\left\{\Phi\in{\rm Aut}({\mathcal{D}}^{b}(X))~{}\middle|~{}\Phi^{H}|_{H^{2,0}(X)}=\mathrm{id}_{H^{2,0}(X)}\right\}
=\displaystyle= {ΦAut(𝒟b(X))|ΦH|T(X)=idT(X)}.\displaystyle\left\{\Phi\in{\rm Aut}({\mathcal{D}}^{b}(X))~{}\middle|~{}\Phi^{H}|_{T(X)}=\mathrm{id}_{T(X)}\right\}.

These two subgroups are normal, and clearly Aut0(𝒟b(X))AutCY(𝒟b(X)){\rm Aut}_{0}({\mathcal{D}}^{b}(X))\subset{\rm Aut}_{{\rm CY}}({\mathcal{D}}^{b}(X)). The group AutCY(𝒟b(X))Aut(𝒟b(X)){\rm Aut}_{{\rm CY}}({\mathcal{D}}^{b}(X))\subset{\rm Aut}({\mathcal{D}}^{b}(X)) is of finite index, and its element is called a Calabi–Yau autoequivalence. We note that ΦAut(𝒟b(X))\Phi\in{\rm Aut}({\mathcal{D}}^{b}(X)) is Calabi–Yau if and only if it respects the Serre duality pairings

Hom(E,F[i])×Hom(F,E[2i]){\rm Hom}(E,F[i])\times{\rm Hom}(F,E[2-i])\to{\mathbb{C}}

induced by a choice of holomorphic volume forms in H2,0(X)H^{2,0}(X), see [BB17, Appendix A].

We recall the definition of centralizer and center groups.

Definition 7.1.

Let GG be a group, and HH a subgroup of GG.

  1. (i)(\rm{i})

    The centralizer group CG(H)C_{G}(H) of HH in GG is defined by

    CG(H):={gGgh=hg for all hH}.C_{G}(H):=\left\{g\in G\mid gh=hg\text{ for all }h\in H\right\}.
  2. (ii)(\rm{ii})

    The center group Z(G)Z(G) of GG is defined by

    Z(G):=CG(G).Z(G):=C_{G}(G).

It is easy to see Autt(X)Z(Aut(X)){\rm Aut}_{t}(X)\subset Z({\rm Aut}(X)), but these are not equal in general.

7.3. Stability conditions on 𝒟b(X){\mathcal{D}}^{b}(X)

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 XX be a K3 surface and fix a norm ||||||\cdot|| on 𝒩(𝒟b(X)){\mathcal{N}}({\mathcal{D}}^{b}(X))\otimes{\mathbb{R}}.

Definition 7.2 ([Bri07, Definition 5.1]).

A (numerical) stability condition σ=(Z,𝒫)\sigma=(Z,{\mathcal{P}}) on 𝒟b(X){\mathcal{D}}^{b}(X) consists of a group homomorphism Z:𝒩(𝒟b(X))Z:{\mathcal{N}}({\mathcal{D}}^{b}(X))\to{\mathbb{C}} called central charge and a family 𝒫={𝒫(ϕ)}ϕ{\mathcal{P}}=\{{\mathcal{P}}(\phi)\}_{\phi\in{\mathbb{R}}} of full additive subcategory of 𝒟b(X){\mathcal{D}}^{b}(X) called slicing, such that

  1. (i)(\rm{i})

    For 0E𝒫(ϕ)0\neq E\in{\mathcal{P}}(\phi), we have Z(v(E))=m(E)exp(iπϕ)Z(v(E))=m(E)\exp(i\pi\phi) for some m(E)>0m(E)\in{\mathbb{R}}_{>0}.

  2. (ii)(\rm{ii})

    For all ϕ\phi\in{\mathbb{R}}, we have 𝒫(ϕ+1)=𝒫(ϕ)[1]{\mathcal{P}}(\phi+1)={\mathcal{P}}(\phi)[1].

  3. (iii)(\rm{iii})

    For ϕ1>ϕ2\phi_{1}>\phi_{2} and Ei𝒫(ϕi)E_{i}\in{\mathcal{P}}(\phi_{i}), we have Hom(E1,E2)=0{\rm Hom}(E_{1},E_{2})=0.

  4. (iv)(\rm{iv})

    For each 0E𝒟b(X)0\neq E\in{\mathcal{D}}^{b}(X), there is a collection of exact triangles called Harder–Narasimhan filtration of EE:

    0=E0\textstyle{0=E_{0}}E1\textstyle{E_{1}}\textstyle{\dots}Ep1\textstyle{E_{p-1}}Ep=E\textstyle{E_{p}=E}A1\textstyle{A_{1}}\textstyle{\dots}Ap\textstyle{A_{p}} (7.1)

    with Ai𝒫(ϕi)A_{i}\in{\mathcal{P}}(\phi_{i}) and ϕ1>ϕ2>>ϕp\phi_{1}>\phi_{2}>\cdots>\phi_{p}.

  5. (v)(\rm{v})

    (support property) There exists a constant C>0C>0 such that for all 0E𝒫(ϕ)0\neq E\in{\mathcal{P}}(\phi), we have

    E<C|Z(E)|.||E||<C|Z(E)|. (7.2)

For any interval II\subset{\mathbb{R}}, define 𝒫(I){\mathcal{P}}(I) to be the extension-closed subcategory of 𝒟b(X){\mathcal{D}}^{b}(X) generated by the subcategories 𝒫(ϕ){\mathcal{P}}(\phi) for ϕI\phi\in I. Then 𝒫((0,1]){\mathcal{P}}((0,1]) is the heart of a bounded t-structure on 𝒟b(X){\mathcal{D}}^{b}(X), hence an abelian category. The full subcategory 𝒫(ϕ)𝒟b(X){\mathcal{P}}(\phi)\subset{\mathcal{D}}^{b}(X) is also shown to be abelian. A non-zero object E𝒫(ϕ)E\in{\mathcal{P}}(\phi) is called σ\sigma-semistable of phase ϕσ(E):=ϕ\phi_{\sigma}(E):=\phi, and especially a simple object in 𝒫(ϕ){\mathcal{P}}(\phi) is called σ\sigma-stable. Taking the Harder–Narasimhan filtration (7.1) of EE, we define ϕσ+(E):=ϕσ(A1)\phi^{+}_{\sigma}(E):=\phi_{\sigma}(A_{1}) and ϕσ(E):=ϕσ(Ap)\phi^{-}_{\sigma}(E):=\phi_{\sigma}(A_{p}). The object AiA_{i} is called σ\sigma-semistable factor of EE. Define Stab(X){\rm Stab}(X) to be the set of numerical stability conditions on 𝒟b(X){\mathcal{D}}^{b}(X).

We prepare some terminologies on the stability on the heart of a tt-structure on 𝒟b(X){\mathcal{D}}^{b}(X).

Definition 7.3.

Let 𝒜{\mathcal{A}} be the heart of a bounded tt-structure on 𝒟b(X){\mathcal{D}}^{b}(X). A stability function on 𝒜{\mathcal{A}} is a group homomorphism Z:𝒩(𝒟b(X))Z:{\mathcal{N}}({\mathcal{D}}^{b}(X))\to{\mathbb{C}} such that for all 0E𝒜𝒟b(X)0\neq E\in{\mathcal{A}}\subset{\mathcal{D}}^{b}(X), the complex number Z(E)Z(E) lies in the semiclosed upper half plane :={reiπϕ|r>0,ϕ(0,1]}{\mathbb{H}}_{-}:=\{re^{i\pi\phi}\in{\mathbb{C}}~{}|~{}r\in{\mathbb{R}}_{>0},\phi\in(0,1]\}\subset{\mathbb{C}}.

Given a stability function Z:𝒩(𝒟b(X))Z:{\mathcal{N}}({\mathcal{D}}^{b}(X))\to{\mathbb{C}} on 𝒜{\mathcal{A}}, the phase of an object 0E𝒜0\neq E\in{\mathcal{A}} is defined to be ϕ(E):=1πargZ(E)(0,1]\phi(E):=\frac{1}{\pi}{\rm arg}Z(E)\in(0,1]. An object 0E𝒜0\neq E\in{\mathcal{A}} is ZZ-semistable (resp. ZZ-stable) if for all subobjects 0AE0\neq A\subset E, we have ϕ(A)ϕ(E)\phi(A)\leq\phi(E) (resp. ϕ(A)<ϕ(E)\phi(A)<\phi(E)). We say that a stability function ZZ satisfies the Harder–Narasimhan property if each object 0E𝒜0\neq E\in{\mathcal{A}} admits a filtration (called Harder–Narasimhan filtration of EE) 0=E0E1E2Em=E0=E_{0}\subset E_{1}\subset E_{2}\subset\cdots\subset E_{m}=E such that Ei/Ei1E_{i}/E_{i-1} is ZZ-semistable for i=1,,mi=1,\cdots,m with ϕ(E1/E0)>ϕ(E2/E1)>>ϕ(Em/Em1)\phi(E_{1}/E_{0})>\phi(E_{2}/E_{1})>\cdots>\phi(E_{m}/E_{m-1}), and the support property if there exists a constant C>0C>0 such that for all ZZ-semistable objects E𝒜E\in{\mathcal{A}}, we have E<C|Z(E)|||E||<C|Z(E)|.

The following proposition shows the relationship between stability conditions and stability functions on the heart of a bounded tt-structure.

Proposition 7.4 ([Bri07, Proposition 5.3]).

To give a stability condition on 𝒟b(X){\mathcal{D}}^{b}(X) is equivalent to giving the heart 𝒜{\mathcal{A}} of a bounded t-structure on 𝒟b(X){\mathcal{D}}^{b}(X), and a stability function ZZ on 𝒜{\mathcal{A}} with the Harder–Narasimhan property and the support property.

For the proof, we construct the slicing 𝒫{\mathcal{P}}, from the pair (Z,𝒜)(Z,{\mathcal{A}}), by

𝒫(ϕ):={E𝒜|E is Z-semistable with ϕ(E)=ϕ} for ϕ(0,1],{\mathcal{P}}(\phi):=\{E\in{\mathcal{A}}~{}|~{}E\text{ is }Z\text{-semistable with }\phi(E)=\phi\}\text{ for }\phi\in(0,1],

and extend for all ϕ\phi\in{\mathbb{R}} by 𝒫(ϕ+1):=𝒫(ϕ)[1]{\mathcal{P}}(\phi+1):={\mathcal{P}}(\phi)[1]. Conversely, for a stability condition σ=(Z,𝒫)\sigma=(Z,{\mathcal{P}}), the heart 𝒜{\mathcal{A}} is given by 𝒜:=𝒫σ((0,1]){\mathcal{A}}:={\mathcal{P}}_{\sigma}((0,1]). We also denote stability conditions by (Z,𝒜)(Z,{\mathcal{A}}).

We recall that the Mukai vector v:𝒩(𝒟b(X))(X)v:{\mathcal{N}}({\mathcal{D}}^{b}(X))\xrightarrow{\sim}{\mathbb{N}}(X) and the Mukai pairing (,)(-,-) on (X)H~(X,){\mathbb{N}}(X)\subset\widetilde{H}(X,{\mathbb{Z}}), then the central charge of a numerical stability condition takes the form Z()=(Ω,v())Z(-)=(\Omega,v(-)) for some Ω(X)\Omega\in{\mathbb{N}}(X)\otimes{\mathbb{C}}. Bridgeland constructed a family of stability conditions on 𝒟b(X){\mathcal{D}}^{b}(X) as follows: Let

V(X)\displaystyle V(X) :=\displaystyle:= {β+iωNS(X)|β,ωNS(X),ω:-ample,\displaystyle\{\beta+i\omega\in{\rm NS}(X)\otimes{\mathbb{C}}~{}|~{}\beta,\omega\in{\rm NS(X)}\otimes{\mathbb{R}},~{}\omega:{\mathbb{R}}\text{-ample},
(exp(β+iω),v(E))0 for all spherical sheaves E}.\displaystyle({\rm exp}(\beta+i\omega),v(E))\notin{\mathbb{R}}_{\leq 0}\text{ for all spherical sheaves }E\}.

For β+iωV(X)\beta+i\omega\in V(X), we set Zβ,ω():=(exp(β+iω),v())Z_{\beta,\omega}(-):=({\rm exp}(\beta+i\omega),v(-)). The category 𝒜β,ω{\mathcal{A}}_{\beta,\omega} is the heart of a t-structure obtained by tilting the standard t-structure with respect to the torsion pair (𝒯β,ω,β,ω)({\mathcal{T}}_{\beta,\omega},{\mathcal{F}}_{\beta,\omega}) on Coh(X){\rm Coh}(X) given by

𝒯β,ω\displaystyle{\mathcal{T}}_{\beta,\omega} :=\displaystyle:= {ECoh(X)|E is a torsion sheaf or μω(E/torsion part)>β.ω}\displaystyle\{E\in{\rm Coh}(X)~{}|~{}E\text{ is a torsion sheaf or }\mu_{\omega}^{-}(E/\text{torsion part})>\beta.\omega\}
β,ω\displaystyle{\mathcal{F}}_{\beta,\omega} :=\displaystyle:= {ECoh(X)|E is torsion free and μω+(E)β.ω},\displaystyle\{E\in{\rm Coh}(X)~{}|~{}E\text{ is torsion free and }\mu_{\omega}^{+}(E)\leq\beta.\omega\},

where for a torsion free sheaf EE, μω+(E)\mu_{\omega}^{+}(E) (resp. μω(E)\mu_{\omega}^{-}(E)) is the maximal (resp. minimal) slope of μω\mu_{\omega}-semistable factors of EE. Then (Zβ,ω,𝒜β,ω)(Z_{\beta,\omega},{\mathcal{A}}_{\beta,\omega}) is a stability condition on 𝒟b(X){\mathcal{D}}^{b}(X) ([Bri08, Lemma 6.2 and Proposition 11.2]).

Let E𝒟b(X)E\in{\mathcal{D}}^{b}(X) be a non-zero object of 𝒟b(X){\mathcal{D}}^{b}(X) and σStab(X)\sigma\in{\rm Stab}(X) be a stability condition on 𝒟b(X){\mathcal{D}}^{b}(X). The mass mσ(E)>0m_{\sigma}(E)\in{\mathbb{R}}_{>0} of EE is defined by

mσ(E):=i=1p|Zσ(Ai)|,m_{\sigma}(E):=\displaystyle\sum_{i=1}^{p}|Z_{\sigma}(A_{i})|,

where A1,,ApA_{1},\cdots,A_{p} are σ\sigma-semistable factors of EE. The following generalized metric (i.e. with values in [0,][0,\infty]) dBd_{B} on Stab(X){\rm Stab}(X) is defined by Bridgeland ([Bri07, Proposition 8.1]):

dB(σ,τ):=supE0{|ϕσ+(E)ϕτ+(E)|,|ϕσ(E)ϕτ(E)|,|logmσ(E)mτ(E)|}[0,].d_{B}(\sigma,\tau):=\sup_{E\neq 0}\left\{|\phi^{+}_{\sigma}(E)-\phi^{+}_{\tau}(E)|,|\phi^{-}_{\sigma}(E)-\phi^{-}_{\tau}(E)|,\middle|\log\frac{m_{\sigma}(E)}{m_{\tau}(E)}\middle|\right\}\in[0,\infty].

This generalized metric induces the topology on Stab(X){\rm Stab}(X). Then the generalized metric dBd_{B} takes a finite value on each connected component Stab(X){\rm Stab}^{\circ}(X) of Stab(X){\rm Stab}(X), thus (Stab(X),dB)({\rm Stab}^{\circ}(X),d_{B}) is a metric space in the strict sense.

Theorem 7.5 ([Bri07, Theorem 7.1]).

The map

Stab(X)(X);σ=((Ω,v()),𝒫)Ω{\rm Stab}(X)\to{\mathbb{N}}(X)\otimes{\mathbb{C}};~{}\sigma=((\Omega,v(-)),{\mathcal{P}})\mapsto\Omega (7.3)

is a local homeomorphism, where (X){\mathbb{N}}(X)\otimes{\mathbb{C}} is equipped with the natural linear topology.

Therefore the space Stab(X){\rm Stab}(X) (and each connected component Stab(X){\rm Stab}^{\circ}(X)) naturally admits a structure of finite dimensional complex manifolds.

There is a left action of Aut(𝒟b(X)){\rm Aut}({\mathcal{D}}^{b}(X)) on Stab(X){\rm Stab}(X) given by

F.σ:=(Zσ(F1()),{F(𝒫σ(ϕ))}) for σStab(X),FAut(𝒟b(X)).F.\sigma:=(Z_{\sigma}(F^{-1}(-)),\{F({\mathcal{P}}_{\sigma}(\phi))\})~{}\text{ for }\sigma\in{\rm Stab}(X),~{}F\in{\rm Aut}({\mathcal{D}}^{b}(X)). (7.4)

This action of Aut(𝒟b(X)){\rm Aut}({\mathcal{D}}^{b}(X)) is isometric with respect to dBd_{B}.

Let Stab(X){\rm Stab}^{\dagger}(X) be the connected component of Stab(X){\rm Stab}(X) 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 (Zβ,ω,𝒜β,ω)(Z_{\beta,\omega},{\mathcal{A}}_{\beta,\omega}) for each β+iωV(X)\beta+i\omega\in V(X) is geometric.

Definition 7.6.

The group Aut(𝒟b(X)){\rm Aut}^{\dagger}({\mathcal{D}}^{b}(X)) is defined as the subgroup of Aut(𝒟b(X)){\rm Aut}({\mathcal{D}}^{b}(X)) which preserves the connected component Stab(X){\rm Stab}^{\dagger}(X).

Proposition 7.7 ([Har12, proof of Proposition 7.9]).

The following autoequivalences preserve the distinguished component Stab(X){\rm Stab}^{\dagger}(X) i.e. are elements in Aut(𝒟b(X)){\rm Aut}^{\dagger}({\mathcal{D}}^{b}(X)):

  • shift [n][n] for nn\in{\mathbb{Z}}

  • line bundle tensor -\otimes\mathcal{L} for Pic(X)\mathcal{L}\in{\rm Pic}(X)

  • pullback ff^{*} for fAut(X)f\in{\rm Aut}(X)

  • The composition

    gΦ1:𝒟b(X)𝒟b(M)𝒟b(X),g^{*}\circ\Phi_{\mathcal{E}}^{-1}:~{}{\mathcal{D}}^{b}(X)\to{\mathcal{D}}^{b}(M)\to{\mathcal{D}}^{b}(X),

    where MM is a 22-dimensional fine compact moduli space of Gieseker-stable torsion free sheaves on XX with the universal family {\mathcal{E}} such that g:XMg:X\xrightarrow{\sim}M.

  • spherical twist TAT_{A} along a Gieseker-stable spherical bundle AA (e.g. 𝒪X\mathcal{O}_{X}).

  • spherical twist T𝒪CT_{\mathcal{O}_{C}} along 𝒪C\mathcal{O}_{C} for any (2)(-2)-curve CC on XX

We additionally define the following groups

AutCY(𝒟b(X))\displaystyle{\rm Aut}^{\dagger}_{{\rm CY}}({\mathcal{D}}^{b}(X)) :=\displaystyle:= AutCY(𝒟b(X))Aut(𝒟b(X))\displaystyle{\rm Aut}_{{\rm CY}}({\mathcal{D}}^{b}(X))\cap{\rm Aut}^{\dagger}({\mathcal{D}}^{b}(X))
Aut0(𝒟b(X))\displaystyle{\rm Aut}^{\dagger}_{0}({\mathcal{D}}^{b}(X)) :=\displaystyle:= Aut0(𝒟b(X))Aut(𝒟b(X))\displaystyle{\rm Aut}_{0}({\mathcal{D}}^{b}(X))\cap{\rm Aut}^{\dagger}({\mathcal{D}}^{b}(X))

Define the open subset 𝒫(X)(X){\mathcal{P}}(X)\subset{\mathbb{N}}(X)\otimes{\mathbb{C}} consisting of vectors Ω(X)\Omega\in{\mathbb{N}}(X)\otimes{\mathbb{C}} whose real and imaginary parts span a positive definite 2-plane (ReΩ,ImΩ,(,))(\langle{\rm Re}\Omega,{\rm Im}\Omega\rangle_{\mathbb{R}},(-,-)) in (X){\mathbb{N}}(X)\otimes{\mathbb{R}}. This subset has two connected components, distinguished by the orientation induced on this 2-plane; let 𝒫+(X){\mathcal{P}}^{+}(X) be the component containing vectors of the form (1,iω,12ω2)(1,i\omega,-\frac{1}{2}\omega^{2}) for an ample class ωNS(X)\omega\in{\rm NS}(X)\otimes{\mathbb{R}}. Consider the root system

Δ(X):={δ(X)|(δ,δ)=2}\Delta(X):=\{\delta\in{\mathbb{N}}(X)~{}|~{}(\delta,\delta)=-2\}

consisting of (2)(-2)-classes in (X){\mathbb{N}}(X), and the corresponding hyperplane complement

𝒫0+(X):=𝒫+(X)\δΔ(X)δ.{\mathcal{P}}_{0}^{+}(X):={\mathcal{P}}^{+}(X)\backslash\bigcup_{\delta\in\Delta(X)}\delta^{\bot}.
Theorem 7.8 ([Bri08, Theorem 1.1]).

The map

π:Stab(X)𝒫0+(X);((Ω,v()),𝒫)Ω\pi:{\rm Stab}^{\dagger}(X)\to{\mathcal{P}}_{0}^{+}(X);~{}((\Omega,v(-)),{\mathcal{P}})\mapsto\Omega

is a normal covering, and the group Aut0(𝒟b(X)){\rm Aut}^{\dagger}_{0}({\mathcal{D}}^{b}(X)) is identified with the group of deck transformations of π\pi.

The following is a Bridgeland conjecture on the space of stability conditions on 𝒟b(X){\mathcal{D}}^{b}(X) and the action on it of Aut(𝒟b(X)){\rm Aut}({\mathcal{D}}^{b}(X)).

Conjecture 7.9 ([Bri08, Conjecture 1.2]).

Let XX be a K3 surface. Then

  1. (i)(\rm{i})

    Stab(X){\rm Stab}^{\dagger}(X) is simply-connected.

  2. (ii)(\rm{ii})

    any autoequivalence of 𝒟b(X){\mathcal{D}}^{b}(X) preserves the distinguished component Stab(X){\rm Stab}^{\dagger}(X), equivalently we have

    Aut(𝒟b(X))=Aut(𝒟b(X)).{\rm Aut}^{\dagger}({\mathcal{D}}^{b}(X))={\rm Aut}({\mathcal{D}}^{b}(X)).

This conjecture clearly implies an isomorphism

π1(𝒫0+(X))Aut0(𝒟b(X)).\pi_{1}({\mathcal{P}}_{0}^{+}(X))\simeq{\rm Aut}_{0}({\mathcal{D}}^{b}(X)).
Theorem 7.10 ([BB17, Theorem 1.3]).

Conjecture 7.9 holds in the case of Picard rank one.

More strongly, Bayer–Bridgeland proved the contractibility of Stab(X){\rm Stab}^{\dagger}(X) in [BB17].

In Section 8, we consider the center groups Z(Aut(𝒟b(X)))Z({\rm Aut}^{\dagger}({\mathcal{D}}^{b}(X))), Z(AutCY(𝒟b(X)))Z({\rm Aut}^{\dagger}_{{\rm CY}}({\mathcal{D}}^{b}(X))), and the centralizer group CAut(𝒟b(X))(AutCY(𝒟b(X)))C_{{\rm Aut}^{\dagger}({\mathcal{D}}^{b}(X))}({\rm Aut}^{\dagger}_{{\rm CY}}({\mathcal{D}}^{b}(X))).

8. The center of autoequivalence groups of K3 surfaces

Let XX be a K3 surface of any Picard rank.

8.1. Main results

We compute the center groups of Aut(𝒟b(X)){\rm Aut}^{\dagger}({\mathcal{D}}^{b}(X)) and AutCY(𝒟b(X)){\rm Aut}_{{\rm CY}}^{\dagger}({\mathcal{D}}^{b}(X)) and the centralizer groups CAut(𝒟b(X))(AutCY(𝒟b(X)))C_{{\rm Aut}^{\dagger}({\mathcal{D}}^{b}(X))}({\rm Aut}_{{\rm CY}}^{\dagger}({\mathcal{D}}^{b}(X))).

Theorem 8.1.

Let XX be a K3 surface, mXm_{X} the order of Autt(X){\rm Aut}_{t}(X), and ftf_{t} a generator of Autt(X){\rm Aut}_{t}(X). Then we have the following

  1. (i)(\rm{i})

    Z(Aut(𝒟b(X)))=CAut(𝒟b(X))(AutCY(𝒟b(X)))=Autt(X)×[1](/mX)×Z({\rm Aut}^{\dagger}({\mathcal{D}}^{b}(X)))=C_{{\rm Aut}^{\dagger}({\mathcal{D}}^{b}(X))}({\rm Aut}^{\dagger}_{{\rm CY}}({\mathcal{D}}^{b}(X)))={\rm Aut}_{t}(X)\times{\mathbb{Z}}[1]\simeq({\mathbb{Z}}/m_{X})\times{\mathbb{Z}}.

  2. (ii)(\rm{ii})
    Z(AutCY(𝒟b(X)))\displaystyle Z({\rm Aut}^{\dagger}_{{\rm CY}}({\mathcal{D}}^{b}(X))) =\displaystyle= {(ft)mX/2[1]if mX is even[2]if mX is odd\displaystyle\begin{cases}\langle~{}(f_{t}^{*})^{m_{X}/2}\circ[1]~{}\rangle&\mbox{if }m_{X}\mbox{ is even}\\ {\mathbb{Z}}[2]&\mbox{if }m_{X}\mbox{ is odd}\end{cases}
    \displaystyle\simeq .\displaystyle{\mathbb{Z}}.
Proof.
  1. (i)(\rm{i})

    We note that Aut(X)Pic(X)Aut(𝒟b(X)){\rm Aut}(X)\ltimes{\rm Pic}(X)\subset{\rm Aut}^{\dagger}({\mathcal{D}}^{b}(X)) and T𝒪XAut(𝒟b(X))T_{\mathcal{O}_{X}}\in{\rm Aut}^{\dagger}({\mathcal{D}}^{b}(X)), see Proposition 7.7. Fix any ΦZ(Aut(𝒟b(X)))\Phi\in Z({\rm Aut}^{\dagger}({\mathcal{D}}^{b}(X))). By Theorem 4.6, the relation ΦT𝒪X=T𝒪XΦ\Phi\circ T_{\mathcal{O}_{X}}=T_{\mathcal{O}_{X}}\circ\Phi implies Φ(𝒪X)=𝒪X[i]\Phi(\mathcal{O}_{X})=\mathcal{O}_{X}[i] for some ii\in{\mathbb{Z}}. For any line bundle Pic(X)\mathcal{L}\in{\rm Pic}(X) on XX, we have (Φ[i])()=(\Phi\circ[-i])(\mathcal{L})=\mathcal{L} by Φ()=()Φ\Phi\circ(-\otimes\mathcal{L})=(-\otimes\mathcal{L})\circ\Phi. By similar arguments as in the proof of [Huy12, Lemma A.2], Φ[i]\Phi\circ[-i] is in Autt(X){\rm Aut}_{t}(X), hence ΦAutt(X),[1]Autt(X)×[1]\Phi\in\langle{\rm Aut}_{t}(X),[1]\rangle\simeq{\rm Aut}_{t}(X)\times{\mathbb{Z}}[1]. It remains to show that Autt(X)Z(Aut(𝒟b(X))){\rm Aut}_{t}(X)\subset Z({\rm Aut}^{\dagger}({\mathcal{D}}^{b}(X))). Fix any ΨAut(𝒟b(X))\Psi\in{\rm Aut}^{\dagger}({\mathcal{D}}^{b}(X)) and set

    Ψt:=ΨftΨ1(ft)1Aut(𝒟b(X)).\Psi_{t}:=\Psi\circ f_{t}^{*}\circ\Psi^{-1}\circ(f_{t}^{*})^{-1}\in{\rm Aut}^{\dagger}({\mathcal{D}}^{b}(X)).

    Since the transcendental part of the Hodge structure H~(X,)\widetilde{H}(X,{\mathbb{Z}}) is equal to T(X)T(X), (X)T(X)H~(X,){\mathbb{N}}(X)\oplus T(X)\subset\widetilde{H}(X,{\mathbb{Z}}) is of finite index, so that ΨtAut0(𝒟b(X))\Psi_{t}\in{\rm Aut}^{\dagger}_{0}({\mathcal{D}}^{b}(X)). By [Huy12, Lemma A.3], ftf_{t}^{*} acts on Stab(X){\rm Stab}^{\dagger}(X) trivially, hence Ψt\Psi_{t} also does. We thus have Ψt=id𝒟b(X)\Psi_{t}={\rm id}_{{\mathcal{D}}^{b}(X)} since Aut0(𝒟b(X)){\rm Aut}^{\dagger}_{0}({\mathcal{D}}^{b}(X)) is isomorphic to the group of deck transformations of the normal cover Stab(X)𝒫0+(X){\rm Stab}^{\dagger}(X)\to{\mathcal{P}}^{+}_{0}(X), see Theorem 7.8. We therefore have ftZ(Aut(𝒟b(X)))f_{t}^{*}\in Z({\rm Aut}^{\dagger}({\mathcal{D}}^{b}(X))), thus Z(Aut(𝒟b(X)))=Autt(X)×[1]Z({\rm Aut}^{\dagger}({\mathcal{D}}^{b}(X)))={\rm Aut}_{t}(X)\times{\mathbb{Z}}[1].

    By T𝒪X,AutCY(𝒟b(X))T_{\mathcal{O}_{X}},~{}-\otimes\mathcal{L}\in{\rm Aut}^{\dagger}_{{\rm CY}}({\mathcal{D}}^{b}(X)), similar arguments give

    CAut(𝒟b(X))(AutCY(𝒟b(X)))Autt(X)×[1].C_{{\rm Aut}^{\dagger}({\mathcal{D}}^{b}(X))}({\rm Aut}^{\dagger}_{{\rm CY}}({\mathcal{D}}^{b}(X)))\subset{\rm Aut}_{t}(X)\times{\mathbb{Z}}[1].

    The other inclusion follows from

    Autt(X)×[1]=Z(Aut(𝒟b(X)))CAut(𝒟b(X))(AutCY(𝒟b(X))).{\rm Aut}_{t}(X)\times{\mathbb{Z}}[1]=Z({\rm Aut}^{\dagger}({\mathcal{D}}^{b}(X)))\subset C_{{\rm Aut}^{\dagger}({\mathcal{D}}^{b}(X))}({\rm Aut}^{\dagger}_{{\rm CY}}({\mathcal{D}}^{b}(X))).
  2. (ii)(\rm{ii})

    By (i), we have

    Z(AutCY(𝒟b(X)))\displaystyle Z({\rm Aut}^{\dagger}_{{\rm CY}}({\mathcal{D}}^{b}(X))) =\displaystyle= CAut(𝒟b(X))(AutCY(𝒟b(X)))AutCY(𝒟b(X))\displaystyle C_{{\rm Aut}^{\dagger}({\mathcal{D}}^{b}(X))}({\rm Aut}^{\dagger}_{{\rm CY}}({\mathcal{D}}^{b}(X)))\cap{\rm Aut}^{\dagger}_{{\rm CY}}({\mathcal{D}}^{b}(X))
    =\displaystyle= (Autt(X)×[1])AutCY(𝒟b(X)).\displaystyle({\rm Aut}_{t}(X)\times{\mathbb{Z}}[1])\cap{\rm Aut}^{\dagger}_{{\rm CY}}({\mathcal{D}}^{b}(X)).

    Each element of Autt(X)×[1]{\rm Aut}_{t}(X)\times{\mathbb{Z}}[1] is of the form (ft)l[i](f_{t}^{*})^{l}\circ[i] for some l/mXl\in{\mathbb{Z}}/m_{X} and ii\in{\mathbb{Z}}. Then ((ft)l[i])H|H2,0(X)=(1)i(ftl)H|H2,0(X)((f_{t}^{*})^{l}\circ[i])^{H}|_{H^{2,0}(X)}=(-1)^{i}(f_{t}^{*l})^{H}|_{H^{2,0}(X)} is equal to idH2,0(X){\rm id}_{H^{2,0}(X)} if and only if (a) ii is odd and (ftl)H|H2,0(X)=idH2,0(X)(f_{t}^{*l})^{H}|_{H^{2,0}(X)}=-{\rm id}_{H^{2,0}(X)}, or (b) ii is even and (ftl)H|H2,0(X)=idH2,0(X)(f_{t}^{*l})^{H}|_{H^{2,0}(X)}={\rm id}_{H^{2,0}(X)} i.e. l=0l=0. By the faithfulness of the action of Autt(X){\rm Aut}_{t}(X) on H2,0(X)H^{2,0}(X), the case (a) is realized only when mXm_{X} is even and l=mX2l=\frac{m_{X}}{2}. We therefore have

    (Autt(X)×[1])AutCY(𝒟b(X))={(ft)mX/2[1]if mX is even[2]if mX is odd,({\rm Aut}_{t}(X)\times{\mathbb{Z}}[1])\cap{\rm Aut}^{\dagger}_{{\rm CY}}({\mathcal{D}}^{b}(X))=\begin{cases}\langle~{}(f_{t}^{*})^{m_{X}/2}\circ[1]~{}\rangle&\mbox{if }m_{X}\mbox{ is even}\\ {\mathbb{Z}}[2]&\mbox{if }m_{X}\mbox{ is odd},\end{cases}

    which completes the proof.

When Conjecture 7.9 (i) and (ii) are true, the group AutCY(𝒟b(X))/[2]{\rm Aut}^{\dagger}_{{\rm CY}}({\mathcal{D}}^{b}(X))/{\mathbb{Z}}[2] is naturally isomorphic to the orbifold fundamental group of the stringy Kähler moduli space of XX ([BB17, Section 7] and [Huy16, Conjecture 3.14]), so this quotient group is important in the context of mirror symmetry.

Corollary 8.2.

Let XX be a K3 surface. Then we have

Z(AutCY(𝒟b(X))/[2])\displaystyle Z({\rm Aut}^{\dagger}_{{\rm CY}}({\mathcal{D}}^{b}(X))/{\mathbb{Z}}[2]) \displaystyle\simeq Z(AutCY(𝒟b(X)))/[2]\displaystyle Z({\rm Aut}^{\dagger}_{{\rm CY}}({\mathcal{D}}^{b}(X)))/{\mathbb{Z}}[2]
\displaystyle\simeq {/2if mX is eventrivialif mX is odd.\displaystyle\begin{cases}{\mathbb{Z}}/2&\mbox{if }m_{X}\mbox{ is even}\\ trivial&\mbox{if }m_{X}\mbox{ is odd}.\end{cases}
Proof.

Let φ:Z(AutCY(𝒟b(X)))Z(AutCY(𝒟b(X))/[2])\varphi:Z({\rm Aut}^{\dagger}_{{\rm CY}}({\mathcal{D}}^{b}(X)))\to Z({\rm Aut}^{\dagger}_{{\rm CY}}({\mathcal{D}}^{b}(X))/{\mathbb{Z}}[2]) be a natural group homomorphism induced by the quotient. By Theorem 8.1(ii), kerφ=[2]{\rm ker}~{}\varphi={\mathbb{Z}}[2] whether mXm_{X} is even or not. We note that Φ¯Z(AutCY(𝒟b(X))/[2])\overline{\Phi}\in Z({\rm Aut}^{\dagger}_{{\rm CY}}({\mathcal{D}}^{b}(X))/{\mathbb{Z}}[2]) if and only if any ΨAutCY(𝒟b(X))\Psi\in{\rm Aut}^{\dagger}_{{\rm CY}}({\mathcal{D}}^{b}(X)) satisfies ΨΦΨ1Φ1=[2l]\Psi\circ\Phi\circ\Psi^{-1}\circ\Phi^{-1}=[2l] for some ll\in{\mathbb{Z}}.

Let

h():×Aut(𝒟b(X));(t,Ψ)ht(Ψ)h_{-}(-):{\mathbb{R}}\times{\rm Aut}({\mathcal{D}}^{b}(X))\to{\mathbb{R}};~{}(t,\Psi^{\prime})\mapsto h_{t}(\Psi^{\prime})

be the categorical entropy ([DHKK14, Definition 2.5]). Then by [KST20, Lemma 2.7(v) and Corollary 2.10], we have

ht(ΨΦΨ1)\displaystyle h_{t}(\Psi\circ\Phi\circ\Psi^{-1}) =\displaystyle= ht(Φ)\displaystyle h_{t}(\Phi)
ht(Φ[2l])\displaystyle h_{t}(\Phi[2l]) =\displaystyle= ht(Φ)+2lt\displaystyle h_{t}(\Phi)+2lt

for all tt\in{\mathbb{R}}, which concludes that l=0l=0. The morphism φ\varphi is therefore surjective, which completes the proof. ∎

8.2. Examples

We here collect several examples of the order mXm_{X} of the finite cyclic group Autt(X){\rm Aut}_{t}(X), which determines the center groups by Theorem 8.1 and Corollary 8.2.

Example 8.3.

Let XX be a K3 surface of odd Picard rank. Then O(T(X))\mathrm{O}(T(X)) is isomorphic to /2{\mathbb{Z}}/2 (cf. [Huy16, Cor 3.3.5]). Therefore mX=2m_{X}=2 (resp. mX=1m_{X}=1) if and only if Autt(X)=O(T(X)){\rm Aut}_{t}(X)=\mathrm{O}(T(X)) (resp. Autt(X)={idX}{\rm Aut}_{t}(X)=\{{\rm id}_{X}\}).

Example 8.4.

Let XX be a K3 surface of Picard rank 1, and HH the ample generator of NS(X)\mathrm{NS}(X). As a special case of Example 8.3, we have the following:

  1. (i)(\rm{i})

    If H2=2H^{2}=2, then Aut(X)=Autt(X)=i=O(T(X)){\rm Aut}(X)={\rm Aut}_{t}(X)=\langle i^{*}\rangle=\mathrm{O}(T(X)), where iAutt(X)i\in{\rm Aut}_{t}(X) is the covering involution of the double cover X2X\to{\mathbb{P}}^{2} branched along a smooth curve of degree six.

  2. (ii)(\rm{ii})

    If H2>2H^{2}>2, then Aut(X)=Autt(X)={idX}O(T(X)){\rm Aut}(X)={\rm Aut}_{t}(X)=\{\mathrm{id}_{X}\}\neq\mathrm{O}(T(X)).

Moreover as mentioned in Theorem 7.10, we have Aut(𝒟b(X))=Aut(𝒟b(X)){\rm Aut}^{\dagger}({\mathcal{D}}^{b}(X))={\rm Aut}({\mathcal{D}}^{b}(X)).

Example 8.5.

Let XX be a K3 surface of Picard rank 2, with infinite automorphism group. Then by [GLP10, Corollary 1], Aut(X){\rm Aut}(X) is isomorphic to {\mathbb{Z}} or /2/2{\mathbb{Z}}/2*{\mathbb{Z}}/2. Since Autt(X){\rm Aut}_{t}(X) is finite and Autt(X)Z(Aut(X)){\rm Aut}_{t}(X)\subset Z({\rm Aut}(X)), one has mX=1m_{X}=1.

Example 8.6.

Let X3X_{3} and X4X_{4} be the K3 surfaces of Picard rank 20 whose transcendental lattices are of the form

T(X3)=(2112) and T(X4)=(2002)T(X_{3})=\begin{pmatrix}2&1\\ 1&2\\ \end{pmatrix}\text{ and }T(X_{4})=\begin{pmatrix}2&0\\ 0&2\\ \end{pmatrix}

respectively, see [SI77] and [Huy16, Corollary 14.3.21]. By Vinberg ([Vin83, Theorem in 2.4 and Theorem in 3.3]), one has Autt(X)U0{\rm Aut}_{t}(X)\simeq U_{0} in his notation, hence mX3=3m_{X_{3}}=3 and mX4=2m_{X_{4}}=2.

Example 8.7.

Let XX be a K3 surface with an unimodular or 2-elementary transcendental lattice. Then there exists an involution σAut(X)\sigma\in{\rm Aut}(X) satisfying H2(σ)|NS(X)=idNS(X)H^{2}(\sigma)|_{{\rm NS}(X)}={\rm id}_{{\rm NS}(X)} and H2(σ)|T(X)=idT(X)H^{2}(\sigma)|_{T(X)}=-{\rm id}_{T(X)} , hence mXm_{X} is even.

Example 8.8.

Let XX be a K3 surface with φ(mX)=rkT(X)\varphi(m_{X})={\rm rk}~{}T(X), where φ\varphi is the Euler function. Then mXm_{X} is in the set

{12,28,36,42,44,66,3k(1k3),5l(l=1,2),7,11,13,17,19}\{12,28,36,42,44,66,3^{k}(1\leq k\leq 3),5^{l}(l=1,2),7,11,13,17,19\}

and XX is uniquely determined by mXm_{X} due to Kondo ([Kon92, Main Theorem]), Vorontsov ([Vor83]), Machida–Oguiso ([MO98, Theorem 3]) and Oguiso–Zhang ([OZ00, Theorem 2]). Especially, mXm_{X} is even (resp. odd) if and only if T(X)T(X) 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 Aut(𝒟(Ad)){\rm Aut}({\mathcal{D}}(A_{d})) and Aut(𝒟(Ad)c){\rm Aut}({\mathcal{D}}(A_{d})^{c}) for the dg-algebra of the graded dual numbers AdA_{d}.

Recall that, for a morphism of dg-algebras ϕ:BB\phi:B\rightarrow B^{\prime}, the functor Indϕ:𝒟(B)𝒟(B)\mathrm{Ind}_{\phi}\colon{\mathcal{D}}(B)\rightarrow{\mathcal{D}}(B^{\prime}) is defined as tensor product with BB seen as an B-BB\text{-}B^{\prime} bimodule. For λ×\lambda\in{\mathbb{C}}^{\times} let ϕλ:AdAd\phi_{\lambda}:A_{d}\rightarrow A_{d} be the morphism of dg-algebras defined by sending ε\varepsilon to λε\lambda\varepsilon.

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 𝒟(Ad)c{\mathcal{D}}(A_{d})^{c} is an autoequivalence, and for every ΦAut(𝒟(Ad))\Phi\in{\rm Aut}({\mathcal{D}}(A_{d})) there exists mm\in{\mathbb{Z}} and λ×\lambda\in{\mathbb{C}}^{\times} such that ΦIndϕλ[m]\Phi\simeq\mathrm{Ind}_{\phi_{\lambda}}\circ[m].

Proof.

If Φ:𝒟(Ad)c𝒟(Ad)c\Phi\colon{\mathcal{D}}(A_{d})^{c}\rightarrow{\mathcal{D}}(A_{d})^{c} is fully faithful, then it must induce an isomorphism of graded algebras Φ:AdHomAd(Φ(Ad),Φ(Ad))=nHom𝒟(Ad)(Φ(Ad),Φ(Ad)[n])[n]\Phi\colon A_{d}\rightarrow{\rm Hom}^{\bullet}_{A_{d}}(\Phi(A_{d}),\Phi(A_{d}))=\oplus_{n\in\mathbb{Z}}{\rm Hom}_{{\mathcal{D}}(A_{d})}(\Phi(A_{d}),\Phi(A_{d})[n])[-n]. As Φ(Ad)\Phi(A_{d}) is compact, by Proposition 6.7 we know that it decomposes as a direct sum of shifts of BnB_{n}’s for n>0n>0. It is easy to check that such a direct sum has endomorphism algebra equal to AdA_{d} if and only if there is just one summand and it is B1[m]B_{1}[m] for some mm\in{\mathbb{Z}}. Hence, Φ(Ad)Ad[m]\Phi(A_{d})\simeq A_{d}[m]. Now notice that the only maps of graded algebras AdAdA_{d}\rightarrow A_{d} (which coincide with the only dg-algebra maps from AdA_{d} to itself) are the ϕλ\phi_{\lambda}’s. Hence, Φ[m]Indϕλ1\Phi\circ[-m]\circ\mathrm{Ind}_{\phi_{\lambda}}^{-1} is an endofunctor of 𝒟(Ad)c{\mathcal{D}}(A_{d})^{c} that acts as the identity of AdA_{d}. Given Proposition 6.7, this implies that Φ[m]Indϕλ1id\Phi\circ[-m]\circ\mathrm{Ind}_{\phi_{\lambda}}^{-1}\simeq\mathrm{id}, and therefore ΦIndϕλ[m]\Phi\simeq\mathrm{Ind}_{\phi_{\lambda}}\circ[m], which is an autoequivalence, and the first claim follows.

For the second claim notice that if ΦAut(𝒟(Ad))\Phi\in{\rm Aut}({\mathcal{D}}(A_{d})) then it induces an autoequivalence of 𝒟(Ad)c{\mathcal{D}}(A_{d})^{c}, and we can repeat the argument above to conclude that ΦIndϕλ[m]\Phi\simeq\mathrm{Ind}_{\phi_{\lambda}}\circ[m] for some mm\in\mathbb{Z}. ∎

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.