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

Exceptional collections on Σ2\Sigma_{2}

Akira Ishii Graduate School of Mathematics, Nagoya University, Furocho, Chikusa-ku, Nagoya, 464-8602, Japan [email protected] Shinnosuke Okawa Department of Mathematics, Graduate School of Science, Osaka University
1-1, Machikaneyamacho, Toyonaka, Osaka 560-0043, Japan
[email protected]
 and  Hokuto Uehara Department of Mathematical Sciences, Graduate School of Science, Tokyo Metropolitan University, 1-1 Minamiohsawa, Hachioji-shi, Tokyo, 192-0397, Japan [email protected]
Abstract.

Structure theorems for exceptional objects and exceptional collections of the bounded derived category of coherent sheaves on del Pezzo surfaces are established by Kuleshov and Orlov in [KO94]. In this paper we propose conjectures which generalize these results to weak del Pezzo surfaces. Unlike del Pezzo surfaces, an exceptional object on a weak del Pezzo surface is not necessarily a shift of a sheaf and is not determined by its class in the Grothendieck group. Our conjectures explain how these complications are taken care of by spherical twists, the categorification of (2)(-2)-reflections acting on the derived category.

This paper is devoted to solving the conjectures for the prototypical weak del Pezzo surface Σ2\Sigma_{2}, the Hirzebruch surface of degree 22. Specifically, we prove the following results: Any exceptional object is sent to the shift of the uniquely determined exceptional vector bundle by a product of spherical twists which acts trivially on the Grothendieck group of the derived category. Any exceptional collection on Σ2\Sigma_{2} is part of a full exceptional collection. We moreover prove that the braid group on 4 strands acts transitively on the set of exceptional collections of length 44 (up to shifts).

1. Introduction

Semiorthogonal decomposition is among the most fundamental notions of triangulated categories. The finest ones, i.e., semiorthogonal decompositions whose components are equivalent to the bounded derived category of a point, are identified with (full) exceptional collections of the triangulated category. If there is an exceptional collection in a triangulated category, then virtually always there are infinitely many of them because of the group action explained below. Hence the classification of exceptional collections is not obvious at all.

Let XX be a smooth projective variety and 𝐃(X)\mathbf{D}(X) be the bounded derived category of coherent sheaves on XX. Concerning the classification of the exceptional collections of 𝐃(X)\mathbf{D}(X), the prototypical results are given in [GR87, Section 5] for X=2X=\mathbb{P}^{2}. The similar results for X=1×1X=\mathbb{P}^{1}\times\mathbb{P}^{1} are given in [Rud88]. Based on these works, Kuleshov and Orlov established in [KO94] the similar results for arbitrary del Pezzo surfaces (see also [GK04, Corollary 4.3.2, Theorem 4.3.3, Theorem 4.6.1]). They are summarized as follows.

Theorem 1.1 ([KO94]).

Let XX be a del Pezzo surface over an algebraically closed field 𝐤\mathbf{k}.

  1. (1)

    Any exceptional object 𝐃(X)\mathcal{E}\in\mathbf{D}(X) is either a vector bundle or a line bundle on a (1)(-1)-curve, up to shifts.

  2. (2)

    The isomorphism class of an exceptional object 𝐃(X)\mathcal{E}\in\mathbf{D}(X) is determined by its class []K0(X)[\mathcal{E}]\in\operatorname{K_{0}}\left(X\right), up to shifts by 22\mathbb{Z}.

  3. (3)

    Any exceptional collection of 𝐃(X)\mathbf{D}(X) can be extended to a full exceptional collection.

  4. (4)

    Conjecture 1.2 below is true for XX; i.e., the action Gr𝖤𝖢r(X)G_{r}\curvearrowright\mathsf{EC}_{r}(X) is transitive.

The symbol K0(X)\operatorname{K_{0}}\left(X\right) denotes the Grothendieck group of the triangulated category 𝐃(X)\mathbf{D}(X). We briefly explain (4). Let 𝖤𝖢r(X)\mathsf{EC}_{r}(X) denote the set of isomorphism classes of exceptional collections of length rr of 𝐃(X)\mathbf{D}(X), where r=rankK0(X)r=\operatorname{rank}\operatorname{K_{0}}\left(X\right). An important fact relevant to the classification is that there is a standard action of the group GrrBrrG_{r}\coloneqq\mathbb{Z}^{r}\rtimes\operatorname{Br}_{r} on 𝖤𝖢r(X)\mathsf{EC}_{r}(X) by shifts and mutations, where Brr\operatorname{Br}_{r} is the braid group on rr strands (see Section 2.3 for details).

Inspired by the earlier works mentioned above, Bondal and Polishchuk gave the following conjecture.

Conjecture 1.2 (==[BP93, Conjecture 2.2]).

Suppose that XX is a smooth projective variety such that 𝐃(X)\mathbf{D}(X) admits a full exceptional collection of length rr. Then the action Gr𝖤𝖢r(𝐃(X))G_{r}\curvearrowright\mathsf{EC}_{r}(\mathbf{D}(X)) is transitive.

In dimensions greater than 22, the classification of exceptional collections is widely open. Even for 3\mathbb{P}^{3}, only partial results seem to be known ([Pos95]) and Conjecture 1.2 is still open.

The aim of this paper is to investigate the generalization of Theorem 1.1 to weak del Pezzo surfaces; i.e., smooth projective surfaces XX whose anti-canonical bundle ωX1\omega_{X}^{-1} is nef and big. As it turns out, the generalization is not straightforward at all.

A weak del Pezzo surface XX which is not a del Pezzo surface admits at least one (and only finitely many) (2)(-2)-curve(s); i.e., a smooth curve CXC\subset X with C2=2C^{2}=-2 and C1C\simeq\mathbb{P}^{1}. Line bundles on CC, as objects of 𝐃(X)\mathbf{D}(X), are 22-spherical objects and hence yield non-trivial autoequivalences of 𝐃(X)\mathbf{D}(X) called spherical twists due to Seidel and Thomas [ST01]. For a 22-spherical object α𝐃(X)\alpha\in\mathbf{D}(X), the corresponding spherical twist TαT_{\alpha} acts as a reflection on K0(X)\operatorname{K_{0}}\left(X\right) whereas it always satisfies Tα2≄id𝐃(X)T_{\alpha}^{2}\not\simeq\operatorname{id}_{\mathbf{D}(X)}. For example, the exceptional object T𝒪C2(𝒪X)T_{\mathcal{O}_{C}}^{2}(\mathcal{O}_{X}) has the same class as 𝒪X\mathcal{O}_{X} in K0(X)\operatorname{K_{0}}\left(X\right) but is not isomorphic to 𝒪X\mathcal{O}_{X}. In fact, by direct computation one can confirm the following.

i(T𝒪C2(𝒪X)){𝒪Xi=0𝒪Ci=1,20otherwise\displaystyle\mathcal{H}^{i}\left(T_{\mathcal{O}_{C}}^{2}(\mathcal{O}_{X})\right)\simeq\begin{cases}\mathcal{O}_{X}&i=0\\ \mathcal{O}_{C}&i=1,2\\ 0&\text{otherwise}\end{cases} (1.1)

Moreover, by applying T𝒪C2T_{\mathcal{O}_{C}}^{2} repeatedly, we obtain a collection of infinitely many exceptional objects of unbounded cohomological amplitudes which share the same class in K0(X)\operatorname{K_{0}}\left(X\right). Hence Theorem 1.1 (1) (2) are not true for XX at all.

Conjecture 1.3, which is the main conjecture of this paper, generalizes Theorem 1.1 to weak del Pezzo surfaces while taking into account all the complications mentioned in the previous paragraph. In a word, it asserts that the failure of Theorem 1.1 (1) (2) on weak del Pezzo surfaces is remedied by spherical twists and that Theorem 1.1 (3) (4) should hold for weak del Pezzo surfaces too. The main theorem of this paper is that Conjecture 1.3 is true for Σ2=1(𝒪𝒪(2))\Sigma_{2}=\mathbb{P}_{\mathbb{P}^{1}}(\mathcal{O}\oplus\mathcal{O}(2)); namely, the Hirzebruch surface of degree 22, a paradigm of weak del Pezzo surfaces. In fact we formulated Conjecture 1.3 by generalizing the results we obtained for Σ2\Sigma_{2}, Theorem 1.1, and [CJ18, Theorem 1.2] simultaneously.

Conjecture 1.3.

Let XX be a weak del Pezzo surface over an algebraically closed field 𝐤\mathbf{k}.

  1. (1)

    For any exceptional object 𝐃(X)\mathcal{E}\in\mathbf{D}(X), there exists a sheaf \mathcal{F} on XX which is either an exceptional vector bundle or a line bundle on a (1)(-1)-curve, a sequence of line bundles on (2)(-2)-curves 1,,n\mathcal{L}_{1},\dots,\mathcal{L}_{n}, and an integer mm\in\mathbb{Z} such that

    (TnT1)()[m].\displaystyle\mathcal{E}\simeq(T_{\mathcal{L}_{n}}\circ\cdots\circ T_{\mathcal{L}_{1}})(\mathcal{F})[m]. (1.2)
  2. (2)

    For any pair of exceptional objects ,𝐃(X)\mathcal{E},\mathcal{E}^{\prime}\in\mathbf{D}(X) such that []=[]K0(X)[\mathcal{E}]=[\mathcal{E}^{\prime}]\in\operatorname{K_{0}}\left(X\right), there are a product bb of spherical twists and inverse spherical twists which acts trivially on K0(X)\operatorname{K_{0}}\left(X\right) and m2m\in 2\mathbb{Z} such that b()[m]\mathcal{E}^{\prime}\simeq b(\mathcal{E})[m]. Moreover, for each exceptional object 𝐃(X)\mathcal{E}\in\mathbf{D}(X), there is a unique exceptional vector bundle \mathcal{F} such that either []=[][\mathcal{E}]=[\mathcal{F}] or []=[][\mathcal{E}]=-[\mathcal{F}] holds.

  3. (3)

    Any exceptional collection on XX can be extended to a full exceptional collection.

  4. (4)

    Conjecture 1.2 is true for XX.

As it is more or less visible, (1), (2), (3), and (4) of Conjecture 1.3 generalizes (1), (2), (3), and (4) of Theorem 1.1, respectively.

To avoid repetition, let us simply state the main result of this paper as follows.

Theorem 1.4 (MAIN THEOREM).

Conjecture 1.3 is true for X=Σ2X=\Sigma_{2}.

More specifically, for X=Σ2X=\Sigma_{2}

  • Conjecture 1.3 (1) is solved affirmatively in Section 3 as Theorem 3.1.

  • Conjecture 1.3 (2) is solved affirmatively in Section 4 as Corollary 4.4.

  • Conjecture 1.3 (3) is solved affirmatively in Section 5 as Corollary 5.6.

  • Conjecture 1.3 (4) is solved affirmatively in Section 6 as Theorem 6.1.

Remark 1.5.

We give some comments on the preceding works which are related to Conjecture 1.3 and Theorem 1.4, and one important consequence of Conjecture 1.3 (4) on the fullness of exceptional collections of maximal length.

  • This paper is a continuation of [OU15] by the 2nd and the 3rd authors and almost completely supersedes it. Conjecture 1.3 (1) for Σ2\Sigma_{2} is stated as [OU15, Conjecture 1.3], and is solved for exceptional sheaves in [OU15, Theorem 1.4].

  • Conjecture 1.3 (1) for torsion exceptional sheaves is partially solved in [CJ18, Theorem 1.2]. A weaker version of Conjecture 1.3 (1) is stated as [CJ18, Conjecture 1.1].

  • Exceptional objects and exceptional collections of vector bundles on weak del Pezzo surfaces is systematically studied in [Kul97]. In fact we use some results of this work in this paper.

  • It follows from the definition of mutations that the triangulated subcategory generated by an exceptional collection is invariant under mutation. In particular, the fullness of an exceptional collection is preserved by mutations. On the other hand, any XX as in Conjecture 1.3 admits a full exceptional collection. Hence Conjecture 1.3 (4) would imply that any exceptional collection of length equal to rankK0(X)=rankPic(X)+2\operatorname{rank}\operatorname{K_{0}}\left(X\right)=\operatorname{rank}\operatorname{Pic}(X)+2 of 𝐃(X)\mathbf{D}(X) is full, though it would also follow from Conjecture 1.3 (3).

1.1. Summary of each section and structure of the paper

Section 2 is a preliminary section. We recall the rudiments of mutations and the group BB of autoequivalences generated by spherical twists from [IU05]. Among others, we prove in Corollary 2.11 that BK0𝗍𝗋𝗂𝗏B^{K_{0}-\mathsf{triv}}, the subgroup of BB acting trivially on K0(Σ2)\operatorname{K_{0}}\left(\Sigma_{2}\right), is generated by squares of spherical twists by line bundles on CC.

From Section 3 till the end of the paper, we restrict ourselves to the proof of Theorem 1.4 and in particular discuss the case of Σ2\Sigma_{2} only.

Section 3 is the main component of the paper and devoted to the proof of Theorem 3.1. Namely, in this section, we prove that for each exceptional object 𝐃(Σ2)\mathcal{E}\in\mathbf{D}(\Sigma_{2}) there is a sequence of integers a1,,ana_{1},\dots,a_{n} such that (TanTa1)()\left(T_{a_{n}}\circ\cdots\circ T_{a_{1}}\right)(\mathcal{E}) is isomorphic to a shift of a vector bundle. In accordance with the steps of the proof, Section 3 is divided into six subsections.

In Section 3.1 we prove some properties of the cohomology sheaf ()=ii()\mathcal{H}^{\bullet}(\mathcal{E})=\bigoplus_{i\in\mathbb{Z}}\mathcal{H}^{i}(\mathcal{E}). Among others we show that there is the index i0i_{0}\in\mathbb{Z} such that Suppi0()=Σ2\operatorname{Supp}\mathcal{H}^{i_{0}}(\mathcal{E})=\Sigma_{2} and for any ii0i\neq i_{0} the cohomology sheaf i()\mathcal{H}^{i}(\mathcal{E}), if not 0, is a pure sheaf whose reduced support is CC. Then in Section 3.2, we prove that actually the schematic support of i()\mathcal{H}^{i}(\mathcal{E}) for ii0i\neq i_{0} is CC. This is the most technical part of the paper, and actually this had been the main obstacle for the whole work. Fortunately one can use the result of this subsection as a black box to read the rest of the paper.

In Section 3.3 we prove that there is a decomposition i0()TE\mathcal{H}^{i_{0}}(\mathcal{E})\simeq T\oplus E, where E=E()E=E(\mathcal{E}) is an exceptional sheaf and TT is a torsion sheaf. We moreover show that if torsE0\operatorname{tors}E\neq 0, then TT is a direct sum of copies of 𝒪C(a)\mathcal{O}_{C}(a) for some aa\in\mathbb{Z} and that tors()\operatorname{tors}\mathcal{H}^{\bullet}(\mathcal{E}) is a direct sum of copies of 𝒪C(a)\mathcal{O}_{C}(a) and 𝒪C(a+1)\mathcal{O}_{C}(a+1). This integer a=a()a=a(\mathcal{E}) plays a central role throughout the paper.

In Section 3.4 we investigate the relationship between the cohomology sheaves of \mathcal{E} and those of \mathcal{E}^{\vee}, the derived dual of \mathcal{E}. We in particular show in Corollary 3.27 that if \mathcal{E} is not isomorphic to a shift of a vector bundle and torsE()=0\operatorname{tors}E(\mathcal{E})=0, then torsE()0\operatorname{tors}E(\mathcal{E}^{\vee})\neq 0. In this paper we mainly discuss the case torsE()0\operatorname{tors}E(\mathcal{E})\neq 0, and by this result we can settle the case where torsE()=0\operatorname{tors}E(\mathcal{E})=0 by passing to \mathcal{E}^{\vee}.

In Section 3.5 we introduce the notion of the length of the “torsion part” of an object at the generic point γ\gamma of CC. Formally speaking, for \mathcal{E} it is defined as ()=ilength𝒪Σ2,γtorsi()γ\ell(\mathcal{E})=\sum_{i\in\mathbb{Z}}\operatorname{length}_{\mathcal{O}_{\Sigma_{2},\gamma}}\operatorname{tors}\mathcal{H}^{i}(\mathcal{E})_{\gamma}. It follows that \mathcal{E} is isomorphic to a shift of a vector bundle if and only if ()=0\ell(\mathcal{E})=0, and hence it suffices to show that (Tc())<()\ell(T_{c}(\mathcal{E}))<\ell(\mathcal{E}) for some c=c()c=c(\mathcal{E})\in\mathbb{Z} if ()>0\ell(\mathcal{E})>0. This is exactly what we achieve in Section 3.6. We show in Theorem 3.32 that c=a()c=a(\mathcal{E}) works if torsE()0\operatorname{tors}E(\mathcal{E})\neq 0 and otherwise c=a()3c=-a(\mathcal{E}^{\vee})-3 does.

Spherical objects on the minimal resolution of type AA singularity is classified in [IU05]. More specifically, the proof of Theorem 3.1 is an adaptation of the proof of [IU05, Proposition 5.1]. It is, however, much more involved than that of [IU05, Proposition 5.1]. This is due to the fact that the support of an exceptional object on Σ2\Sigma_{2} is never concentrated in the (2)(-2)-curve CC. On the contrary the reduced support of a spherical object on Σ2\Sigma_{2} is concentrated in CC, and it immediately implies that the schematic supports of the cohomology sheaves of the spherical object coincide with CC.

Section 4 is devoted to the proof of Theorem 4.3. It is almost immediately obtained by combining Theorem 3.1 with a small trick on squares of spherical twists (Proposition 2.8) and the fact that an exceptional vector bundle is uniquely determined by its class in K0(Σ2)\operatorname{K_{0}}\left(\Sigma_{2}\right) (Lemma 2.33).

Section 5 is devoted to the proof of Corollary 5.6. Take an exceptional collection ¯\underline{\mathcal{E}}. We first show in Theorem 5.4 that there is a product of spherical twists bb such that b(¯)b(\underline{\mathcal{E}}) consists of vector bundles up to shifts. This is achieved in a one-by-one manner. The key is that if (,)(\mathcal{B},\mathcal{E}) is an exceptional pair such that \mathcal{B} is a vector bundle and ()>0\ell(\mathcal{E})>0, then, surprisingly enough, Tc()()T_{c(\mathcal{E})}(\mathcal{B}) remains to be a vector bundle (though it may not be isomorphic to \mathcal{B}). Recall that c()c(\mathcal{E})\in\mathbb{Z} depends only on \mathcal{E} and that Tc()T_{c(\mathcal{E})} strictly decreases the length of \mathcal{E}.

Corollary 5.6 is known for exceptional collections of vector bundles by a slight generalization Theorem 5.5 of a result by Kuleshov in [Kul97]. Applying it to b(¯)b(\underline{\mathcal{E}}), we immediately obtain the proof of Corollary 5.6 for ¯\underline{\mathcal{E}}.

Section 6 is devoted to the proof of Theorem 6.1. We use the deformation of Σ2\Sigma_{2} to 1×1\mathbb{P}^{1}\times\mathbb{P}^{1}. The difficulty is that there are infinitely many exceptional objects on Σ2\Sigma_{2} which deform to the same exceptional object on 1×1\mathbb{P}^{1}\times\mathbb{P}^{1} (non-uniqueness of the specialization, which is translated into the non-separatedness of the moduli space of semiorthogonal decompositions introduced in [BOR20]). Actually, two exceptional objects have the same deformation to 1×1\mathbb{P}^{1}\times\mathbb{P}^{1} if and only if they have the same class in K0(Σ2)\operatorname{K_{0}}\left(\Sigma_{2}\right) (up to shifts by 22\mathbb{Z}).

In Step 1 of the proof, we use the fact that the corresponding result is already known by Theorem 1.1 (4) for the del Pezzo surface 1×1\mathbb{P}^{1}\times\mathbb{P}^{1}. Since deformation of exceptional collections commutes with mutations, the result for 1×1\mathbb{P}^{1}\times\mathbb{P}^{1} immediately implies that for any exceptional collection of length 44 on Σ2\Sigma_{2} there is a sequence of mutations which brings it to a collection which is numerically equivalent to the standard collection 𝗌𝗍𝖽\mathcal{E}^{\mathsf{std}} (i.e., having the same classes as 𝗌𝗍𝖽\mathcal{E}^{\mathsf{std}} in K0(Σ2)\operatorname{K_{0}}\left(\Sigma_{2}\right). See (2.78) for the definition of 𝗌𝗍𝖽\mathcal{E}^{\mathsf{std}}).

It remains to show that an exceptional collection ¯\underline{\mathcal{E}} on Σ2\Sigma_{2} which is numerically equivalent to 𝗌𝗍𝖽\mathcal{E}^{\mathsf{std}} can be sent to 𝗌𝗍𝖽\mathcal{E}^{\mathsf{std}} by mutations. In Step 2, as an intermediate step, we find bBb\in B such that ¯=b(𝗌𝗍𝖽)\underline{\mathcal{E}}=b(\mathcal{E}^{\mathsf{std}}). We construct such bb again in one-by-one manner.

In Step 3 we prove that bb can be replaced by a sequence of mutations. Thanks to the fact that mutations commute with autoequivalences, it is enough to show the assertion only for b{T0,T1}b\in\{T_{0},T_{-1}\}. Recall that T0,T1T_{0},T_{-1} generate BB. At this point the problem is concrete enough to be settled by hand.

1.2. Some words on future directions

Though Conjecture 1.3 is stated for arbitrary weak del Pezzo surfaces, in this paper we restrict ourselves to the study of the case of Σ2\Sigma_{2}. This is partly because the proof is rather involved already in this case. We nevertheless think that the case of Σ2\Sigma_{2} should serve as a paradigm for the further investigations of Conjecture 1.3.

The proof of Theorem 1.4 is based on the classification of rigid sheaves on the (2)(-2)-curve; i.e., the fundamental cycle of the minimal resolution of the A1A_{1}-singularity. So far such classification is achieved only for the minimal resolution of type AA singularities by [IU05]. If one wants to push the strategy of this paper, it seems inevitable to establish the similar classification for the minimal resolution of type DD and type EE singularities. See [Kaw19] for results in this direction.

A weak del Pezzo surface over an algebraically closed field 𝐤\mathbf{k} is isomorphic to either 1×1,Σ2\mathbb{P}^{1}\times\mathbb{P}^{1},\Sigma_{2}, or a blowup of 2\mathbb{P}^{2} in at most eight points in almost general positions (see, say, [Dol12, Theorem 8.1.15, Corollary 8.1.24]). Hence our strategy based on the deformation to del Pezzo surfaces, in principle, is applicable to all weak del Pezzo surfaces.

Structure theorems for exceptional collections on 2\mathbb{P}^{2} play an important role in showing the contractibility of (the main component of) the space of Bridgeland stability conditions Stab(2)\operatorname{Stab}(\mathbb{P}^{2}) in [Li17]. Our results should be similarly useful for studying Stab(Σ2)\operatorname{Stab}(\Sigma_{2}). More specifically, they should be very closely related to the relationship between Stab(Σ2)\operatorname{Stab}(\Sigma_{2}) and Stab(1×1)\operatorname{Stab}(\mathbb{P}^{1}\times\mathbb{P}^{1}); see Remark 5.11 for details.

1.3. Notation and convention

We work over an algebraically closed field 𝐤\mathbf{k}, unless otherwise stated. To ease notation, we will write exti=dim𝐤Exti,hi=dim𝐤Hi,e2p,q=dim𝐤E2p,q,\operatorname{ext}^{i}=\dim_{\mathbf{k}}\operatorname{Ext}^{i},h^{i}=\dim_{\mathbf{k}}H^{i},e^{p,q}_{2}=\dim_{\mathbf{k}}E_{2}^{p,q}, and so on. Below is a list of frequently used symbols.

Σd\Sigma_{d} the Hirzebruch surface of degree dd (2.1)
CC the (2)(-2)-curve of Σ2\Sigma_{2}
ff the divisor class of a fiber of the morphism Σ21\Sigma_{2}\to\mathbb{P}^{1} (2.4)
\ast^{\vee} the derived dual of 𝐃(Σ2)\ast\in\mathbf{D}(\Sigma_{2}) (2.6)
Ta(resp. Ta)Auteq(Σ2)T_{a}\ (\text{resp. }T^{\prime}_{a})\in\operatorname{Auteq}(\Sigma_{2}) the (inverse) spherical twist by 𝒪C(a)\mathcal{O}_{C}(a) (2.15)
B<Auteq(Σ2)B<\operatorname{Auteq}(\Sigma_{2}) the group of autoequivalences generated by spherical twists (Definition 2.10)
𝖤𝖢N,𝖤𝖢𝖵𝖡N,𝖥𝖤𝖢,𝖥𝖤𝖢𝖵𝖡\mathsf{EC}_{N},\mathsf{ECVB}_{N},\mathsf{FEC},\mathsf{FECVB} various sets of exceptional collections (Definition 2.14)
BrN\operatorname{Br}_{N} (resp. GNG_{N}) the braid group on NN strands (2.55) (resp. the extension of BrN\operatorname{Br}_{N} by N\mathbb{Z}^{N} (2.60))
𝗀𝖾𝗇\operatorname{\mathsf{gen}} the generalization map for exceptional collections from the central fiber to the generic fiber (2.65)
𝗇𝗎𝗆𝖤𝖢N,𝗇𝗎𝗆𝖥𝖤𝖢\mathsf{numEC}_{N},\mathsf{numFEC} various sets of numerical exceptional collections (Definition 2.34)
𝗌𝗍𝖽𝖥𝖤𝖢𝖵𝖡(Σ2)\mathcal{E}^{\mathsf{std}}\in\mathsf{FECVB}(\Sigma_{2}) the standard full exceptional collection of 𝐃(Σ2)\mathbf{D}(\Sigma_{2}) (Definition 2.36)
ξ𝗌𝗍𝖽𝖥𝖤𝖢𝖵𝖡(𝒳𝗀𝖾𝗇)\mathcal{E}^{\mathsf{std}}_{\xi}\in\mathsf{FECVB}(\mathcal{X}_{\operatorname{\mathsf{gen}}}) the standard full exceptional collection of 𝐃(𝒳𝗀𝖾𝗇)\mathbf{D}(\mathcal{X}_{\operatorname{\mathsf{gen}}}) (2.79)
()(resp. i())\mathcal{H}^{\bullet}(\mathcal{E})\ (\text{resp. }\mathcal{H}^{i}(\mathcal{E})) the total (resp. ii-th) cohomology of 𝐃(X)\mathcal{E}\in\mathbf{D}(X) with respect to the standard t-structure (3.3)
i0=i0()i_{0}=i_{0}(\mathcal{E}) the unique index such that Suppi0()=Σ2\operatorname{Supp}\mathcal{H}^{i_{0}}(\mathcal{E})=\Sigma_{2} (Definition 3.7)
(R,𝔪)(R,\mathfrak{m}) the complete local ring of the A1A_{1}-singularity (Notation 3.10)
(0:I)MM(0:I)_{M}\subseteq M the maximal submodule of MM annihilated by the ideal IRI\subseteq R (3.28)
C𝒪Σ2\mathcal{I}_{C}\subset\mathcal{O}_{\Sigma_{2}} the ideal sheaf of CΣ2C\subset\Sigma_{2} (3.30)
D()D(\ast) the dual 𝐤\mathbf{k}-vector space of \ast (3.36)
i0()E()T()\mathcal{H}^{i_{0}}(\mathcal{E})\simeq E(\mathcal{E})\oplus T(\mathcal{E}) the canonical decomposition into an exceptional sheaf and a torsion sheaf (Lemma 3.21)
𝒯(),()\mathcal{T}(\mathcal{E}),\ \mathcal{F}(\mathcal{E}) the torsion (resp. the torsion free) part of i0()\mathcal{H}^{i_{0}}(\mathcal{E}) (Definition 3.22)
a,s,ta,s,t\in\mathbb{Z} integers specified by the irreducible decomposition of 𝒯\mathcal{T} (3.52) (see also Lemma 3.21)
()\ell(\ast) “length of the torsion part of \ast” at the generic point γ\gamma of CC (Definition 3.28)
b,r,sb,r,s\in\mathbb{Z} integers specified by the irreducible decomposition of an exceptional vector bundle restricted to CC (Lemma 3.20)

Acknowledgements

During the preparation of this paper, A.I. was partially supported by JSPS Grants-in-Aid for Scientific Research (19K03444). S.O. was partially supported by JSPS Grants-in-Aid for Scientific Research (16H05994, 16H02141, 16H06337, 18H01120, 20H01797, 20H01794). H.U. was partially supported by JSPS Grants-in-Aid for Scientific Research (18K03249).

2. Preliminaries

2.1. The Hirzebruch surfaces

The Hirzebruch surface of degree d0d\in\mathbb{Z}_{\geq 0} is the ruled surface

p:Σd1(𝒪1𝒪1(d))1.\displaystyle p\colon\Sigma_{d}\coloneqq\mathbb{P}_{\mathbb{P}^{1}}\left(\mathcal{O}_{\mathbb{P}^{1}}\oplus\mathcal{O}_{\mathbb{P}^{1}}(d)\right)\to\mathbb{P}^{1}. (2.1)

The history of Hirzebruch surfaces goes back, at least, to the first paper by Hirzebruch [Hir51]. An explicit isotrivial degeneration of Σd\Sigma_{d} to Σd\Sigma_{d^{\prime}} for d>dd>d^{\prime} and dd2d-d^{\prime}\in 2\mathbb{Z} is constructed in [Kod63, p.86 Example]. As a special case, there exists a smooth projective morphism (defined over Spec\operatorname{Spec}\mathbb{Z})

𝒳𝔸t1\displaystyle\mathcal{X}\to\mathbb{A}^{1}_{t} (2.2)

such that the fiber over t=0t=0 is isomorphic to Σ2\Sigma_{2} and the restriction of the family over the open subscheme 𝔾m𝔸1\mathbb{G}_{m}\hookrightarrow\mathbb{A}^{1} is isomorphic to the trivial family (1×1)×𝔾mpr2𝔾m\left(\mathbb{P}^{1}\times\mathbb{P}^{1}\right)\times\mathbb{G}_{m}\stackrel{{\scriptstyle\mathrm{pr}_{2}}}{{\to}}\mathbb{G}_{m}.

In this paper we investigate the bounded derived category of coherent sheaves on Σ2\Sigma_{2}, which is the most basic example of weak del Pezzo surfaces. We let CΣ2C\subset\Sigma_{2} denote the unique negative curve, and ff the (linear equivalence class of) the fiber of pp. Recall that

PicΣ2=Cf,\displaystyle\operatorname{Pic}\Sigma_{2}=\mathbb{Z}C\oplus\mathbb{Z}f, (2.3)

where

C2=2,f2=0,C.f=1.\displaystyle C^{2}=-2,f^{2}=0,C.f=1. (2.4)

The anti-canonical bundle is given by

KΣ2=2C+4f.\displaystyle-K_{\Sigma_{2}}=2C+4f. (2.5)

2.2. Derived category, spherical twist, and the autoequivalence group of Σ2\Sigma_{2}

Definition 2.1.

For a quasi-compact scheme YY, we let PerfY\operatorname{Perf}Y denote the perfect derived category of YY with the standard structure of a triangulated category. When YY is equipped with a morphism to Spec𝐤\operatorname{Spec}\mathbf{k}, we think of PerfY\operatorname{Perf}Y as a triangulated 𝐤\mathbf{k}-linear category. It comes with the natural symmetric monoidal structure given by the tensor product over 𝒪Y\mathcal{O}_{Y}, but we do not take it into account unless otherwise stated.

When YY is a smooth and projective variety over a field 𝐤\mathbf{k}, we identify PerfY\operatorname{Perf}Y with the bounded derived category 𝐃(Y)\mathbf{D}(Y) of coherent sheaves on YY.

The following equivalence of tensor triangulated categories

:(PerfY)opPerfY;omY(,𝒪Y)\displaystyle{}^{\vee}\colon\left(\operatorname{Perf}Y\right)^{\mathrm{op}}\xrightarrow{\sim}\operatorname{Perf}Y;\quad\mathcal{E}\mapsto\mathcal{E}^{\vee}\coloneqq\mathbb{R}\mathop{{\mathcal{H}}om}\nolimits_{Y}(\mathcal{E},\mathcal{O}_{Y}) (2.6)

will be called the derived dual.

One can easily verify that there exits a canonical natural isomorphism

id.\displaystyle\operatorname{id}\stackrel{{\scriptstyle\sim}}{{\Rightarrow}}{}^{\vee\vee}. (2.7)
Definition 2.2.

For smooth projective varieties X,YX,Y over Spec𝐤\operatorname{Spec}\mathbf{k} and an object K𝐃(X×𝐤Y)K\in\mathbf{D}(X\times_{\mathbf{k}}Y), the integral transform by the kernel KK will be denoted and defined as follows.

ΦKΦKXY:𝐃(X)𝐃(Y);EpY(pXEX×𝐤Y𝕃K)\displaystyle\Phi_{K}\coloneqq\Phi_{K}^{X\to Y}\colon\mathbf{D}(X)\to\mathbf{D}(Y);\quad E\mapsto\mathbb{R}p_{Y\ast}\left(p_{X}^{\ast}E\otimes_{X\times_{\mathbf{k}}Y}^{\mathbb{L}}K\right) (2.8)

Let XX be a smooth projective variety over a field 𝐤\mathbf{k}. Recall that an object α𝐃(X)\alpha\in\mathbf{D}(X) is spherical if α𝒪XωXα\alpha\otimes_{\mathcal{O}_{X}}\omega_{X}\simeq\alpha and HomX(α,α)𝐤𝐤[dimX]\mathop{\mathbb{R}\mathrm{Hom}}\nolimits_{X}(\alpha,\alpha)\simeq\mathbf{k}\oplus\mathbf{k}[-\dim X]. The spherical twist by α\alpha is the endofunctor

TαΦKα\displaystyle T_{\alpha}\coloneqq\Phi_{K_{\alpha}} (2.9)

defined by the kernel Kαcone(ααev𝒪ΔX)K_{\alpha}\coloneqq\operatorname{cone}\left(\alpha^{\vee}\boxtimes\alpha\xrightarrow{\operatorname{ev}}\mathcal{O}_{\Delta_{X}}\right).

Consider the exchange automorphism

𝗌𝗐𝖺𝗉:X×XX×X;(x,y)(y,x).\displaystyle\operatorname{\mathsf{swap}}\colon X\times X\to X\times X;\,(x,y)\mapsto(y,x). (2.10)

Recall that

(𝗌𝗐𝖺𝗉Kα)p1ωX[dimX]\displaystyle\left(\operatorname{\mathsf{swap}}^{*}K_{\alpha}\right)^{\vee}\otimes p_{1}^{\ast}\omega_{X}[\dim X] (2.11)

is called the right adjoint kernel of KαK_{\alpha} and it enjoys the following adjoint property (see, say, [Huy06, Definition 5.7]).

Tα=ΦKαΦ(𝗌𝗐𝖺𝗉Kα)p1ωX[dimX]Tα\displaystyle T_{\alpha}=\Phi_{K_{\alpha}}\dashv\Phi_{\left(\operatorname{\mathsf{swap}}^{*}K_{\alpha}\right)^{\vee}\otimes p_{1}^{\ast}\omega_{X}[\dim X]}\eqqcolon T^{\prime}_{\alpha} (2.12)

It follows that TαT_{\alpha} is an autoequivalence, so that TαTα1T^{\prime}_{\alpha}\simeq T_{\alpha}^{-1}. Note that there exists the obvious isomorphism

𝗌𝗐𝖺𝗉KαKα,\displaystyle\operatorname{\mathsf{swap}}^{\ast}K_{\alpha}\simeq K_{\alpha^{\vee}}, (2.13)

so that

TαΦKαp1ωX[dimX].\displaystyle T^{\prime}_{\alpha^{\vee}}\simeq\Phi_{K_{\alpha}^{\vee}\otimes p_{1}^{\ast}\omega_{X}[\dim X]}. (2.14)

A typical example of a spherical object on Σ2\Sigma_{2} is 𝒪C(a)\mathcal{O}_{C}(a) for aa\in\mathbb{Z}. The corresponding (inverse) spherical twist will be denoted as follows, for short.

TaT𝒪C(a),TaT𝒪C(a)1\displaystyle T_{a}\coloneqq T_{\mathcal{O}_{C}(a)},\quad T^{\prime}_{a}\coloneqq T_{\mathcal{O}_{C}(a)}^{-1} (2.15)

By definition, for each spherical object α\alpha and any object E𝐃(X)E\in\mathbf{D}(X), there are standard triangles as follows. To ease notation, we simply let 𝒪X\otimes_{\mathcal{O}_{X}} denote the derived tensor product. The morphisms ε\varepsilon and η\eta are the evaluation and the coevaluation maps respectively, both of which are obtained from the standard adjoint pair of functors 𝒪XHomX(,)-\otimes_{\mathcal{O}_{X}}\ast\dashv\mathop{\mathbb{R}\mathrm{Hom}}\nolimits_{X}(\ast,-).

HomX(α,E)𝐤α𝜀ETα(E)+1\displaystyle\mathop{\mathbb{R}\mathrm{Hom}}\nolimits_{X}\left(\alpha,E\right)\otimes_{\mathbf{k}}\alpha\xrightarrow{\varepsilon}E\to T_{\alpha}(E)\xrightarrow{+1} (2.16)
Tα(E)E𝜂HomX(E,α)𝐤α+1\displaystyle T^{\prime}_{\alpha}(E)\to E\xrightarrow{\eta}\mathop{\mathbb{R}\mathrm{Hom}}\nolimits_{X}\left(E,\alpha\right)^{\vee}\otimes_{\mathbf{k}}\alpha\xrightarrow{+1} (2.17)
Lemma 2.3 ([IU05, Lemma 4.14]).

For each ΦAuteq(X)\Phi\in\operatorname{Auteq}(X), there is an isomorphism of autoequivalences as follows.

ΦTαΦ1TΦ(α)\displaystyle\Phi\circ T_{\alpha}\circ\Phi^{-1}\simeq T_{\Phi(\alpha)} (2.18)

In particular, for any a,ma,m\in\mathbb{Z} it holds that

(𝒪Σ2(mC)𝒪Σ2)TaTa2m(𝒪Σ2(mC)𝒪Σ2).\displaystyle\left(\mathcal{O}_{\Sigma_{2}}(mC)\otimes_{\mathcal{O}_{\Sigma_{2}}}-\right)\circ T_{a}\simeq T_{a-2m}\circ\left(\mathcal{O}_{\Sigma_{2}}(mC)\otimes_{\mathcal{O}_{\Sigma_{2}}}-\right). (2.19)
Proof.

Since XX is a smooth projective variety over 𝐤\mathbf{k}, any autoequivalence Ψ\Psi of 𝐃(X)\mathbf{D}(X) is a Fourier-Mukai transform; i.e., it is isomorphic to the integral transform by an appropriate kernel by [Orl97, Theorem 2.2]. It then follows that Ψ\Psi is isomorphic to id𝐃(X)\operatorname{id}_{\mathbf{D}(X)} if and only if Ψ(E)E\Psi(E)\simeq E holds for any E𝐃(X)E\in\mathbf{D}(X). With this in mind, one can confirm the assertion by using (2.16). ∎

The operation of taking duals is related to (inverse) spherical twists nicely.

Lemma 2.4.
  1. (1)

    For any spherical object α𝐃(X)\alpha\in\mathbf{D}(X), there exists a natural isomorphism of functors

    (Tα)Tα().\displaystyle\left(T_{\alpha}-\right)^{\vee}\simeq T^{\prime}_{\alpha^{\vee}}\left(-^{\vee}\right). (2.20)
  2. (2)

    When X=Σ2X=\Sigma_{2}, for any aa\in\mathbb{Z}, there exists a natural isomorphism of functors

    (Ta)T2a().\displaystyle\left(T_{a}-\right)^{\vee}\simeq T^{\prime}_{-2-a}\left(-^{\vee}\right). (2.21)
Proof.

There is a natural isomorphism

(Tα)=(ΦKα())\displaystyle\left(T_{\alpha}-\right)^{\vee}=\left(\Phi_{K_{\alpha}}(-)\right)^{\vee} =omX(p2(p1X×kX𝕃Kα),𝒪X)\displaystyle=\mathop{\mathbb{R}\mathcal{H}om}\nolimits_{X}\left(\mathbb{R}p_{2\ast}(p_{1}^{\ast}-\otimes^{\mathbb{L}}_{X\times_{k}X}K_{\alpha}),\mathcal{O}_{X}\right) (2.22)
omX(p2(ωXX×kX𝕃Kα[dimX]),ωX[dimX])\displaystyle\simeq\mathop{\mathbb{R}\mathcal{H}om}\nolimits_{X}\left(\mathbb{R}p_{2\ast}\left(-\boxtimes\omega_{X}\otimes^{\mathbb{L}}_{X\times_{k}X}K_{\alpha}[\dim X]\right),\omega_{X}[\dim X]\right) (2.23)
p2omX×𝐤X(ωXX×kX𝕃Kα[dimX],ωXωX[2dimX])\displaystyle\stackrel{{\scriptstyle\ast}}{{\simeq}}\mathbb{R}p_{2\ast}\mathop{\mathbb{R}\mathcal{H}om}\nolimits_{X\times_{\mathbf{k}}X}\left(-\boxtimes\omega_{X}\otimes^{\mathbb{L}}_{X\times_{k}X}K_{\alpha}[\dim X],\omega_{X}\boxtimes\omega_{X}[2\dim X]\right) (2.24)
p2(p1X×𝐤X(Kαp1ωX[dimX]))\displaystyle\simeq\mathbb{R}p_{2\ast}\left(p_{1}^{\ast}-^{\vee}\otimes_{X\times_{\mathbf{k}}X}\left(K_{\alpha}^{\vee}\otimes p_{1}^{\ast}\omega_{X}[\dim X]\right)\right) (2.25)
ΦKαp1ωΣ2[dimX](),\displaystyle\simeq\Phi_{K_{\alpha}^{\vee}\otimes p_{1}^{\ast}\omega_{\Sigma_{2}}[\dim X]}(-^{\vee}), (2.26)

where the isomorphism \stackrel{{\scriptstyle\ast}}{{\simeq}} follows from [Har66, Chapter VII, Corollary 4.3 b)].

Comparing this with (2.14), we obtain (2.20).

The second item follows from (2.20) and the isomorphism 𝒪C(a)[1]𝒪C(2a)\mathcal{O}_{C}(a)^{\vee}[1]\simeq\mathcal{O}_{C}(-2-a). ∎

Lemma 2.5 ([IU05, Lemma 4.15 (i) (2)]).

There is an isomorphism of autoequivalences

TaTa+1𝒪Σ2(C)𝒪Σ2,\displaystyle T_{a}\circ T_{a+1}\simeq\mathcal{O}_{\Sigma_{2}}(C)\otimes_{\mathcal{O}_{\Sigma_{2}}}-, (2.27)

which implies

TaTa+1𝒪Σ2(C)𝒪Σ2,Ta2Ta+1Ta+3𝒪Σ2(2C)𝒪Σ2.\displaystyle\begin{aligned} T_{a}^{\prime}\simeq T_{a+1}\circ\mathcal{O}_{\Sigma_{2}}(-C)\otimes_{\mathcal{O}_{\Sigma_{2}}}-,\\ {T_{a}^{\prime}}^{2}\simeq T_{a+1}\circ T_{a+3}\circ\mathcal{O}_{\Sigma_{2}}(-2C)\otimes_{\mathcal{O}_{\Sigma_{2}}}-.\end{aligned} (2.28)
Definition 2.6.

For a triangulated category 𝐃\mathbf{D}, we define the group AuteqK0𝗍𝗋𝗂𝗏(𝐃)\operatorname{Auteq}^{K_{0}-\mathsf{triv}}(\mathbf{D}) of K0K_{0}-trivial autoequivalences by the following exact sequence.

1AuteqK0𝗍𝗋𝗂𝗏(𝐃)Auteq(𝐃)Aut(K0(𝐃))\displaystyle 1\to\operatorname{Auteq}^{K_{0}-\mathsf{triv}}(\mathbf{D})\to\operatorname{Auteq}(\mathbf{D})\to\operatorname{Aut}\left(\operatorname{K_{0}}\left(\mathbf{D}\right)\right) (2.29)

The following well-known lemma, which follows from (2.16) and (2.17), asserts that spherical twists are categorification of root reflections.

Lemma 2.7.

For any spherical object α𝐃(Σ2)\alpha\in\mathbf{D}(\Sigma_{2}), it holds that

Tα2,Tα2AuteqK0𝗍𝗋𝗂𝗏(𝐃(Σ2)).\displaystyle T_{\alpha}^{2},{T^{\prime}_{\alpha}}^{2}\in\operatorname{Auteq}^{K_{0}-\mathsf{triv}}(\mathbf{D}(\Sigma_{2})). (2.30)
Proposition 2.8.

For any a,ba,b\in\mathbb{Z}, there exists a sequence a1,,aa_{1},\dots,a_{\ell}\in\mathbb{Z} and mm\in\mathbb{Z} such that

TaTb=𝒪Σ2(mC)𝒪Σ2Ta±2Ta1±2.\displaystyle T_{a}T_{b}=\mathcal{O}_{\Sigma_{2}}(mC)\otimes_{\mathcal{O}_{\Sigma_{2}}}-\circ T_{a_{\ell}}^{\pm 2}\cdots T_{a_{1}}^{\pm 2}. (2.31)
Proof.

By (2.19), we may and will reduce the proof to that of the following claim.

Claim 2.9.

TaTbT_{a}T_{b} is contained in the subgroup of Auteq(𝐃(Σ2))\operatorname{Auteq}(\mathbf{D}(\Sigma_{2})) generated by 𝒪Σ2(C)𝒪Σ2\mathcal{O}_{\Sigma_{2}}(C)\otimes_{\mathcal{O}_{\Sigma_{2}}}- and Ta2T_{a}^{2} for all aa\in\mathbb{Z}.

In the rest, we prove this claim. Consider first the case aba\leq b. We induct on ba0b-a\geq 0. If it is 0, we have nothing to show. In general, we have

TaTb(TaTa+1)(Ta+12)(Ta+1Tb).\displaystyle T_{a}T_{b}\simeq\left(T_{a}T_{a+1}\right)\left({T^{\prime}_{a+1}}^{2}\right)\left(T_{a+1}T_{b}\right). (2.32)

By (2.27), the first term of the right hand side is isomorphic to 𝒪Σ2(C)𝒪Σ2\mathcal{O}_{\Sigma_{2}}(C)\otimes_{\mathcal{O}_{\Sigma_{2}}}-. By the induction hypothesis, the 3rd term is also contained in the subgroup mentioned in the claim. Hence we are done with this case.

Now suppose that a>ba>b. Then

TaTbTa2(TbTa)Tb2,\displaystyle T_{a}T_{b}\simeq T_{a}^{2}\left(T_{b}T_{a}\right)^{\prime}T_{b}^{2}, (2.33)

and we know that the middle term is contained in the subgroup as shown in the previous paragraph. ∎

Definition 2.10.

The subgroup of Auteq(𝐃(Σ2))\operatorname{Auteq}(\mathbf{D}(\Sigma_{2})) generated by spherical twists will be denoted by BB.

Corollary 2.11.

Fix any a0a_{0}\in\mathbb{Z}. Then elements of BB are exhausted by those of the following form, where a1,,an,ma_{1},\dots,a_{n},m\in\mathbb{Z}.

(𝒪Σ2(mC)𝒪Σ2)(Tan±2Ta1±2)\displaystyle\left(\mathcal{O}_{\Sigma_{2}}(mC)\otimes_{\mathcal{O}_{\Sigma_{2}}}-\right)\circ\left(T_{a_{n}}^{\pm 2}\circ\cdots\circ T_{a_{1}}^{\pm 2}\right) (2.34)
(𝒪Σ2(mC)𝒪Σ2)Ta0(Tan±2Ta1±2)\displaystyle\left(\mathcal{O}_{\Sigma_{2}}(mC)\otimes_{\mathcal{O}_{\Sigma_{2}}}-\right)\circ T_{a_{0}}\circ\left(T_{a_{n}}^{\pm 2}\circ\cdots\circ T_{a_{1}}^{\pm 2}\right) (2.35)

Moreover, the normal subgroup

BK0𝗍𝗋𝗂𝗏BAuteqK0𝗍𝗋𝗂𝗏(𝐃(X))B\displaystyle B^{K_{0}-\mathsf{triv}}\coloneqq B\cap\operatorname{Auteq}^{K_{0}-\mathsf{triv}}(\mathbf{D}(X))\triangleleft B (2.36)

is generated by {Ta2a}BK0𝗍𝗋𝗂𝗏\{T_{a}^{2}\mid a\in\mathbb{Z}\}\subset B^{K_{0}-\mathsf{triv}}.

Proof.

The first assertion immediately follows from (2.28), (2.19), and Proposition 2.8.

For the second assertion, take bBK0𝗍𝗋𝗂𝗏b\in B^{K_{0}-\mathsf{triv}}. If bb is as in (2.34), then obviously m=0m=0. Also one can verify that bb is never like (2.35), say, by using that it must preserve the classes [𝒪Σ2((a0+1)f)][\mathcal{O}_{\Sigma_{2}}((a_{0}+1)f)] and [𝒪Σ2((a0+2)f)][\mathcal{O}_{\Sigma_{2}}((a_{0}+2)f)]. ∎

Lemma 2.3 implies that BB is a normal subgroup of Auteq(𝐃(Σ2))\operatorname{Auteq}(\mathbf{D}(\Sigma_{2})). As we explain next, sets of generators of BB are well understood.

Theorem 2.12.

For any aa\in\mathbb{Z}, the group BB is generated by the two spherical twists Ta,Ta+1T_{a},T_{a+1}.

Proof.

The case a=2a=-2 is explicitly mentioned in [IU05, Lemma 4.6]. The general case inductively follows from this by the isomorphism of functors

Ta1Ta𝒪Σ2(C)𝒪Σ2TaTa+1,\displaystyle T_{a-1}T_{a}\simeq\mathcal{O}_{\Sigma_{2}}(C)\otimes_{\mathcal{O}_{\Sigma_{2}}}-\simeq T_{a}T_{a+1}, (2.37)

which is (2.27). ∎

2.3. Deformation and mutation of exceptional collections

Let

f:𝒳B\displaystyle f\colon\mathcal{X}\to B (2.38)

be a smooth projective morphism of Noetherian schemes with a closed point 0B0\in B, and let

X0=𝒳×f,B{0}\displaystyle X_{0}=\mathcal{X}\times_{f,B}\{0\} (2.39)

be the central fiber. Note that the properness and smoothness of ff implies it is a perfect morphism, in the sense that the derived pushforward ff_{\ast} respects perfect complexes ([LN07, Proposition 2.1]).

Definition 2.13.
  1. (1)

    An object Perf(𝒳)\mathcal{E}\in\operatorname{Perf}(\mathcal{X}) is ff-exceptional if fom(,)\mathbb{R}f_{\ast}\mathbb{R}\mathop{{\mathcal{H}}om}\nolimits(\mathcal{E},\mathcal{E}) is a line bundle on BB.

  2. (2)

    A collection of ff-exceptional objects 1,,NPerf𝒳\mathcal{E}_{1},\dots,\mathcal{E}_{N}\in\operatorname{Perf}\mathcal{X} is an ff-exceptional collection if fom(j,i)=0\mathbb{R}f_{\ast}\mathbb{R}\mathop{{\mathcal{H}}om}\nolimits(\mathcal{E}_{j},\mathcal{E}_{i})=0 for any (i,j)(i,j) with 1i<jN1\leq i<j\leq N.

  3. (3)

    An ff-exceptional collection as above is said to be strong if moreover fom(j,i)\mathbb{R}f_{\ast}\mathbb{R}\mathop{{\mathcal{H}}om}\nolimits(\mathcal{E}_{j},\mathcal{E}_{i}) is isomorphic to a locally free sheaf (regarded as a complex concentrated in degree 0) for any (i,j)(i,j).

  4. (4)

    An ff-exceptional collection as above is said to be full if the minimal BB-linear (i.e., closed under f-\otimes f^{\ast}\mathcal{F} by any PerfB\mathcal{F}\in\operatorname{Perf}B) triangulated subcategory which contains all of the objects in the collection is equivalent to Perf𝒳\operatorname{Perf}\mathcal{X}.

Definition 2.14.

We say that two ff-exceptional collections 1,,N\mathcal{E}_{1},\dots,\mathcal{E}_{N} and 1,,N\mathcal{E}^{\prime}_{1},\dots,\mathcal{E}^{\prime}_{N} are isomorphic if ii\mathcal{E}_{i}\simeq\mathcal{E}^{\prime}_{i} for each i=1,,Ni=1,\dots,N. We will let 𝖤𝖢N(f)\mathsf{EC}_{N}(f) (or 𝖤𝖢N(𝒳)\mathsf{EC}_{N}(\mathcal{X}), if ff is obvious from the context) denote the set of isomorphism classes of ff-exceptional collections of length NN. The set of isomorphism classes of ff-exceptional collections of length NN consisting entirely of locally free sheaves will be denoted by 𝖤𝖢𝖵𝖡N(f)\mathsf{ECVB}_{N}(f) (or 𝖤𝖢𝖵𝖡(𝒳)\mathsf{ECVB}(\mathcal{X})), which comes with the obvious injection 𝖤𝖢𝖵𝖡N(f)𝖤𝖢N(f)\mathsf{ECVB}_{N}(f)\hookrightarrow\mathsf{EC}_{N}(f). Similarly, the set of equivalence classes of full ff-exceptional collections will be denoted by either 𝖥𝖤𝖢(f)\mathsf{FEC}(f) or 𝖥𝖤𝖢(𝒳)\mathsf{FEC}(\mathcal{X}), and the set of isomorphism classes of full ff-exceptional collections consisting entirely of locally free sheaves will be denoted by 𝖥𝖤𝖢𝖵𝖡(f)\mathsf{FECVB}(f) (or 𝖥𝖤𝖢𝖵𝖡(𝒳)\mathsf{FECVB}(\mathcal{X})), which comes with the obvious injection 𝖥𝖤𝖢𝖵𝖡(f)𝖥𝖤𝖢(f)\mathsf{FECVB}(f)\hookrightarrow\mathsf{FEC}(f).

Lemma 2.15.
  1. (1)

    Let Perf𝒳\mathcal{E}\in\operatorname{Perf}\mathcal{X} be an ff-exceptional object. Then the functor

    Φ:PerfBPerf𝒳;FfF\displaystyle\Phi_{\mathcal{E}}\colon\operatorname{Perf}B\to\operatorname{Perf}\mathcal{X};\quad F\mapsto f^{\ast}F\otimes\mathcal{E} (2.40)

    is fully faithful and admits a right adjoint as follows.

    ΦϕRf()\displaystyle\Phi_{\mathcal{E}}\dashv\phi_{\mathcal{E}}^{R}\coloneqq f_{\ast}\left(-\otimes\mathcal{E}^{\vee}\right) (2.41)
  2. (2)

    Let (1,,N)𝖤𝖢N(𝒳)\left(\mathcal{E}_{1},\dots,\mathcal{E}_{N}\right)\in\mathsf{EC}_{N}(\mathcal{X}) be an ff-exceptional collection. Then the smallest BB-linear triangulated subcategory of Perf𝒳\operatorname{Perf}\mathcal{X} which contains 1,,N\mathcal{E}_{1},\dots,\mathcal{E}_{N} admits a BB-linear semiorthogonal decomposition Φ1(PerfB),ΦN(PerfB)\langle\Phi_{\mathcal{E}_{1}}(\operatorname{Perf}B),\dots\Phi_{\mathcal{E}_{N}}(\operatorname{Perf}B)\rangle.

We say that a semiorthogonal decomposition Perf𝒳=𝒜1,,𝒜N\operatorname{Perf}\mathcal{X}=\langle\mathcal{A}_{1},\dots,\mathcal{A}_{N}\rangle is BB-linear if 𝒜ifb𝒜i\mathcal{A}_{i}\otimes f^{\ast}b\subseteq\mathcal{A}_{i} holds for any i=1,,Ni=1,\dots,N and bPerfBb\in\operatorname{Perf}B (see [Kuz11, Section 2.3]).

We will freely use the following very useful base change theorem from [Bon06, Corollary 2.1.4]. See also [Kuz06, Section 2.4] and [Sta16, Tag 08IB] for treatise from different points of view.

Lemma 2.16.

Consider the following Cartesian square of finite dimensional noetherian schemes, where f,gf,g are perfect.

Y{Y}𝒳{\mathcal{X}}C{C}B{B}g\scriptstyle{g}ψ\scriptstyle{\psi}f\scriptstyle{f}φ\scriptstyle{\varphi} (2.42)

Then the standard natural transformation of functors

φfgψ:Perf𝒳PerfC\displaystyle\varphi^{\ast}\circ f_{\ast}\Rightarrow g_{\ast}\circ\psi^{\ast}\colon\operatorname{Perf}\mathcal{X}\to\operatorname{Perf}C (2.43)

is an isomorphism if either ff or φ\varphi is flat.

Corollary 2.17.

Consider a morphism of schemes as in (2.38) and a (strong) ff-exceptional collection (1,,N)𝖤𝖢N(𝒳)\left(\mathcal{E}_{1},\dots,\mathcal{E}_{N}\right)\in\mathsf{EC}_{N}(\mathcal{X}). Take any morphism φ:CB\varphi\colon C\to B, where CC is a finite dimensional noetherian scheme, and consider the Cartesian diagram as in Lemma 2.16. Then

(ψ1,,ψN)\displaystyle\left(\psi^{\ast}\mathcal{E}_{1},\dots,\psi^{\ast}\mathcal{E}_{N}\right) (2.44)

is a (strong) gg-exceptional collection on YY.

Proof.

As ff is flat, the assertion immediately follows from the following computation. All functors are derived.

gomY(ψi,ψj)gψom𝒳(i,j)Lemma 2.16φfom𝒳(i,j)\displaystyle g_{\ast}\mathop{{\mathcal{H}}om}\nolimits_{Y}\left(\psi^{\ast}\mathcal{E}_{i},\psi^{\ast}\mathcal{E}_{j}\right)\simeq g_{\ast}\psi^{\ast}\mathop{{\mathcal{H}}om}\nolimits_{\mathcal{X}}\left(\mathcal{E}_{i},\mathcal{E}_{j}\right)\stackrel{{\scriptstyle\text{Lemma~{}\ref{lm:base change}}}}{{\simeq}}\varphi^{\ast}f_{\ast}\mathop{{\mathcal{H}}om}\nolimits_{\mathcal{X}}\left(\mathcal{E}_{i},\mathcal{E}_{j}\right) (2.45)

Lemma 2.18.

Suppose that B=SpecRB=\operatorname{Spec}R for a complete local Noetherian ring (R,𝔪,𝐤)(R,\mathfrak{m},\mathbf{k}). Then the natural restriction maps

𝖤𝖢N(f)𝖤𝖢N(X0),\displaystyle\mathsf{EC}_{N}(f)\to\mathsf{EC}_{N}(X_{0}), (2.46)
𝖤𝖢𝖵𝖡N(f)𝖤𝖢𝖵𝖡N(X0)\displaystyle\mathsf{ECVB}_{N}(f)\to\mathsf{ECVB}_{N}(X_{0}) (2.47)

obtained in Corollary 2.17 are bijections for any NN.

Proof.

Let us first show that any exceptional object PerfX0\mathcal{E}\in\operatorname{Perf}X_{0} deforms to an object RPerf𝒳\mathcal{E}_{R}\in\operatorname{Perf}\mathcal{X}. By definition of exceptional object, we know that ExtX01(,)=ExtX02(,)=0\operatorname{Ext}^{1}_{X_{0}}(\mathcal{E},\mathcal{E})=\operatorname{Ext}^{2}_{X_{0}}(\mathcal{E},\mathcal{E})=0. By the deformation theory of objects (see, say, [HT10, Corollary 3.4]), for each n1n\geq 1 one finds the unique lift in Perf(𝒳RR/𝔪n+1)\operatorname{Perf}(\mathcal{X}\otimes_{R}R/\mathfrak{m}^{n+1}) of EE. Then it algebrizes uniquely to an actual object RPerf(𝒳)\mathcal{E}_{R}\in\operatorname{Perf}(\mathcal{X}) by [Lie06, Proposition 3.6.1].

We next show that R\mathcal{E}_{R} is an ff-exceptional object. This is equivalent to the assertion 𝒞=0\mathcal{C}=0, where 𝒞PerfB\mathcal{C}\in\operatorname{Perf}B is defined as the cone of the following standard morphism.

𝒪Bfom𝒳(R,R)\displaystyle\mathcal{O}_{B}\to\mathbb{R}f_{\ast}\mathbb{R}\mathop{{\mathcal{H}}om}\nolimits_{\mathcal{X}}(\mathcal{E}_{R},\mathcal{E}_{R}) (2.48)

Consider the following Cartesian diagram.

X0{X_{0}}𝒳{\mathcal{X}}Spec𝐤{\operatorname{Spec}\mathbf{k}}SpecR=B{\operatorname{Spec}R=B}ι\scriptstyle{\iota}f0\scriptstyle{f_{0}}{\lrcorner}f\scriptstyle{f}0\scriptstyle{0} (2.49)

By Lemma 2.16 it follows that

𝕃0fom𝒳(R,R)HomX0(,)=𝐤id[0],\displaystyle\mathbb{L}0^{\ast}\mathbb{R}f_{\ast}\mathop{\mathbb{R}\mathcal{H}om}\nolimits_{\mathcal{X}}(\mathcal{E}_{R},\mathcal{E}_{R})\simeq\mathop{\mathbb{R}\mathrm{Hom}}\nolimits_{X_{0}}(\mathcal{E},\mathcal{E})=\mathbf{k}\operatorname{id}_{\mathcal{E}}[0], (2.50)

so that 𝕃0𝒞=0\mathbb{L}0^{\ast}\mathcal{C}=0. By Nakayama’s lemma, this implies that 𝒞=0\mathcal{C}=0. The semiorthogonality of the collection 1,R,,N,R\mathcal{E}_{1,R},\dots,\mathcal{E}_{N,R} is shown by similar arguments. ∎

Remark 2.19.

One can similarly show that the deformation of a strong collection is also strong. This follows from the fact that if PPerfBP\in\operatorname{Perf}B satisfies 0P𝐤r[0]0^{\ast}P\simeq\mathbf{k}^{\oplus r}[0] for some r0r\geq 0, then PRr[0]P\simeq R^{\oplus r}[0]. Also, it follows from Lemma 2.15 (2) that the deformation of a full exceptional collection is again full.

Definition 2.20.

Let ff be a morphism as in (2.38). An (ff-)exceptional pair is an (ff-)exceptional collection of length 2. For an ff-exceptional pair ,\mathcal{E},\mathcal{F}, the left mutation LL_{\mathcal{E}}\mathcal{F} of \mathcal{F} through \mathcal{E} and the right mutation RR_{\mathcal{F}}\mathcal{E} of \mathcal{E} through \mathcal{F} are defined by the following distinguished triangles.

ffom𝒳(,)𝒪𝒳𝜀L,\displaystyle f^{\ast}\mathbb{R}f_{\ast}\mathop{\mathbb{R}\mathcal{H}om}\nolimits_{\mathcal{X}}(\mathcal{E},\mathcal{F})\otimes_{\mathcal{O}_{\mathcal{X}}}\mathcal{E}\xrightarrow{\varepsilon}\mathcal{F}\to L_{\mathcal{E}}\mathcal{F}, (2.51)
R𝜂ffom𝒳(,)𝒪𝒳\displaystyle R_{\mathcal{F}}\mathcal{E}\to\mathcal{E}\xrightarrow{\eta}f^{\ast}\mathbb{R}f_{\ast}\mathop{\mathbb{R}\mathcal{H}om}\nolimits_{\mathcal{X}}(\mathcal{E},\mathcal{F})^{\vee}\otimes_{\mathcal{O}_{\mathcal{X}}}\mathcal{F} (2.52)
Remark 2.21.

The definition of mutations given above differs by shifts from the one in [Bon89, Section 2], but is slightly simpler in that for an orthogonal exceptional pair, the mutations just exchange the two objects without any shift.

By the base change theorem Lemma 2.16, one can easily verify that mutations commute with base change.

Lemma 2.22.

Under the notation and the assumptions of Lemma 2.16, suppose that ff is flat and hence the natural transformation (2.43) is an isomorphism. For any ff-exceptional pair (,)\left(\mathcal{E},\mathcal{F}\right), it follows that (ψ,ψ)\left(\psi^{\ast}\mathcal{E},\psi^{\ast}\mathcal{F}\right) is an gg-exceptional pair and the following isomorphisms hold.

Lψ(ψ)ψ(L)\displaystyle L_{\psi^{\ast}\mathcal{E}}\left(\psi^{\ast}\mathcal{F}\right)\simeq\psi^{\ast}(L_{\mathcal{E}}\mathcal{F}) (2.53)
Rψ(ψ)ψ(R)\displaystyle R_{\psi^{\ast}\mathcal{F}}\left(\psi^{\ast}\mathcal{E}\right)\simeq\psi^{\ast}(R_{\mathcal{F}}\mathcal{E}) (2.54)

Next, let us recall a group action on 𝖤𝖢N(f)\mathsf{EC}_{N}(f) from [BP93, Proposition 2.1]. Let BrN\operatorname{Br}_{N} be the braid group on NN strands, which admits the following famous presentation by generators and relations.

BrN=σ1,,σN1\displaystyle\operatorname{Br}_{N}=\langle\sigma_{1},\dots,\sigma_{N-1}\mid\, σiσi+1σi=σi+1σiσi+1i=1,,N1\displaystyle\sigma_{i}\sigma_{i+1}\sigma_{i}=\sigma_{i+1}\sigma_{i}\sigma_{i+1}\quad i=1,\dots,N-1 (2.55)
σiσj=σjσi|ij|2\displaystyle\sigma_{i}\sigma_{j}=\sigma_{j}\sigma_{i}\quad|i-j|\geq 2\rangle (2.56)

Consider the action

BrN𝖤𝖢N(f)\displaystyle\operatorname{Br}_{N}\curvearrowright\mathsf{EC}_{N}(f) (2.57)

given by

σi:i,i+1i+1,Ri+1i,\displaystyle\sigma_{i}\colon\mathcal{E}_{i},\mathcal{E}_{i+1}\mapsto\mathcal{E}_{i+1},R_{\mathcal{E}_{i+1}}\mathcal{E}_{i}, (2.58)

so that

σi1:i,i+1Lii+1,i.\displaystyle\sigma_{i}^{-1}\colon\mathcal{E}_{i},\mathcal{E}_{i+1}\mapsto L_{\mathcal{E}_{i}}\mathcal{E}_{i+1},\mathcal{E}_{i}. (2.59)

On the other hand, through the standard surjective homomorphism BrN𝔖N;σi(i,i+1)\operatorname{Br}_{N}\twoheadrightarrow\mathfrak{S}_{N};\quad\sigma_{i}\mapsto(i,i+1) to the symmetric group of degree NN, the group BrN\operatorname{Br}_{N} acts naturally on the abelian group N=𝖬𝖺𝗉({1,,N},)\mathbb{Z}^{N}=\operatorname{\mathsf{Map}}(\left\{1,\dots,N\right\},\mathbb{Z}) from the left. Let

GNNBrN\displaystyle G_{N}\coloneqq\mathbb{Z}^{N}\rtimes\operatorname{Br}_{N} (2.60)

be the semi-direct product corresponding to the action.

One can verify that this, together with the action N𝖤𝖢N\mathbb{Z}^{N}\curvearrowright\mathsf{EC}_{N}, where [a1,,aN]TN[a_{1},\dots,a_{N}]^{T}\in\mathbb{Z}^{N} sends a collection (1,,N)𝖤𝖢N\left(\mathcal{E}_{1},\dots,\mathcal{E}_{N}\right)\in\mathsf{EC}_{N} to (1[a1],,N[aN])\left(\mathcal{E}_{1}[a_{1}],\dots,\mathcal{E}_{N}[a_{N}]\right), extends to an action GN𝖤𝖢N(f)G_{N}\curvearrowright\mathsf{EC}_{N}(f). In particular, one has the following induced action.

BrN𝖤𝖢N(f)/N\displaystyle\operatorname{Br}_{N}\curvearrowright\mathsf{EC}_{N}(f)/\mathbb{Z}^{N} (2.61)

When Perf𝒳\operatorname{Perf}\mathcal{X} admits a full ff-exceptional collection of length rr, then one similarly obtains the action Gr𝖥𝖤𝖢(𝒳)G_{r}\curvearrowright\mathsf{FEC}(\mathcal{X}).

Remark 2.23.

[BP93, Conjecture 2.2] asserts that this action should be transitive. Note that this conjecture is equivalent to the transitivity of the action (2.61). Theorem 6.1 below is nothing but the affirmative answer to [BP93, Conjecture 2.2] for 𝒳=Σ2\mathcal{X}=\Sigma_{2}.

Let XX be a smooth projective variety over the field 𝐤\mathbf{k}. The autoequivalences of 𝐃(X)\mathbf{D}(X) and the notion of exceptional collections are nicely compatible as we explain next.

Lemma 2.24.

If 1,,N𝐃(X)\mathcal{E}_{1},\dots,\mathcal{E}_{N}\in\mathbf{D}(X) is an exceptional collection and ΦAuteq(𝐃(X))\Phi\in\operatorname{Auteq}(\mathbf{D}(X)), then so is Φ(1),,Φ(N)𝐃(X)\Phi(\mathcal{E}_{1}),\dots,\Phi(\mathcal{E}_{N})\in\mathbf{D}(X). In particular, there is the natural action

Auteq(𝐃(X))𝖤𝖢N(X).\displaystyle\operatorname{Auteq}(\mathbf{D}(X))\curvearrowright\mathsf{EC}_{N}(X). (2.62)

The following lemma is easy to verify and plays an important role in this paper.

Lemma 2.25.

The actions (2.57) and (2.62) commute.

2.4. Results obtained via deformation of Σ2\Sigma_{2} to 1×1\mathbb{P}^{1}\times\mathbb{P}^{1}

Consider the (isotrivial) degeneration (2.2) over an algebraically closed field 𝐤\mathbf{k}. Consider the discrete valuation ring R=(𝐤[[t]],(t),𝐤)R=\left(\mathbf{k}[[t]],(t),\mathbf{k}\right) and take the base change by B=SpecR𝔸t1B=\operatorname{Spec}R\to\mathbb{A}^{1}_{t} of the family. We write

f:𝒳=𝒳×𝔸1BB\displaystyle f\colon\mathcal{X}=\mathcal{X}\times_{\mathbb{A}^{1}}B\to B (2.63)

by abuse of notation. The central fiber 𝒳0\mathcal{X}_{0} of ff is isomorphic to Σ2\Sigma_{2}. Also, let

ξ:K¯𝐤((t))¯B\displaystyle\xi\colon{\overline{K}}\coloneqq\overline{\mathbf{k}((t))}\to B (2.64)

be the geometric generic point of BB. The isotriviality of the family (2.2) outside the origin implies that the geometric generic fiber 𝒳ξSpecK¯\mathcal{X}_{\xi}\to\operatorname{Spec}{\overline{K}} is isomorphic to 1×1\mathbb{P}^{1}\times\mathbb{P}^{1} over K¯{\overline{K}}. Throughout this section, we freely use the symbols introduced in this paragraph.

Since the generic fiber 1×1\mathbb{P}^{1}\times\mathbb{P}^{1} is a del Pezzo surface, the properties of exceptional collections on it is very well known by [KO94]. We list the known properties.

Theorem 2.26.
  1. (1)

    Any exceptional object on 1×1\mathbb{P}^{1}\times\mathbb{P}^{1} is isomorphic to a shift of an exceptional vector bundle, so that the natural map 𝖤𝖢N(1×1)𝖤𝖢𝖵𝖡N(1×1)\mathsf{EC}_{N}(\mathbb{P}^{1}\times\mathbb{P}^{1})\hookrightarrow\mathsf{ECVB}_{N}(\mathbb{P}^{1}\times\mathbb{P}^{1}) is a bijection.

  2. (2)

    𝖤𝖢4(1×1)=𝖥𝖤𝖢(1×1)\mathsf{EC}_{4}(\mathbb{P}^{1}\times\mathbb{P}^{1})=\mathsf{FEC}(\mathbb{P}^{1}\times\mathbb{P}^{1}), and the action G4𝖥𝖤𝖢(1×1)G_{4}\curvearrowright\mathsf{FEC}(\mathbb{P}^{1}\times\mathbb{P}^{1}) is transitive.

  3. (3)

    Any exceptional collection on 1×1\mathbb{P}^{1}\times\mathbb{P}^{1} can be extended to a full exceptional collection.

The non-triviality of the group AuteqK0𝗍𝗋𝗂𝗏(𝐃(Σ2))\operatorname{Auteq}^{K_{0}-\mathsf{triv}}(\mathbf{D}(\Sigma_{2})) implies that an exceptional object on Σ2\Sigma_{2} is not uniquely determined by its class in K0K_{0}, even modulo shifts by [2]\mathbb{Z}[2]. However, if one considers only exceptional vector bundles, then it is the case:

Lemma 2.27 (==a weaker version of [OU15, Lemma 3.5]).

Let ,\mathcal{E},\mathcal{E}^{\prime} be exceptional vector bundles on Σ2\Sigma_{2} such that []=[]K0(Σ2)[\mathcal{E}]=[\mathcal{E}^{\prime}]\in\operatorname{K_{0}}\left(\Sigma_{2}\right). Then \mathcal{E}\simeq\mathcal{E}^{\prime}.

The degeneration (2.63) allows one to compare various invariants of Σ2\Sigma_{2} to those of 1×1\mathbb{P}^{1}\times\mathbb{P}^{1}. Recall that 𝒳0Σ2\mathcal{X}_{0}\simeq\Sigma_{2}.

Definition 2.28.

For an exceptional object 𝐃(𝒳0)\mathcal{E}\in\mathbf{D}(\mathcal{X}_{0}), let R,𝗀𝖾𝗇()\mathcal{E}_{R},\operatorname{\mathsf{gen}}(\mathcal{E}) denote the unique deformation of \mathcal{E} to 𝒳\mathcal{X} and its restriction to the geometric generic fiber 𝒳ξ\mathcal{X}_{\xi}, respectively. For an exceptional collection ¯\underline{\mathcal{E}} on the central fiber, we will similarly write ¯R,𝗀𝖾𝗇(¯)\underline{\mathcal{E}}_{R},\operatorname{\mathsf{gen}}(\underline{\mathcal{E}}) to mean its (unique) deformation to 𝒳\mathcal{X} and its restriction to 𝒳ξ\mathcal{X}_{\xi}, respectively.

Let

𝗀𝖾𝗇:𝖤𝖢4(𝒳0)𝖥𝖤𝖢(𝒳ξ)\displaystyle\operatorname{\mathsf{gen}}\colon\mathsf{EC}_{4}(\mathcal{X}_{0})\to\mathsf{FEC}(\mathcal{X}_{\xi}) (2.65)

be the map which sends (an isomorphism class of) an exceptional collection ¯\underline{\mathcal{E}} of length 4 on Σ2\Sigma_{2} to 𝗀𝖾𝗇(¯)𝖥𝖤𝖢(𝒳ξ)\operatorname{\mathsf{gen}}(\underline{\mathcal{E}})\in\mathsf{FEC}(\mathcal{X}_{\xi}), which is obtained by restricting the deformation of ¯\underline{\mathcal{E}} to the ff-exceptional collection, whose existence and uniqueness is guaranteed by Lemma 2.18, to the geometric generic fiber (see Corollary 2.17). See Corollary 2.37 below for the surjectivity of 𝗀𝖾𝗇\operatorname{\mathsf{gen}}.

One similarly defines the map 𝗀𝖾𝗇|𝖤𝖢𝖵𝖡4(𝒳0):𝖤𝖢𝖵𝖡4(𝒳0)𝖥𝖤𝖢𝖵𝖡(𝒳ξ)\operatorname{\mathsf{gen}|_{\mathsf{ECVB}_{4}(\mathcal{X}_{0})}}\colon\mathsf{ECVB}_{4}(\mathcal{X}_{0})\to\mathsf{FECVB}(\mathcal{X}_{\xi}), to obtain the following diagram.

𝖤𝖢4(𝒳0){\mathsf{EC}_{4}(\mathcal{X}_{0})}𝖥𝖤𝖢(𝒳ξ){\mathsf{FEC}(\mathcal{X}_{\xi})}𝖤𝖢𝖵𝖡4(𝒳0){\mathsf{ECVB}_{4}(\mathcal{X}_{0})}𝖥𝖤𝖢𝖵𝖡(𝒳ξ){\mathsf{FECVB}(\mathcal{X}_{\xi})}𝗀𝖾𝗇\scriptstyle{\operatorname{\mathsf{gen}}}𝗀𝖾𝗇|𝖤𝖢𝖵𝖡4(𝒳0)\scriptstyle{\operatorname{\mathsf{gen}|_{\mathsf{ECVB}_{4}(\mathcal{X}_{0})}}} (2.66)

We next compare K0(Perf)\operatorname{K_{0}}\left(\operatorname{Perf}\right) of the surfaces. Note that we have the following diagram of schemes (the labels of the arrows in the diagram will be freely used).

𝒳0{\mathcal{X}_{0}}𝒳{\mathcal{X}}𝒳ξ{\mathcal{X}_{\xi}}Spec𝐤{\operatorname{Spec}\mathbf{k}}SpecR{\operatorname{Spec}R}SpecK¯{\operatorname{Spec}{\overline{K}}}i\scriptstyle{i}f0\scriptstyle{f_{0}}{\lrcorner}f\scriptstyle{f}j\scriptstyle{j}fξ\scriptstyle{f_{\xi}}{\llcorner}ι\scriptstyle{\iota}ȷ¯\scriptstyle{{\overline{\jmath}}} (2.67)

Applying the functor K0(Perf())\operatorname{K_{0}}\left(\operatorname{Perf}(-)\right), we obtain the first two rows of Figure 2.68, which is a commutative diagram of commutative rings with units.

K0(𝒳0){\operatorname{K_{0}}\left(\mathcal{X}_{0}\right)}K0(𝒳){\operatorname{K_{0}}\left(\mathcal{X}\right)}K0(𝒳ξ){\operatorname{K_{0}}\left(\mathcal{X}_{\xi}\right)}K0(𝐤){\operatorname{K_{0}}\left(\mathbf{k}\right)}K0(R){\operatorname{K_{0}}\left(R\right)}K0(K¯){\operatorname{K_{0}}\left({\overline{K}}\right)}{\mathbb{Z}}{\mathbb{Z}}{\mathbb{Z}}f0\scriptstyle{f_{0\ast}}\scriptstyle{\simeq}i\scriptstyle{i^{\ast}}j\scriptstyle{j^{\ast}}\scriptstyle{\simeq}f\scriptstyle{f_{\ast}}fξ\scriptstyle{f_{\xi\ast}}\scriptstyle{\simeq}ι\scriptstyle{\iota^{\ast}}ȷ¯\scriptstyle{{\overline{\jmath}}^{\ast}}\scriptstyle{\simeq}\scriptstyle{\simeq}=\scriptstyle{=}=\scriptstyle{=} (2.68)

The derived dual defined in (2.6) induces an automorphism of commutative rings

K0(Perf𝒳)=K0((Perf𝒳)op)K0(Perf𝒳);[][].\displaystyle\operatorname{K_{0}}\left(\operatorname{Perf}\mathcal{X}\right)=\operatorname{K_{0}}\left(\left(\operatorname{Perf}\mathcal{X}\right)^{\mathrm{op}}\right)\xrightarrow[\simeq]{\vee}\operatorname{K_{0}}\left(\operatorname{Perf}\mathcal{X}\right);\quad[\mathcal{E}]\mapsto[\mathcal{E}^{\vee}]. (2.69)

The Euler pairing on K0(Perf𝒳)=K0(𝒳)\operatorname{K_{0}}\left(\operatorname{Perf}\mathcal{X}\right)=\operatorname{K_{0}}\left(\mathcal{X}\right) (note that 𝒳\mathcal{X} is a regular scheme) is the following bilinear pairing.

χ𝒳:K0(𝒳)×K0(𝒳)K0(R);(v,w)f(vw)\displaystyle\chi_{\mathcal{X}}\colon\operatorname{K_{0}}\left(\mathcal{X}\right)\times\operatorname{K_{0}}\left(\mathcal{X}\right)\to\operatorname{K_{0}}\left(R\right);\quad(v,w)\mapsto f_{\ast}\left(v^{\vee}\cdot w\right) (2.70)

We can similarly define

χ𝒳0:K0(𝒳0)×K0(𝒳0)K0(𝐤);(v,w)f0(vw),\displaystyle\chi_{\mathcal{X}_{0}}\colon\operatorname{K_{0}}\left(\mathcal{X}_{0}\right)\times\operatorname{K_{0}}\left(\mathcal{X}_{0}\right)\to\operatorname{K_{0}}\left(\mathbf{k}\right);\quad(v,w)\mapsto f_{0\ast}\left(v^{\vee}\cdot w\right), (2.71)
χ𝒳ξ:K0(𝒳ξ)×K0(𝒳ξ)K0(K¯);(v,w)fξ(vw).\displaystyle\chi_{\mathcal{X}_{\xi}}\colon\operatorname{K_{0}}\left(\mathcal{X}_{\xi}\right)\times\operatorname{K_{0}}\left(\mathcal{X}_{\xi}\right)\to\operatorname{K_{0}}\left({\overline{K}}\right);\quad(v,w)\mapsto f_{\xi\ast}\left(v^{\vee}\cdot w\right). (2.72)
Lemma 2.29.
ιχ𝒳=χ𝒳0(i×i),\displaystyle\iota^{\ast}\circ\chi_{\mathcal{X}}=\chi_{\mathcal{X}_{0}}\circ(i^{\ast}\times i^{\ast}), (2.73)
ȷ¯χ𝒳=χ𝒳ξ(j×j)\displaystyle{\overline{\jmath}}^{\ast}\circ\chi_{\mathcal{X}}=\chi_{\mathcal{X}_{\xi}}\circ(j^{\ast}\times j^{\ast}) (2.74)

Let ¯\underline{\mathcal{E}} be a full exceptional collection of 𝒳0\mathcal{X}_{0}, and ¯R,¯ξ,\underline{\mathcal{E}}_{R},\underline{\mathcal{E}}_{\xi}, be the deformation of ¯\underline{\mathcal{E}} to 𝒳\mathcal{X} and the restriction of ¯R\underline{\mathcal{E}}_{R} to 𝒳ξ\mathcal{X}_{\xi} as defined in Definition 2.28. As pointed out in Remark 2.19, both ¯R,¯ξ\underline{\mathcal{E}}_{R},\underline{\mathcal{E}}_{\xi} are full exceptional collections.

Lemma 2.30.

K0(¯),K0(¯R).K0(¯ξ)\operatorname{K_{0}}\left(\underline{\mathcal{E}}\right),\operatorname{K_{0}}\left(\underline{\mathcal{E}}_{R}\right).\operatorname{K_{0}}\left(\underline{\mathcal{E}}_{\xi}\right) are bases of K0(𝒳0),K0(𝒳),K0(𝒳ξ)\operatorname{K_{0}}\left(\mathcal{X}_{0}\right),\operatorname{K_{0}}\left(\mathcal{X}\right),\operatorname{K_{0}}\left(\mathcal{X}_{\xi}\right), respectively. In particular, the horizontal maps i,ji^{\ast},j^{\ast} in the first row of (2.68) are isomorphisms.

Proof.

Immediately follows from the fact that the collections ¯,¯R,¯ξ\underline{\mathcal{E}},\underline{\mathcal{E}}_{R},\underline{\mathcal{E}}_{\xi} are full exceptional collections of the triangulated categories over the base of length 44. ∎

On the other hand we obtain the bottom row of the diagram Figure 2.68, where the vertical maps to the bottom row are isomorphisms of rings. Let

K0(𝗀𝖾𝗇):K0(𝒳0)K0(𝒳ξ)\displaystyle\operatorname{K_{0}}\left(\operatorname{\mathsf{gen}}\right)\colon\operatorname{K_{0}}\left(\mathcal{X}_{0}\right)\to\operatorname{K_{0}}\left(\mathcal{X}_{\xi}\right) (2.75)

be the isomorphism of abelian groups obtained from the diagram Figure 2.68. Also, regard χ𝒳0,χ𝒳ξ\chi_{\mathcal{X}_{0}},\chi_{\mathcal{X}_{\xi}} as \mathbb{Z}-valued bilinear pairings by the diagram Figure 2.68. With all the preparations above, we can show the desired properties of the map K0(𝗀𝖾𝗇)\operatorname{K_{0}}\left(\operatorname{\mathsf{gen}}\right).

Proposition 2.31.

The isomorphism K0(𝗀𝖾𝗇)\operatorname{K_{0}}\left(\operatorname{\mathsf{gen}}\right) of (2.75) respects the pairings χ𝒳0,χ𝒳ξ\chi_{\mathcal{X}_{0}},\chi_{\mathcal{X}_{\xi}} on the source and the target abelian groups. Moreover, it fits in the following commutative diagram.

𝖤𝖢1(𝒳0){\mathsf{EC}_{1}(\mathcal{X}_{0})}𝖤𝖢1(𝒳ξ){\mathsf{EC}_{1}(\mathcal{X}_{\xi})}K0(𝒳0){\operatorname{K_{0}}\left(\mathcal{X}_{0}\right)}K0(𝒳ξ){\operatorname{K_{0}}\left(\mathcal{X}_{\xi}\right)}𝗀𝖾𝗇\scriptstyle{\operatorname{\mathsf{gen}}}K0()\scriptstyle{\operatorname{K_{0}}\left(\right)}K0()\scriptstyle{\operatorname{K_{0}}\left(\right)}K0(𝗀𝖾𝗇)\scriptstyle{\operatorname{K_{0}}\left(\operatorname{\mathsf{gen}}\right)}\scriptstyle{\simeq}
Proposition 2.32.

Let ,𝐃(𝒳0)\mathcal{E},\mathcal{E}^{\prime}\in\mathbf{D}(\mathcal{X}_{0}) be exceptional objects. The following conditions are equivalent.

  1. (1)

    []=[]K0(𝒳0)[\mathcal{E}]=[\mathcal{E}^{\prime}]\in\operatorname{K_{0}}\left(\mathcal{X}_{0}\right).

  2. (2)

    𝗀𝖾𝗇()𝗀𝖾𝗇()[2m]\operatorname{\mathsf{gen}}(\mathcal{E})\simeq\operatorname{\mathsf{gen}}(\mathcal{E}^{\prime})[2m] for some mm\in\mathbb{Z}.

Proof.

(2) \Rightarrow (1) is an consequence of the commutative diagram above and that K0(𝗀𝖾𝗇)\operatorname{K_{0}}\left(\operatorname{\mathsf{gen}}\right) is an isomorphism.

Conversely, assume (1). It then follows that [𝗀𝖾𝗇()]=[𝗀𝖾𝗇()]K0(𝒳ξ)[\operatorname{\mathsf{gen}}(\mathcal{E})]=[\operatorname{\mathsf{gen}}(\mathcal{E}^{\prime})]\in\operatorname{K_{0}}\left(\mathcal{X}_{\xi}\right), which in turn implies (2) by Theorem 1.1 (2). ∎

For exceptional vector bundles on Σ2𝒳0\Sigma_{2}\simeq\mathcal{X}_{0}, we have the following reconstruction result. This is an immediate corollary of Lemma 2.27.

Lemma 2.33.

For any N=1,,4N=1,\dots,4, the map 𝖤𝖢𝖵𝖡N(Σ2)K0()𝗇𝗎𝗆𝖤𝖢N(Σ2)\mathsf{ECVB}_{N}(\Sigma_{2})\xrightarrow{\operatorname{K_{0}}\left(\right)}\mathsf{numEC}_{N}(\Sigma_{2}) is injective.

See Definition 2.34 for the definition of the set 𝗇𝗎𝗆𝖤𝖢N(Σ2)\mathsf{numEC}_{N}(\Sigma_{2}).

Definition 2.34.

Let SS be a smooth projective variety, and for simplicity let us assume that K0(S)\operatorname{K_{0}}\left(S\right) is isomorphic to the numerical Grothendieck group; i.e., the Euler pairing χ\chi is non-degenerate. This in particular implies that K0(S)\operatorname{K_{0}}\left(S\right) is a free abelian group of finite rank.

An exceptional vector is an element eK0(S)e\in\operatorname{K_{0}}\left(S\right) such that e2=χS(e,e)=1e^{2}=\chi_{S}(e,e)=1.

A numerical exceptional collection on SS is a sequence of exceptional vectors e1,,eNK0(S)e_{1},\dots,e_{N}\in\operatorname{K_{0}}\left(S\right) such that χS(ej,ei)=0\chi_{S}(e_{j},e_{i})=0 for any 1i<jN1\leq i<j\leq N. A numerical exceptional collection is said to be full if it is a basis of K0(S)\operatorname{K_{0}}\left(S\right); i.e, when N=rankK0(S)N=\operatorname{rank}\operatorname{K_{0}}\left(S\right).

The set of numerical exceptional collections of length NN (resp. full) on SS will be denoted by 𝗇𝗎𝗆𝖤𝖢N(S)\mathsf{numEC}_{N}(S) and 𝗇𝗎𝗆𝖥𝖤𝖢(S)\mathsf{numFEC}(S), respectively.

For a numerical exceptional collection (e,f)(e,f) of length 22, which will also be called a numerical exceptional pair, its right and left mutations are the new numerical exceptional pairs defined and denoted as follows.

(f,Rf(e)eχS(e,f)f)\displaystyle\left(f,R_{f}(e)\coloneqq e-\chi_{S}(e,f)f\right) (2.76)
(Le(f)fχS(e,f)e,e)\displaystyle\left(L_{e}(f)\coloneqq f-\chi_{S}(e,f)e,e\right) (2.77)

By similar arguments for the action GN𝖤𝖢NG_{N}\curvearrowright\mathsf{EC}_{N}, one can verify that for a surface SS with K0(S)N\operatorname{K_{0}}\left(S\right)\simeq\mathbb{Z}^{N} there is an action GN𝗇𝗎𝗆𝖥𝖤𝖢(S)G_{N}\curvearrowright\mathsf{numFEC}(S), where the subgroup BrN\operatorname{Br}_{N} acts by the mutations (2.76) (2.77) and N\mathbb{Z}^{N} by the change of signs (hence the action descends to the quotient (/2)NBrN\left(\mathbb{Z}/2\mathbb{Z}\right)^{N}\rtimes\operatorname{Br}_{N}).

At last we claim that everything goes together.

Proposition 2.35.

There exists the following commutative diagram of sets equipped with the action of the group G4=4Br4G_{4}=\mathbb{Z}^{4}\rtimes\operatorname{Br}_{4}.

𝖤𝖢4(𝒳0){\mathsf{EC}_{4}(\mathcal{X}_{0})}𝖥𝖤𝖢(𝒳ξ){\mathsf{FEC}(\mathcal{X}_{\xi})}𝗇𝗎𝗆𝖥𝖤𝖢(𝒳0){\mathsf{numFEC}(\mathcal{X}_{0})}𝗇𝗎𝗆𝖥𝖤𝖢(𝒳ξ){\mathsf{numFEC}(\mathcal{X}_{\xi})}𝗀𝖾𝗇\scriptstyle{\operatorname{\mathsf{gen}}}K0()\scriptstyle{\operatorname{K_{0}}\left(\right)}K0()\scriptstyle{\operatorname{K_{0}}\left(\right)}K0(𝗀𝖾𝗇)\scriptstyle{\operatorname{K_{0}}\left(\operatorname{\mathsf{gen}}\right)}\scriptstyle{\simeq}

Let us now introduce a particular (ff-)exceptional collection of invertible sheaves, which will be called the standard collection and serve as the base point of the sets 𝖥𝖤𝖢\mathsf{FEC} and 𝖥𝖤𝖢𝖵𝖡\mathsf{FECVB} in this paper.

Definition 2.36.

The standard full exceptional collection of invertible sheaves on Σ2𝒳0\Sigma_{2}\simeq\mathcal{X}_{0} is defined as follows.

𝗌𝗍𝖽(𝒪Σ2,𝒪Σ2(f),𝒪Σ2(C+2f),𝒪Σ2(C+3f))𝖥𝖤𝖢𝖵𝖡(Σ2).\displaystyle\mathcal{E}^{\mathsf{std}}\coloneqq\left(\mathcal{O}_{\Sigma_{2}},\mathcal{O}_{\Sigma_{2}}(f),\mathcal{O}_{\Sigma_{2}}(C+2f),\mathcal{O}_{\Sigma_{2}}(C+3f)\right)\in\mathsf{FECVB}(\Sigma_{2}). (2.78)

By Lemma 2.18 and Remark 2.19, it uniquely deforms to a full ff-exceptional collection of invertible sheaves. We will write it R𝗌𝗍𝖽\mathcal{E}^{\mathsf{std}}_{R}, and its pullback to 𝒳ξ\mathcal{X}_{\xi} will be denoted by ξ𝗌𝗍𝖽\mathcal{E}^{\mathsf{std}}_{\xi}. Using the deformation invariance of the intersection numbers, one can easily confirm that

ξ𝗌𝗍𝖽=(𝒪𝒳ξ,𝒪𝒳ξ(1,0),𝒪𝒳ξ(1,1),𝒪𝒳ξ(2,1))\displaystyle\mathcal{E}^{\mathsf{std}}_{\xi}=\left(\mathcal{O}_{\mathcal{X}_{\xi}},\mathcal{O}_{\mathcal{X}_{\xi}}(1,0),\mathcal{O}_{\mathcal{X}_{\xi}}(1,1),\mathcal{O}_{\mathcal{X}_{\xi}}(2,1)\right) (2.79)

under an isomorphism 𝒳ξ1×1\mathcal{X}_{\xi}\simeq\mathbb{P}^{1}\times\mathbb{P}^{1}.

Corollary 2.37.

The map 𝗀𝖾𝗇:𝖤𝖢4(𝒳0)𝖥𝖤𝖢(𝒳ξ)\operatorname{\mathsf{gen}}\colon\mathsf{EC}_{4}(\mathcal{X}_{0})\to\mathsf{FEC}(\mathcal{X}_{\xi}) is surjective.

Proof.

By Theorem 2.26 (2), we know that 𝖥𝖤𝖢(𝒳ξ)=G4ξ𝗌𝗍𝖽\mathsf{FEC}(\mathcal{X}_{\xi})=G_{4}\cdot\mathcal{E}^{\mathsf{std}}_{\xi}. Since the diagram of Proposition 2.35 is G4G_{4}-equivariant and 𝗀𝖾𝗇(𝗌𝗍𝖽)=ξ𝗌𝗍𝖽\operatorname{\mathsf{gen}}(\mathcal{E}^{\mathsf{std}})=\mathcal{E}^{\mathsf{std}}_{\xi}, we obtain the conclusion. ∎

Corollary 2.38.

For any ¯𝖤𝖢4(𝒳0)\underline{\mathcal{E}}\in\mathsf{EC}_{4}(\mathcal{X}_{0}), there exists σG4\sigma\in G_{4} such that K0(σ(¯))=K0(𝗌𝗍𝖽)𝗇𝗎𝗆𝖥𝖤𝖢(𝒳0)\operatorname{K_{0}}\left(\sigma(\underline{\mathcal{E}})\right)=\operatorname{K_{0}}\left(\mathcal{E}^{\mathsf{std}}\right)\in\mathsf{numFEC}(\mathcal{X}_{0}).

Proof.

Again by Theorem 2.26 (2), one can find an element σG4\sigma\in G_{4} such that σ𝗀𝖾𝗇(¯)=ξ𝗌𝗍𝖽\sigma\operatorname{\mathsf{gen}}(\underline{\mathcal{E}})=\mathcal{E}^{\mathsf{std}}_{\xi}. Again by the equivariance, it follows that 𝗀𝖾𝗇(σ(¯))=ξ𝗌𝗍𝖽=𝗀𝖾𝗇(𝗌𝗍𝖽)\operatorname{\mathsf{gen}}(\sigma(\underline{\mathcal{E}}))=\mathcal{E}^{\mathsf{std}}_{\xi}=\operatorname{\mathsf{gen}}(\mathcal{E}^{\mathsf{std}}), which means that K0(σ(¯))=K0(𝗌𝗍𝖽)\operatorname{K_{0}}\left(\sigma(\underline{\mathcal{E}})\right)=\operatorname{K_{0}}\left(\mathcal{E}^{\mathsf{std}}\right) by Proposition 2.32. ∎

3. Twisting exceptional objects down to exceptional vector bundles

The purpose of this section is to show the following theorem.

Theorem 3.1.

For any exceptional object 𝐃(Σ2)\mathcal{E}\in\mathbf{D}(\Sigma_{2}), there exists an exceptional vector bundle \mathcal{F} and a sequence of integers a1,,an,ma_{1},\dots,a_{n},m such that

(TanTa1)()[m].\displaystyle\mathcal{E}\simeq\left(T_{a_{n}}\circ\cdots\circ T_{a_{1}}\right)(\mathcal{F})[m]. (3.1)

The similar result for spherical objects is given in [IU05, Proposition 1.6]. In fact, we prove Theorem 3.1 by suitably modifying the proof of [IU05, Proposition 1.6].

Notation 3.2.

Let XX be an integral noetherian scheme. For cohX\mathcal{E}\in\operatorname{coh}X, we define

SuppSpecIm(𝒪XndX())X\displaystyle\operatorname{Supp}\mathcal{E}\coloneqq\operatorname{Spec}\operatorname{Im}\left(\mathcal{O}_{X}\to\mathop{{\mathcal{E}}nd}\nolimits_{X}(\mathcal{E})\right)\subset X (3.2)

and call it the schematic support of \mathcal{E}. It is universal among the closed subschemes ι:ZX\iota\colon Z\hookrightarrow X which admits a coherent sheaf cohZ\mathcal{E}^{\prime}\in\operatorname{coh}Z such that ι\iota_{\ast}\mathcal{E}^{\prime}\simeq\mathcal{E}. The underlying closed subset of XX, or equivalently the reduced closed subscheme (Supp)𝗋𝖾𝖽X(\operatorname{Supp}\mathcal{E})_{\operatorname{\mathsf{red}}}\subset X, is called the reduced support of \mathcal{E}. Also we let tors\operatorname{tors}\mathcal{E}\subset\mathcal{E} be the maximum torsion subsheaf of \mathcal{E}. For an object 𝐃(X)\mathcal{E}\in\mathbf{D}(X), we use the following notation.

()ii()Suppi(Suppi())𝗋𝖾𝖽\displaystyle\mathcal{H}^{\bullet}(\mathcal{E})\coloneqq\bigoplus_{i\in\mathbb{Z}}\mathcal{H}^{i}(\mathcal{E})\quad\operatorname{Supp}\mathcal{E}\coloneqq\bigcup_{i\in\mathbb{Z}}\left(\operatorname{Supp}\mathcal{H}^{i}(\mathcal{E})\right)_{\operatorname{\mathsf{red}}} (3.3)

3.1. First properties of ()\mathcal{H}^{\bullet}(\mathcal{E})

As the first step toward the proof of Theorem 3.1, in this subsection we prove some basic properties of ()\mathcal{H}^{\bullet}(\mathcal{E}). Part of them concern (Suppi())𝗋𝖾𝖽\left(\operatorname{Supp}\mathcal{H}^{i}(\mathcal{E})\right)_{\operatorname{\mathsf{red}}}, which can be summarized as follows.

There exists the unique integer i0i_{0}\in\mathbb{Z} with the following properties.

  • Suppi0()=Σ2\operatorname{Supp}\mathcal{H}^{i_{0}}(\mathcal{E})=\Sigma_{2}

  • (Supptorsi0())𝗋𝖾𝖽=C\left(\operatorname{Supp}\operatorname{tors}\mathcal{H}^{i_{0}}(\mathcal{E})\right)_{\operatorname{\mathsf{red}}}=C

  • (Suppi())𝗋𝖾𝖽=C\left(\operatorname{Supp}\mathcal{H}^{i}(\mathcal{E})\right)_{\operatorname{\mathsf{red}}}=C for ii0\forall i\neq i_{0}

Based on this, the similar statements for the schematic supports will be proved in the next subsection; namely, we will remove 𝗋𝖾𝖽\operatorname{\mathsf{red}} from the second and the third items. In fact, this step has been the main obstacle for the project.

The similar results for spherical objects appear in [IU05, Lemma 4.8], where it is rather easily shown that the schematic support of the cohomology sheaves of a spherical object coincides with CC. However, unfortunately, the proof of [IU05, Lemma 4.8] does not immediately apply to our situation. The fact Supp=Σ2\operatorname{Supp}\mathcal{E}=\Sigma_{2} prevents us from studying the problem locally around the curve CC.

Lemma 3.3.

The cohomology sheaves of an exceptional object 𝐃(Σ2)\mathcal{E}\in\mathbf{D}(\Sigma_{2}) enjoy the following properties.

  1. (1)

    ()\mathcal{H}^{\bullet}(\mathcal{E}) is rigid (E21,q=0q\iff E_{2}^{1,q}=0\ \forall q\in\mathbb{Z} in the spectral sequence (3.4)).

  2. (2)

    There exists a unique integer i0i_{0}\in\mathbb{Z} such that

    • Suppi0()=Σ2\operatorname{Supp}\mathcal{H}^{i_{0}}(\mathcal{E})=\Sigma_{2}.

    • i()\mathcal{H}^{i}(\mathcal{E}) for ii0i\neq i_{0} and torsi0()\operatorname{tors}\mathcal{H}^{i_{0}}(\mathcal{E}) are pure sheaves with (Supp())𝗋𝖾𝖽=C\left(\operatorname{Supp}(-)\right)_{\operatorname{\mathsf{red}}}=C, unless =0=0.

  3. (3)

    tors(i0())\operatorname{tors}(\mathcal{H}^{i_{0}}(\mathcal{E})) is rigid, and i0()/tors(i0())\mathcal{H}^{i_{0}}(\mathcal{E})/\operatorname{tors}(\mathcal{H}^{i_{0}}(\mathcal{E})) is an exceptional vector bundle.

See the following definition for the notion of rigidity.

Definition 3.4.

An object 𝐃(X)\mathcal{E}\in\mathbf{D}(X) is said to be rigid if ExtX1(,)=0\operatorname{Ext}_{X}^{1}(\mathcal{E},\mathcal{E})=0.

We will freely use the following standard fact on rigid objects.

Lemma 3.5.

Let 𝐃(X)\mathcal{E}\in\mathbf{D}(X) be a rigid object. Then gg\mathcal{E}\simeq\mathcal{E} for any gAutX/𝐤0g\in\operatorname{Aut}^{0}_{X/\mathbf{k}}. In particular, Suppg=SuppX\operatorname{Supp}g\mathcal{E}=\operatorname{Supp}\mathcal{E}\subset X for any such gg.

We need the following spectral sequence, which also plays the central role for the study of spherical objects in [IU05].

Lemma 3.6.

For any object 𝐃(X)\mathcal{E}\in\mathbf{D}(X), there exists the following spectral sequence.

E2p,q=iExtΣ2p(i(),i+q())Ep+q=HomΣ2p+q(,)\displaystyle E_{2}^{p,q}=\bigoplus_{i}\operatorname{Ext}_{\Sigma_{2}}^{p}(\mathcal{H}^{i}(\mathcal{E}),\mathcal{H}^{i+q}(\mathcal{E}))\Rightarrow E^{p+q}=\operatorname{Hom}_{\Sigma_{2}}^{p+q}(\mathcal{E},\mathcal{E}) (3.4)

Moreover, using the classes ei=ei()ExtΣ22(i(),i1())e^{i}=e^{i}(\mathcal{E})\in\operatorname{Ext}_{\Sigma_{2}}^{2}(\mathcal{H}^{i}(\mathcal{E}),\mathcal{H}^{i-1}(\mathcal{E})) canonically determined by \mathcal{E}, the d2d_{2} maps of (3.4) are given by

d2p,q:(ϕi)i((1)p+qϕi1eiei+qϕi)i.\displaystyle d_{2}^{p,q}\colon({\phi_{i}})_{i}\mapsto\left((-1)^{p+q}\phi_{i-1}\circ e^{i}-e^{i+q}\circ\phi_{i}\right)_{i}. (3.5)
Proof.

See [IU05, Section 4.1], in particular [IU05, Proposition 4.1]. ∎

Proof of Lemma 3.3.

Throughout the proof, we consider the spectral sequence (3.4) for the exceptional object \mathcal{E}. We prove the four items one by one.

(1) The exceptionality of \mathcal{E} is translated into the following conditions.

En={𝐤idn=00n0\displaystyle E^{n}=\begin{cases}\mathbf{k}\operatorname{id}_{\mathcal{E}}&n=0\\ 0&n\neq 0\end{cases} (3.6)

Since Σ2\Sigma_{2} is a smooth projective surface, E2p,q=0E_{2}^{p,q}=0 unless 0p20\leq p\leq 2. Hence (3.4) is E3E_{3}-degenerate, and moreover is E2E_{2}-degenerate at (1,q)(1,q) for any qq\in\mathbb{Z}. Combined with (3.6), this implies that E21,q=0E_{2}^{1,q}=0 for any q1q\neq-1. In the next paragraph we also confirm E21,1=0E_{2}^{1,-1}=0, thereby concluding the rigidity of ()\mathcal{H}^{\bullet}(\mathcal{E}).

It follows from the explicit description (3.5) of d2d_{2} maps that 0iidi()E20,00\neq\sum_{i\in\mathbb{Z}}\operatorname{id}_{\mathcal{H}^{i}(\mathcal{E})}\in E_{2}^{0,0} is in fact contained in Kerd20,0E30,0E0,0\operatorname{Ker}d_{2}^{0,0}\simeq E_{3}^{0,0}\simeq E_{\infty}^{0,0}, which implies that E0,00E_{\infty}^{0,0}\neq 0. Combined with the isomorphism E0𝐤E^{0}\simeq\mathbf{k} from (3.6) and the epimorphism E0E0/F1E0E0,0E^{0}\twoheadrightarrow E^{0}/F^{1}E^{0}\simeq E_{\infty}^{0,0}, this implies that E0,0E0(𝐤)E_{\infty}^{0,0}\simeq E^{0}(\simeq\mathbf{k}) and hence E21,1E1,1F1E0/F2E0=0E_{2}^{1,-1}\simeq E_{\infty}^{1,-1}\simeq F^{1}E^{0}/F^{2}E^{0}=0.

(2) We further obtain the following equalities from (3.6).

e20,q\displaystyle e_{2}^{0,q} =e22,q1(if q0)\displaystyle=e_{2}^{2,q-1}\quad(\text{if $q\neq 0$}) (3.7)
e20,0\displaystyle e_{2}^{0,0} =e22,1+1\displaystyle=e_{2}^{2,-1}+1 (3.8)

To see (3.7) for q=1q=-1, note that the arguments in the previous paragraph implies 0=E2,2E32,20=E_{\infty}^{2,-2}\simeq E_{3}^{2,-2}. Moreover, (1) implies that either Supp(torsi())𝗋𝖾𝖽=Σ2,C,or\operatorname{Supp}\left(\operatorname{tors}\mathcal{H}^{i}(\mathcal{E})\right)_{\operatorname{\mathsf{red}}}=\Sigma_{2},C,\ \text{or}\ \emptyset. This follows from Lemma 3.5 and the orbit decomposition Σ2=(Σ2C)C\Sigma_{2}=\left(\Sigma_{2}\setminus C\right)\coprod C for the action AutΣ2/𝐤Σ2\operatorname{Aut}_{\Sigma_{2}/\mathbf{k}}\curvearrowright\Sigma_{2}.

Suppose for a contradiction that Suppi()𝗋𝖾𝖽Σ2\operatorname{Supp}\mathcal{H}^{i}(\mathcal{E})_{\operatorname{\mathsf{red}}}\neq\Sigma_{2} for every ii. Then Supp\operatorname{Supp}\mathcal{E} coincides with CC and hence there is an isomorphism ωΣ2\mathcal{E}\otimes\omega_{\Sigma_{2}}\simeq\mathcal{E}, from which we deduce extΣ22(,)=extΣ20(,)\operatorname{ext}_{\Sigma_{2}}^{2}(\mathcal{E},\mathcal{E})=\operatorname{ext}_{\Sigma_{2}}^{0}(\mathcal{E},\mathcal{E}) by the Serre duality. This contradicts (3.6) and hence there must be at least one integer i0i_{0} with Suppi0()𝗋𝖾𝖽=Σ2\operatorname{Supp}\mathcal{H}^{i_{0}}(\mathcal{E})_{\operatorname{\mathsf{red}}}=\Sigma_{2}.

Now consider the Serre duality

(E22,q)iHomΣ2(i(),iq()ωΣ2).\displaystyle(E_{2}^{2,q})^{\vee}\simeq\bigoplus_{i}\operatorname{Hom}_{\Sigma_{2}}(\mathcal{H}^{i}(\mathcal{E}),\mathcal{H}^{i-q}(\mathcal{E})\otimes\omega_{\Sigma_{2}}). (3.9)

Fix a non-trivial morphism s:ωΣ2𝒪Σ2s\colon\omega_{\Sigma_{2}}\hookrightarrow\mathcal{O}_{\Sigma_{2}} such that the support of its cokernel is disjoint from the set of associated points of i()\mathcal{H}^{i}(\mathcal{E}) for all ii\in\mathbb{Z}. This is possible, as the linear system |KΣ2||-K_{\Sigma_{2}}| is base point free and there are only finitely many (schematic) points to be avoided (see [HL10, p. 8]). The property which we required for ss implies that for each ii\in\mathbb{Z} the natural morphism

iq()ωΣ2idsiq()\displaystyle\mathcal{H}^{i-q}(\mathcal{E})\otimes\omega_{\Sigma_{2}}\stackrel{{\scriptstyle\operatorname{id}\otimes s}}{{\to}}\mathcal{H}^{i-q}(\mathcal{E}) (3.10)

is injective. Thus we obtain an injection of vector spaces

ϕq:(E22,q)E20,q\phi_{q}\colon(E_{2}^{2,q})^{\vee}\hookrightarrow E_{2}^{0,-q}

and hence an inequality

e22,qe20,q\displaystyle e_{2}^{2,q}\leq e_{2}^{0,-q} (3.11)

for any qq.

Next we prove that ϕ0\phi_{0} is not surjective on the direct summands indexed by those ii with Suppi()=Σ2\operatorname{Supp}\mathcal{H}^{i}(\mathcal{E})=\Sigma_{2}. To see this, (by slight abuse of notation) let i0i_{0} be one of such indices and apply the functor Hom(i0(),)\operatorname{Hom}(\mathcal{H}^{i_{0}}(\mathcal{E}),-) to the following short exact sequence.

0i0()ωΣ2idsi0()i0()|Z(s)0.\displaystyle 0\to\mathcal{H}^{i_{0}}(\mathcal{E})\otimes\omega_{\Sigma_{2}}\xrightarrow[]{\operatorname{id}\otimes s}\mathcal{H}^{i_{0}}(\mathcal{E})\to\mathcal{H}^{i_{0}}(\mathcal{E})|_{Z(s)}\to 0. (3.12)

The rigidity of i0()\mathcal{H}^{i_{0}}(\mathcal{E}), which we confirmed in (1), implies that we have the following short exact sequence.

0Hom(i0(),i0()ωΣ2)Hom(i0(),i0())Hom(i0(),i0()|Z(s))0\displaystyle 0\to\operatorname{Hom}\left(\mathcal{H}^{i_{0}}(\mathcal{E}),\mathcal{H}^{i_{0}}(\mathcal{E})\otimes\omega_{\Sigma_{2}}\right)\to\operatorname{Hom}\left(\mathcal{H}^{i_{0}}(\mathcal{E}),\mathcal{H}^{i_{0}}(\mathcal{E})\right)\to\operatorname{Hom}\left(\mathcal{H}^{i_{0}}(\mathcal{E}),\mathcal{H}^{i_{0}}(\mathcal{E})|_{Z(s)}\right)\to 0 (3.13)

The 3rd term is not 0 by the assumption Suppi0()=Σ2\operatorname{Supp}\mathcal{H}^{i_{0}}(\mathcal{E})=\Sigma_{2} and hence ids\operatorname{id}\otimes s, which is identified with the direct summand of ϕ0\phi_{0} of interest, is not surjective.

Since we showed above that there is at least one such index i0i_{0}, we have confirmed the non-surjectivity of ϕ0\phi_{0} and hence the following inequality.

e22,0<e20,0\displaystyle e_{2}^{2,0}<e_{2}^{0,0} (3.14)

Summarizing the results so far, we obtain the following sequence of (in)equalities.

e22,1(3.11)(q=1)e20,1=(3.7)(q=1)e22,0=(3.14)e20,0rankϕ0=(3.8)e22,1+1rankϕ0e22,1,e_{2}^{2,-1}\stackrel{{\scriptstyle\eqref{eq:e_2^{2,q} <= e_2^{0, -q}}(q=-1)}}{{\leq}}e_{2}^{0,1}\stackrel{{\scriptstyle\eqref{eq:e^{0,q} = e^{2,q-1}}(q=1)}}{{=}}e_{2}^{2,0}\stackrel{{\scriptstyle\eqref{eq:e_2^{2,0} < e_2^{0,0}}}}{{=}}e_{2}^{0,0}-\operatorname{rank}\phi_{0}\stackrel{{\scriptstyle\eqref{eq:e^{0,0} = e^{2,-1}+1}}}{{=}}e_{2}^{2,-1}+1-\operatorname{rank}\phi_{0}\leq e_{2}^{2,-1},

which implies that

rankϕ0=e20,0e22,0=1.\displaystyle\operatorname{rank}\phi_{0}=e_{2}^{0,0}-e_{2}^{2,0}=1. (3.15)

This means that the number of the indices i0i_{0} with Suppi0()=Σ2\operatorname{Supp}\mathcal{H}^{i_{0}}(\mathcal{E})=\Sigma_{2} is exactly one.

Finally, recall that rigid sheaf with one-dimensional support is pure by [Kul97, Corollary 2.2.3]. We already confirmed the rigidity of i()\mathcal{H}^{i}(\mathcal{E}) for ii0i\neq i_{0} above, and for torsi0()\operatorname{tors}\mathcal{H}^{i_{0}}(\mathcal{E}) it is proven below.

(3) Put i0()\mathcal{H}\coloneqq\mathcal{H}^{i_{0}}(\mathcal{E}), 𝒯tors()\mathcal{T}\coloneqq\operatorname{tors}(\mathcal{H}) and /𝒯\mathcal{F}\coloneqq\mathcal{H}/\mathcal{T}. Consider the spectral sequence

E1p,q=jExtΣ2p+q(Gj,Gj+p)Ep+q=ExtΣ2p+q(,)\displaystyle E_{1}^{p,q}=\bigoplus_{j}\operatorname{Ext}_{\Sigma_{2}}^{p+q}(G_{j},G_{j+p})\Rightarrow E^{p+q}=\operatorname{Ext}_{\Sigma_{2}}^{p+q}(\mathcal{H},\mathcal{H}) (3.16)

arising from the short exact sequence 0G2G100\to G_{2}\to\mathcal{H}\to G_{1}\to 0, where G1=G_{1}=\mathcal{F} and G2=𝒯G_{2}=\mathcal{T}. For obvious reasons we see E1p,q=0E_{1}^{p,q}=0 unless 0p+q20\leq p+q\leq 2 and 1p1-1\leq p\leq 1.

On the other hand, since \mathcal{H} is rigid, it is stable under the action of AutΣ2/𝐤=AutΣ2/𝐤0\operatorname{Aut}_{\Sigma_{2}/\mathbf{k}}=\operatorname{Aut}_{\Sigma_{2}/\mathbf{k}}^{0} by Lemma 3.5. Since the torsion part of a coherent sheaf is uniquely determined by the sheaf, it follows that 𝒯\mathcal{T} is also stable under the same group action. Hence it follows that (Supp𝒯)𝗋𝖾𝖽C\left(\operatorname{Supp}\mathcal{T}\right)_{\operatorname{\mathsf{red}}}\subset C, so that

𝒯ωΣ2𝒯.\displaystyle\mathcal{T}\otimes\omega_{\Sigma_{2}}\simeq\mathcal{T}. (3.17)

Combined with the Serre duality, this implies the equality

e1p,q=e1p,2q\displaystyle e_{1}^{p,q}=e_{1}^{-p,2-q} (3.18)

for p0p\neq 0.

It then follows that 0=E11,1=E11,10=E_{1}^{-1,1}=E_{1}^{1,1}, where the first equality is the consequence of the fact that G2G_{2} is torsion and G1G_{1} is torsion free, and the second equality is the case (p,q)=(1,1)(p,q)=(1,1) of (3.18).

Thus we have confirmed that the spectral sequence (3.16) is E1E_{1}-degenerate at (0,1)(0,1). Hence E10,1E0,1E1=0E_{1}^{0,1}\simeq E_{\infty}^{0,1}\simeq E^{1}=0, where the last vanishing is nothing but the rigidity of \mathcal{H}, which we confirmed in (1). Thus we have shown the rigidity of 𝒯\mathcal{T} and \mathcal{F}.

Again in the spectral sequence (3.16), the vanishing E1=0E^{1}=0 implies 0=E1,0=E21,00=E_{\infty}^{1,0}=E_{2}^{1,0}. Hence d10,0d_{1}^{0,0} is surjective. Also the vanishing E1=0E^{1}=0 implies E1,2=0E_{\infty}^{-1,2}=0. Since 0=E11,1=E21,10=E_{1}^{1,1}=E_{2}^{1,1}, the spectral sequence is E2E_{2}-degenerate at (1,2)(-1,2) and hence E1,2E21,2E_{\infty}^{-1,2}\simeq E_{2}^{-1,2}. This implies that d11,2d_{1}^{-1,2} is injective. Thus we obtain

hom(,)\displaystyle\hom(\mathcal{H},\mathcal{H}) =e11,1+kerd10,0\displaystyle=e_{1}^{1,-1}+\ker d_{1}^{0,0} =e11,1+e10,0e11,0\displaystyle=e_{1}^{1,-1}+e_{1}^{0,0}-e_{1}^{1,0} (3.19)
ext2(,)\displaystyle\operatorname{ext}^{2}(\mathcal{H},\mathcal{H}) =e11,3+cokerd11,2\displaystyle=e_{1}^{-1,3}+\operatorname{coker}d_{1}^{-1,2} =e11,3+e10,2e11,2.\displaystyle=e_{1}^{-1,3}+e_{1}^{0,2}-e_{1}^{-1,2}. (3.20)

Therefore substituting (3.19) and (3.20) into

(3.15)hom(,)=ext2(,)+1\displaystyle\eqref{eq:the difference is 1}\iff\hom(\mathcal{H},\mathcal{H})=\operatorname{ext}^{2}(\mathcal{H},\mathcal{H})+1 (3.21)

and using (3.18), we obtain

1=hom(,)ext2(,)=e10,0e10,2=(3.17)hom(,)ext2(,).1=\hom(\mathcal{H},\mathcal{H})-\operatorname{ext}^{2}(\mathcal{H},\mathcal{H})=e_{1}^{0,0}-e_{1}^{0,2}\stackrel{{\scriptstyle\eqref{eq:cT is stable under Serre functor up to shift}}}{{=}}\hom(\mathcal{F},\mathcal{F})-\operatorname{ext}^{2}(\mathcal{F},\mathcal{F}).

Since \mathcal{F} is rigid, the same proof as in [OU15, Lemma 2.2] shows that \mathcal{F} is an exceptional vector bundle. ∎

Definition 3.7.

For an exceptional object 𝐃(Σ2)\mathcal{E}\in\mathbf{D}(\Sigma_{2}), we will write i0()=i0i_{0}(\mathcal{E})=i_{0}\in\mathbb{Z} for the unique integer such that Suppi0()=Σ2\operatorname{Supp}\mathcal{H}^{i_{0}}(\mathcal{E})=\Sigma_{2}.

Let rank:K0(Σ2)\operatorname{rank}\colon\operatorname{K_{0}}\left(\Sigma_{2}\right)\to\mathbb{Z} be the rank function. If we let ι:Spec𝐤(Σ2)Σ2\iota\colon\operatorname{Spec}\mathbf{k}(\Sigma_{2})\to\Sigma_{2} denote the embedding of the generic point, rank\operatorname{rank} is concisely defined as the composition of the map K0(ι)\operatorname{K_{0}}\left(\iota^{\ast}\right) and the isomorphism K0(Spec𝐤(Σ2))\operatorname{K_{0}}\left(\operatorname{Spec}\mathbf{k}(\Sigma_{2})\right)\simeq\mathbb{Z} which sends the class of 𝐤(Σ2)\mathbf{k}(\Sigma_{2}) to 11. As a corollary of Lemma 3.3, we obtain the following

Corollary 3.8.

The equality

rank=(1)i0ranki0()\displaystyle\operatorname{rank}\mathcal{E}=(-1)^{i_{0}}\operatorname{rank}\mathcal{H}^{i_{0}}(\mathcal{E}) (3.22)

holds for any exceptional object 𝐃(Σ2)\mathcal{E}\in\mathbf{D}(\Sigma_{2}). In particular, rank0\operatorname{rank}\mathcal{E}\neq 0.

3.2. Properties of the schematic support of tors()\operatorname{tors}\mathcal{H}^{\bullet}(\mathcal{E})

The aim of this subsection is to prove Proposition 3.18, which asserts that the schematic support of tors()\operatorname{tors}\mathcal{H}^{\bullet}(\mathcal{E}) coincides with CC. This is the most technical part of this paper.

Remark 3.9.

In this paper, Proposition 3.18 will be used only as a black box. Hence one can first assume Proposition 3.18 to understand the rest of the paper and then later come back to its proof.

To describe the schematic support of the cohomology sheaves of \mathcal{E}, we consider the anti-canonical morphism f:Σ2P(1,1,2)f\colon\Sigma_{2}\to P(1,1,2) to the weighted projective plane of weight (1,1,2)(1,1,2), which contracts the (2)(-2)-curve CC to the singularity (0:0:1)(0:0:1).

Notation 3.10.

Let (R,𝔪)(R,\mathfrak{m}) denote the local ring of P(1,1,2)P(1,1,2) at the singular point. It is isomorphic to 𝐤[x,y,z](x,y,z)/(z2xy)\mathbf{k}[x,y,z]_{(x,y,z)}/(z^{2}-xy), the A1A_{1} singularity.

Our first goal is to give an RR-module structure on E2p,qE_{2}^{p,q} for (p,q)(0,0)(p,q)\neq(0,0) of the spectral sequence (3.4) with respect to which the differentials of the spectral sequence are RR-linear. The similar fact for spherical objects is used in [IU05], in which case the existence of such an RR-module structure is trivial. In fact, since the support of a spherical object is concentrated in CC, one can use the pushforward along f|SpecRf|_{\operatorname{Spec}R}.

Remark 3.11.

For any cohΣ2\mathcal{M}\in\operatorname{coh}\Sigma_{2} with (Supp)𝗋𝖾𝖽C\left(\operatorname{Supp}\mathcal{M}\right)_{\operatorname{\mathsf{red}}}\subset C, there is a natural homomorphism of 𝐤\mathbf{k}-algebras REndΣ2()R\to\operatorname{End}_{\Sigma_{2}}(\mathcal{M}) which kills 𝔪\mathfrak{m}^{\ell} for some >0\ell>0. Hence for ii\in\mathbb{Z} and 𝒩cohΣ2\mathcal{N}\in\operatorname{coh}\Sigma_{2}, there are natural RR-module structures of finite length on ExtΣ2i(,𝒩)Hom𝐃(Σ2)(,𝒩[i])\operatorname{Ext}_{\Sigma_{2}}^{i}(\mathcal{M},\mathcal{N})\simeq\operatorname{Hom}_{\mathbf{D}(\Sigma_{2})}(\mathcal{M},\mathcal{N}[i]) and ExtΣ2i(𝒩,)Hom𝐃(Σ2)(𝒩,[i])\operatorname{Ext}_{\Sigma_{2}}^{i}(\mathcal{N},\mathcal{M})\simeq\operatorname{Hom}_{\mathbf{D}(\Sigma_{2})}(\mathcal{N},\mathcal{M}[i]) given by precomposition and postcomposition, respectively. If 𝒩\mathcal{N} is also supported in CC, we may use REndΣ2(𝒩)R\to\operatorname{End}_{\Sigma_{2}}(\mathcal{N}) instead. However, we end up with the same RR-module structures defined via REndΣ2()R\to\operatorname{End}_{\Sigma_{2}}(\mathcal{M}).

For q0q\neq 0, the reduced support of either i()\mathcal{H}^{i}(\mathcal{E}) or i+q()\mathcal{H}^{i+q}(\mathcal{E}) is contained in CC and therefore E2p,qE_{2}^{p,q} has a canonical RR-module structure by Remark 3.11. For q=0q=0 and p0,2p\neq 0,2, E2p,0=0E_{2}^{p,0}=0.

For p=2p=2, we have

E22,0ii0ExtΣ22(i(),i())ExtΣ22(i0(),i0()).E_{2}^{2,0}\simeq\bigoplus_{i\neq i_{0}}\operatorname{Ext}_{\Sigma_{2}}^{2}(\mathcal{H}^{i}(\mathcal{E}),\mathcal{H}^{i}(\mathcal{E}))\oplus\operatorname{Ext}_{\Sigma_{2}}^{2}(\mathcal{H}^{i_{0}}(\mathcal{E}),\mathcal{H}^{i_{0}}(\mathcal{E})).

Since the reduced support of i()\mathcal{H}^{i}(\mathcal{E}) for ii0i\neq i_{0} is contained in CC, all the direct summands but the last one have canonical RR-module structures again by Remark 3.11.

Recall the following short exact sequence from the proof of Lemma 3.3 (3).

0𝒯torsi0()ιi0()i0()/𝒯0\displaystyle 0\to\mathcal{T}\coloneqq\operatorname{tors}\mathcal{H}^{i_{0}}(\mathcal{E})\stackrel{{\scriptstyle\iota}}{{\to}}\mathcal{H}^{i_{0}}(\mathcal{E})\to\mathcal{F}\coloneqq\mathcal{H}^{i_{0}}(\mathcal{E})/\mathcal{T}\to 0 (3.23)

Take HomX(,i0())\operatorname{Hom}_{X}(-,\mathcal{H}^{i_{0}}(\mathcal{E})) to obtain the following exact sequence

ExtΣ22(,i0())ExtΣ22(i0(),i0())ιExtΣ22(𝒯,i0())0,\operatorname{Ext}_{\Sigma_{2}}^{2}(\mathcal{F},\mathcal{H}^{i_{0}}(\mathcal{E}))\to\operatorname{Ext}_{\Sigma_{2}}^{2}(\mathcal{H}^{i_{0}}(\mathcal{E}),\mathcal{H}^{i_{0}}(\mathcal{E}))\stackrel{{\scriptstyle\iota^{\ast}}}{{\to}}\operatorname{Ext}_{\Sigma_{2}}^{2}(\mathcal{T},\mathcal{H}^{i_{0}}(\mathcal{E}))\to 0,

where

ExtΣ22(i0(),i0())ExtΣ22(𝒯,i0())xxι\displaystyle\begin{aligned} \operatorname{Ext}_{\Sigma_{2}}^{2}(\mathcal{H}^{i_{0}}(\mathcal{E})&,\mathcal{H}^{i_{0}}(\mathcal{E}))&&\to&\operatorname{Ext}_{\Sigma_{2}}^{2}(&\mathcal{T},\mathcal{H}^{i_{0}}(\mathcal{E}))\\ &x&&\mapsto&&x\circ\iota\end{aligned} (3.24)

is the precomposition (think of xx as a morphism i0()i0()[2]\mathcal{H}^{i_{0}}(\mathcal{E})\to\mathcal{H}^{i_{0}}(\mathcal{E})[2] in 𝐃(Σ2)\mathbf{D}(\Sigma_{2})). The following vanishing follows from the exceptionality of \mathcal{F}.

ExtΣ22(,i0())ExtΣ22(,)=0\operatorname{Ext}_{\Sigma_{2}}^{2}(\mathcal{F},\mathcal{H}^{i_{0}}(\mathcal{E}))\simeq\operatorname{Ext}_{\Sigma_{2}}^{2}(\mathcal{F},\mathcal{F})=0

Hence we conclude that the map ι\iota^{\ast} is an isomorphism of 𝐤\mathbf{k}-vector spaces. Since the reduced support of 𝒯\mathcal{T} is contained in CC, the right hand side of (3.24) has a canonical RR-module structure.

Definition 3.12.

We transfer the RR-module structure on ExtΣ22(𝒯,i0())\operatorname{Ext}_{\Sigma_{2}}^{2}(\mathcal{T},\mathcal{H}^{i_{0}}(\mathcal{E})) to ExtΣ22(i0(),i0())\operatorname{Ext}_{\Sigma_{2}}^{2}(\mathcal{H}^{i_{0}}(\mathcal{E}),\mathcal{H}^{i_{0}}(\mathcal{E})) via the isomorphism ι\iota^{\ast} of (3.24), thereby giving an RR-module structure on E22,0E_{2}^{2,0}.

Lemma 3.13.

For q0q\neq 0, the maps d20,q:E20,qE22,q1d_{2}^{0,q}\colon E_{2}^{0,q}\to E_{2}^{2,q-1} in the spectral sequence (3.4) are RR-linear.

Proof.

For q0,1q\neq 0,1, as already explained, the RR-module structures on the source and the target of d20,qd_{2}^{0,q} are naturally defined and hence the RR-linearity of d20,qd_{2}^{0,q} are rather obvious.

Let us show the RR-linearity of d20,1d_{2}^{0,1}. Take an arbitrary element

(ϕi)iE20,1=iHomΣ2(i(),i+1()).\displaystyle(\phi_{i})_{i}\in E_{2}^{0,1}=\bigoplus_{i}\operatorname{Hom}_{\Sigma_{2}}(\mathcal{H}^{i}(\mathcal{E}),\mathcal{H}^{i+1}(\mathcal{E})). (3.25)

It suffices to show that the maps as follows, which appear in the description of d20,1d_{2}^{0,1} given in Lemma 3.6, are RR-linear.

ϕi01\displaystyle\phi_{i_{0}-1} ϕi01ei0ExtΣ22(i0(),i0())\displaystyle\mapsto\phi_{i_{0}-1}\circ e^{i_{0}}\in\operatorname{Ext}_{\Sigma_{2}}^{2}(\mathcal{H}^{i_{0}}(\mathcal{E}),\mathcal{H}^{i_{0}}(\mathcal{E})) (3.26)
ϕi0\displaystyle\phi_{i_{0}} ei0+1ϕi0ExtΣ22(i0(),i0())\displaystyle\mapsto e^{i_{0}+1}\circ\phi_{i_{0}}\in\operatorname{Ext}_{\Sigma_{2}}^{2}(\mathcal{H}^{i_{0}}(\mathcal{E}),\mathcal{H}^{i_{0}}(\mathcal{E})) (3.27)

Under the isomorphism ι\iota^{\ast} of (3.24), the map (3.26) is identified with

ϕi01ϕi01ei0ιExtΣ22(𝒯,0()),\phi_{i_{0}-1}\mapsto\phi_{i_{0}-1}\circ e^{i_{0}}\circ\iota\in\operatorname{Ext}_{\Sigma_{2}}^{2}(\mathcal{T},\mathcal{H}^{0}(\mathcal{E})),

and, similarly, the map (3.27) is identified with

ϕi0ei0+1ϕi0ιExtΣ22(𝒯,i0()).\phi_{i_{0}}\mapsto e^{i_{0}+1}\circ\phi_{i_{0}}\circ\iota\in\operatorname{Ext}_{\Sigma_{2}}^{2}(\mathcal{T},\mathcal{H}^{i_{0}}(\mathcal{E})).

These maps are RR-linear, for the reason that the RR-module structures given in Remark 3.11 are by means of precompositions. As the RR-module structure on ExtΣ22(i0(),i0())\operatorname{Ext}_{\Sigma_{2}}^{2}(\mathcal{H}^{i_{0}}(\mathcal{E}),\mathcal{H}^{i_{0}}(\mathcal{E})) is given in Definition 3.12 by transferring the RR-module structure on ExtΣ22(𝒯,i0())\operatorname{Ext}_{\Sigma_{2}}^{2}(\mathcal{T},\mathcal{H}^{i_{0}}(\mathcal{E})) via ι\iota^{\ast}, this is exactly what we had to prove. ∎

Suppose \mathcal{M} is a pure coherent sheaf on Σ2\Sigma_{2} with (Supp)𝗋𝖾𝖽=C\left(\operatorname{Supp}\mathcal{M}\right)_{\operatorname{\mathsf{red}}}=C. Then by Remark 3.11, MH0(Σ2,)HomΣ2(𝒪Σ2,)M\coloneqq H^{0}(\Sigma_{2},\mathcal{M})\simeq\operatorname{Hom}_{\Sigma_{2}}(\mathcal{O}_{\Sigma_{2}},\mathcal{M}) has a standard RR-module structure of finite length and hence there is an integer \ell such that 𝔪M=0\mathfrak{\mathfrak{m}}^{\ell}M=0. In general, for an RR-module MM and an ideal IRI\subset R, we let (0:I)M(0:I)_{M} denote the annihilator

(0:I)M{xMIx=0}M.\displaystyle(0:I)_{M}\coloneqq\{x\in M\mid Ix=0\}\subseteq M. (3.28)

Geometrically speaking, this is the maximum submodule of MM which is “supported on the closed subscheme SpecR/I\operatorname{Spec}R/I”.

Lemma 3.14.

For a pure coherent sheaf \mathcal{M} on Σ2\Sigma_{2} with (Supp)𝗋𝖾𝖽=C\left(\operatorname{Supp}\mathcal{M}\right)_{\operatorname{\mathsf{red}}}=C, assume

nmin{𝔪M=0}>1,n\coloneqq\min\{\ell\mid\mathfrak{\mathfrak{m}}^{\ell}M=0\}>1,

where M=H0(Σ2,)modRM=H^{0}(\Sigma_{2},\mathcal{M})\in\operatorname{mod}R as above. Then the following strict inequality holds.

dim𝐤M(0:𝔪n1)M<dim𝐤𝔪n1M\displaystyle\dim_{\mathbf{k}}\frac{M}{(0:\mathfrak{m}^{n-1})_{M}}<\dim_{\mathbf{k}}\mathfrak{m}^{n-1}M (3.29)
Proof.

Without loss of generality, we may assume that ffηf^{\ast}f_{\ast}\mathcal{M}\stackrel{{\scriptstyle\eta}}{{\to}}\mathcal{M} is surjective; i.e., \mathcal{M} is ff-globally generated. In fact, if η\eta is not surjective, we can replace \mathcal{M} with the image Imη\operatorname{Im}\eta. By standard arguments on the adjoint pair fff^{\ast}\dashv f_{\ast}, there is a canonical isomorphism fImηf=Mf_{\ast}\operatorname{Im}\eta\xrightarrow{\sim}f_{\ast}\mathcal{M}=M and Imη\operatorname{Im}\eta is automatically ff-globally generated.

The following fact about the A1A_{1}-singularity RR is known well: for each i>0i>0, the ii-th powers of the ideal sheaf

C=𝒪Σ2(C)𝒪Σ2\displaystyle\mathcal{I}_{C}=\mathcal{O}_{\Sigma_{2}}(-C)\subset\mathcal{O}_{\Sigma_{2}} (3.30)

of CC and 𝔪\mathfrak{m} are related to each other as follows.

𝔪i𝒪Σ2=𝒪Σ2(iC)=Ci𝒪Σ2(f|SpecR)Ci=𝔪iR\displaystyle\begin{aligned} \mathfrak{m}^{i}\mathcal{O}_{\Sigma_{2}}=\mathcal{O}_{\Sigma_{2}}(-iC)=\mathcal{I}_{C}^{i}\subseteq\mathcal{O}_{\Sigma_{2}}\\ \left(f|_{\operatorname{Spec}R}\right)_{\ast}\mathcal{I}_{C}^{i}=\mathfrak{m}^{i}\subseteq R\end{aligned} (3.31)

The vanishing 𝔪nM=0\mathfrak{m}^{n}M=0 means that M=fM=f_{\ast}\mathcal{M} is a sheaf on Spec(R/𝔪n)\operatorname{Spec}(R/\mathfrak{m}^{n}), from which we deduce that fff^{*}f_{*}\mathcal{M} and hence its quotient \mathcal{M} are supported on f1(Spec(R/𝔪n))=(3.31)nCf^{-1}(\operatorname{Spec}(R/\mathfrak{m}^{n}))\stackrel{{\scriptstyle\eqref{eq:powers of cO ( - C ) vs frakm}}}{{=}}nC. This means Cn=0\mathcal{I}_{C}^{n}\mathcal{M}=0.

On the other hand, note that the image of the canonical morphism

H0(Σ2,Cn1)𝐤H0(Σ2,)H0(Σ2,Cn1),\displaystyle H^{0}\left(\Sigma_{2},\mathcal{I}_{C}^{n-1}\right)\otimes_{\mathbf{k}}H^{0}(\Sigma_{2},\mathcal{M})\to H^{0}\left(\Sigma_{2},\mathcal{I}_{C}^{n-1}\mathcal{M}\right), (3.32)

as a submodule of H0(Σ2,)=MH^{0}(\Sigma_{2},\mathcal{M})=M, is 𝔪n1M\mathfrak{m}^{n-1}M by (3.31). Thus we see H0(Σ2,Cn1)𝔪n1M0H^{0}\left(\Sigma_{2},\mathcal{I}_{C}^{n-1}\mathcal{M}\right)\supset\mathfrak{m}^{n-1}M\neq 0, hence Cn10\mathcal{I}_{C}^{n-1}\mathcal{M}\neq 0. Therefore we conclude

n=min{C=0}.n=\min\{\ell\mid\mathcal{I}_{C}^{\ell}\mathcal{M}=0\}.

Since C(Cn1)=0\mathcal{I}_{C}(\mathcal{I}_{C}^{n-1}\mathcal{M})=0, we can naturally think of Cn1\mathcal{I}_{C}^{n-1}\mathcal{M} as an object of cohC\operatorname{coh}C. Since we assumed that \mathcal{M} is pure, so is its subsheaf Cn1\mathcal{I}_{C}^{n-1}\mathcal{M}. Hence Cn1\mathcal{I}_{C}^{n-1}\mathcal{M} is a vector bundle on CC of positive rank.

Likewise (0:Cn1)\frac{\mathcal{M}}{(0:\mathcal{I}_{C}^{n-1})_{\mathcal{M}}} can be thought of an object of cohC\operatorname{coh}C, where (0:Cn1)(0:\mathcal{I}_{C}^{n-1})_{\mathcal{M}} is the sheaf version of the annihilator (3.28); i.e., the maximum subsheaf of \mathcal{M} whose schematic support is contained in (n1)C(n-1)C. We claim that there is an isomorphism

φ:𝒪C((n1)C)𝒪C(0:Cn1)Cn1\displaystyle\varphi\colon\mathcal{O}_{C}(-(n-1)C)\otimes_{\mathcal{O}_{C}}\frac{\mathcal{M}}{(0:\mathcal{I}_{C}^{n-1})_{\mathcal{M}}}\xrightarrow{\sim}\mathcal{I}_{C}^{n-1}\mathcal{M} (3.33)

of vector bundles on CC. Indeed φ\varphi is induced from the product morphism Cn1𝒪Σ2Cn1\mathcal{I}_{C}^{n-1}\otimes_{\mathcal{O}_{\Sigma_{2}}}\mathcal{M}\to\mathcal{I}_{C}^{n-1}\mathcal{M}, whose surjectivity implies that of φ\varphi. At the generic point ξ\xi of CC, the stalk (C)ξ(\mathcal{I}_{C})_{\xi} is the maximal ideal of the discrete valuation ring 𝒪Σ2,ξ\mathcal{O}_{\Sigma_{2},\xi} and ξ\mathcal{M}_{\xi} is a finite length 𝒪Σ2,ξ/(Cn)ξ\mathcal{O}_{\Sigma_{2},\xi}/(\mathcal{I}_{C}^{n})_{\xi}-module. Hence one sees that φ\varphi is an isomorphism at ξ\xi (use the structure theorem for finitely generated modules over a discrete valuation ring). Therefore it is enough to show that the left hand side of (3.33) is pure; i.e., there is no zero-dimensional subsheaf.

Assume for a contradiction that /(0:Cn1)\mathcal{M}/(0:\mathcal{I}_{C}^{n-1})_{\mathcal{M}} is not pure. Then there is a subsheaf (0:Cn1)𝒮(0:\mathcal{I}_{C}^{n-1})_{\mathcal{M}}\subsetneq\mathcal{S}\subset\mathcal{M} such that Supp(𝒮/(0:Cn1))\operatorname{Supp}\left(\mathcal{S}/(0:\mathcal{I}_{C}^{n-1})_{\mathcal{M}}\right) is zero-dimensional. Then Cn1𝒮\mathcal{I}_{C}^{n-1}\mathcal{S}\subset\mathcal{M} is non-zero and zero-dimensional, which contradicts the purity of \mathcal{M}. To see dimCn1𝒮=0\dim\mathcal{I}_{C}^{n-1}\mathcal{S}=0, note that there is an epimorphism

Cn1𝒪Σ2𝒮/(0:Cn1)Cn1𝒮\displaystyle\mathcal{I}_{C}^{n-1}\otimes_{\mathcal{O}_{\Sigma_{2}}}\mathcal{S}/(0:\mathcal{I}_{C}^{n-1})_{\mathcal{M}}\twoheadrightarrow\mathcal{I}_{C}^{n-1}\mathcal{S} (3.34)

and that dim𝒮/(0:Cn1)=0\dim\mathcal{S}/(0:\mathcal{I}_{C}^{n-1})_{\mathcal{M}}=0.

Now consider the following RR-module.

VMH0(Σ2,(0:Cn1))=M(0:𝔪n1)MV\coloneqq\frac{M}{H^{0}\left(\Sigma_{2},(0:\mathcal{I}_{C}^{n-1})_{\mathcal{M}}\right)}=\frac{M}{(0:\mathfrak{m}^{n-1})_{M}}

To see the equality, note that the denominator of the left hand side, as a submodule of M=H0(Σ2,)M=H^{0}(\Sigma_{2},\mathcal{M}), coincides with the following subset.

{sMsP(0:C,Pn1)PPC}.\{s\in M\mid s_{P}\in(0:\mathcal{I}_{C,P}^{n-1})_{\mathcal{M}_{P}}\quad\forall P\in C\}.

Since C,Pn1=𝔪n1𝒪Σ2,P\mathcal{I}_{C,P}^{n-1}=\mathfrak{m}^{n-1}\mathcal{O}_{\Sigma_{2},P}, this coincides with the denominator of the right hand side.

If we put L𝒪C((n1)C)L\coloneqq\mathcal{O}_{C}(-(n-1)C) and E/(0:Cn1)E\coloneqq\mathcal{M}/(0:\mathcal{I}_{C}^{n-1})_{\mathcal{M}}, then (3.33) is rewritten as

L𝒪CECn1.L\otimes_{\mathcal{O}_{C}}E\simeq\mathcal{I}_{C}^{n-1}\mathcal{M}.

Noting H0(C,L)𝔪n1/𝔪nH^{0}(C,L)\simeq\mathfrak{m}^{n-1}/\mathfrak{m}^{n}, we see that 𝔪n1M\mathfrak{m}^{n-1}M is the image of the product map

ψ:H0(C,L)𝐤VH0(C,Cn1).\psi\colon H^{0}(C,L)\otimes_{\mathbf{k}}V\to H^{0}(C,\mathcal{I}_{C}^{n-1}\mathcal{M}).

Then the assertion follows from Lemma 3.15 below. ∎

Lemma 3.15.

Let EE and LL be a vector bundle and a line bundle on a smooth projective curve CC, respectively. Suppose dimH0(C,L)>1\dim H^{0}(C,L)>1 and let VH0(C,E)V\subset H^{0}(C,E) be a non-zero linear subspace. Then we have a strict inequality

dimV<rankψ,\dim V<\operatorname{rank}\psi,

where ψ:H0(C,L)𝐤VH0(C,L𝒪CE)\psi\colon H^{0}(C,L)\otimes_{\mathbf{k}}V\to H^{0}(C,L\otimes_{\mathcal{O}_{C}}E) denotes the product map.

Proof.

Without loss of generality, we may replace EE with the subsheaf generated by VV. Take a pair of linearly independent global sections s,tH0(C,L)s,t\in H^{0}(C,L). It follows that 𝒪Ct𝒪Cs\mathcal{O}_{C}\cdot t\not\subseteq\mathcal{O}_{C}\cdot s as subsheaves of LL. Since EE is locally free, this implies that (𝒪Ct)𝒪CE(𝒪Cs)𝒪CE(\mathcal{O}_{C}\cdot t)\otimes_{\mathcal{O}_{C}}E\not\subseteq(\mathcal{O}_{C}\cdot s)\otimes_{\mathcal{O}_{C}}E as subsheaves of L𝒪CEL\otimes_{\mathcal{O}_{C}}E. Since (𝒪Ct)𝒪CE(\mathcal{O}_{C}\cdot t)\otimes_{\mathcal{O}_{C}}E and (𝒪Cs)𝒪CE(\mathcal{O}_{C}\cdot s)\otimes_{\mathcal{O}_{C}}E are the subsheaves of L𝒪CEL\otimes_{\mathcal{O}_{C}}E generated by ψ(𝐤tV)\psi(\mathbf{k}t\otimes V) and ψ(𝐤sV)\psi(\mathbf{k}s\otimes V) respectively, we conclude ψ(𝐤tV)ψ(𝐤sV)\psi(\mathbf{k}t\otimes V)\not\subseteq\psi(\mathbf{k}s\otimes V). Thus we see

Vψ(𝐤sV)ψ(𝐤sV)+ψ(𝐤tV)Imψ.\displaystyle V\simeq\psi(\mathbf{k}s\otimes V)\subsetneq\psi(\mathbf{k}s\otimes V)+\psi(\mathbf{k}t\otimes V)\subseteq\operatorname{Im}\psi. (3.35)

Taking dim𝐤\dim_{\mathbf{k}}, we obtain the assertion. ∎

Let mod0RmodR\operatorname{mod}_{0}R\subset\operatorname{mod}R be the category of Artinian (dim𝐤<\iff\dim_{\mathbf{k}}<\infty) RR-modules. There is an (anti-)involution of categories

D:(mod0R)opmod0RD\colon\left(\operatorname{mod}_{0}R\right)^{\mathrm{op}}\xrightarrow{\sim}\operatorname{mod}_{0}R

which is defined as follows.

DMHom𝐤(M,𝐤)\displaystyle DM\coloneqq\operatorname{Hom}_{\mathbf{k}}(M,\mathbf{k}) (3.36)
Corollary 3.16.

Under the same notation and assumption as in Lemma 3.14, put WDMW\coloneqq DM. Then the following strict inequality holds.

dim𝐤W(0:𝔪n1)W>dim𝐤𝔪n1W\dim_{\mathbf{k}}\frac{W}{(0:\mathfrak{m}^{n-1})_{W}}>\dim_{\mathbf{k}}\mathfrak{m}^{n-1}W
Proof.

To obtain the assertion, apply the anti-involution DD to both sides of the equality (3.29) (before applying dim𝐤\dim_{\mathbf{k}}) and then use the following Lemma 3.17. Note that DD preserves dimensions. ∎

Lemma 3.17.

For Mmod0RM\in\operatorname{mod}_{0}R and integer 0\ell\geq 0, there are isomorphisms as follows.

D(𝔪M)\displaystyle D(\mathfrak{m}^{\ell}M) DM/(0:𝔪)DM\displaystyle\simeq DM/(0:\mathfrak{m}^{\ell})_{DM} (3.37)
D(M/(0:𝔪)M)\displaystyle D\left(M/(0:\mathfrak{m}^{\ell})_{M}\right) 𝔪DM\displaystyle\simeq\mathfrak{m}^{\ell}DM (3.38)
Proof.

To obtain the second isomorphism, replace MM with DMDM in the first isomorphism and then use D2idD^{2}\simeq\operatorname{id}. To see the first isomorphism, consider the following short exact sequence. The inclusion ii is the canonical one.

0(0:𝔪)DMiDMCcokeri0\displaystyle 0\to(0:\mathfrak{m}^{\ell})_{DM}\stackrel{{\scriptstyle i}}{{\to}}DM\to C\coloneqq\operatorname{coker}i\to 0 (3.39)

By applying the anti-involution DD to this, we obtain the following short exact sequence.

0DCMD(0:𝔪)DM0\displaystyle 0\to DC\to M\to D(0:\mathfrak{m}^{\ell})_{DM}\to 0 (3.40)

Then, as a submodule of MM, DCDC is computed as follows.

DC={xMy(0:𝔪)DMi(y)(x)=0}={xMyDM,𝔪y=0y(x)=0}.\displaystyle DC=\{x\in M\mid y\in(0:\mathfrak{m}^{\ell})_{DM}\Rightarrow i(y)(x)=0\}=\{x\in M\mid y\in DM,\mathfrak{m}^{\ell}y=0\Rightarrow y(x)=0\}. (3.41)

Note that yDMy\in DM satisfies 𝔪y=0\mathfrak{m}^{\ell}y=0 if and only if y(𝔪M)=0y(\mathfrak{m}^{\ell}M)=0; i.e., yD(M/𝔪M)qDMy\in D(M/\mathfrak{m}^{\ell}M)\stackrel{{\scriptstyle q^{\ast}}}{{\hookrightarrow}}DM, where q:MM/𝔪Mq\colon M\twoheadrightarrow M/\mathfrak{m}^{\ell}M is the quotient map. Hence

DC={xMyD(M/𝔪M)y(x)=0}=D(M/𝔪M)=𝔪M,\displaystyle DC=\{x\in M\mid y\in D(M/\mathfrak{m}^{\ell}M)\Rightarrow y(x)=0\}=D(M/\mathfrak{m}^{\ell}M)^{\perp}=\mathfrak{m}^{\ell}M, (3.43)

so that CD(𝔪M)C\simeq D(\mathfrak{m}^{\ell}M). ∎

Now we are ready to prove the following

Proposition 3.18.

For an exceptional object 𝐃(Σ2)\mathcal{E}\in\mathbf{D}(\Sigma_{2}), it holds that

Supptorsi0()=Suppi()(ii0)=C or .\operatorname{Supp}\operatorname{tors}\mathcal{H}^{i_{0}}(\mathcal{E})=\operatorname{Supp}\mathcal{H}^{i}(\mathcal{E})\ (i\neq i_{0})=C\text{ or }\emptyset.
Proof.

We first discuss Suppi()\operatorname{Supp}\mathcal{H}^{i}(\mathcal{E}) for ii0i\neq i_{0}. We may and will assume ii0i()0\bigoplus_{i\neq i_{0}}\mathcal{H}^{i}(\mathcal{E})\neq 0, since otherwise there is nothing to prove. Put

nminii0{Ci()=0}1.\displaystyle n\coloneqq\min\bigcap_{i\neq i_{0}}\left\{\ell\mid\mathcal{I}_{C}^{\ell}\mathcal{H}^{i}(\mathcal{E})=0\right\}\geq 1. (3.44)

In the spectral sequence (3.4) for the exceptional object \mathcal{E}, we have the isomorphism

d20,1:E20,1=iHomΣ2(i(),i+1())E22,0=iExtΣ22(i(),i())\displaystyle d_{2}^{0,1}\colon E_{2}^{0,1}=\bigoplus_{i}\operatorname{Hom}_{\Sigma_{2}}(\mathcal{H}^{i}(\mathcal{E}),\mathcal{H}^{i+1}(\mathcal{E}))\xrightarrow{\sim}E_{2}^{2,0}=\bigoplus_{i}\operatorname{Ext}_{\Sigma_{2}}^{2}(\mathcal{H}^{i}(\mathcal{E}),\mathcal{H}^{i}(\mathcal{E})) (3.45)

of RR-modules by Lemma 3.13. Note that 𝔪nE20,1=0\mathfrak{m}^{n}E_{2}^{0,1}=0, since for each ii either Cni()=0\mathcal{I}_{C}^{n}\mathcal{H}^{i}(\mathcal{E})=0 or Cni+1()=0\mathcal{I}_{C}^{n}\mathcal{H}^{i+1}(\mathcal{E})=0 holds.

On the other hand, for each ii0i\neq i_{0} there is an isomorphism of RR-modules given by the Serre duality:

ExtΣ22(i(),i())DHom(i(),i())\operatorname{Ext}^{2}_{\Sigma_{2}}(\mathcal{H}^{i}(\mathcal{E}),\mathcal{H}^{i}(\mathcal{E}))\simeq D\operatorname{Hom}(\mathcal{H}^{i}(\mathcal{E}),\mathcal{H}^{i}(\mathcal{E}))

This implies that 𝔪n1E22,00\mathfrak{m}^{n-1}E_{2}^{2,0}\neq 0. Combining this with the isomorphism of RR-modules (3.45), we obtain

n=min{𝔪E20,1=0}=min{𝔪E22,0=0}.n=\min\{\ell\mid\mathfrak{m}^{\ell}E_{2}^{0,1}=0\}=\min\{\ell\mid\mathfrak{m}^{\ell}E_{2}^{2,0}=0\}.

Now assume for a contradiction that n>1n>1. Let

iomΣ2(i(),i+1()),\displaystyle\mathcal{M}\coloneqq\bigoplus_{i}\mathop{{\mathcal{H}}om}\nolimits_{\Sigma_{2}}(\mathcal{H}^{i}(\mathcal{E}),\mathcal{H}^{i+1}(\mathcal{E})), (3.46)

so that E20,1fE_{2}^{0,1}\simeq f_{\ast}\mathcal{M}. We can apply Lemma 3.14 to \mathcal{M}, to obtain the strict inequality

dim𝐤E20,1(0:𝔪n1)E20,1<dim𝐤𝔪n1E20,1.\displaystyle\dim_{\mathbf{k}}\frac{E_{2}^{0,1}}{(0:\mathfrak{m}^{n-1})_{E_{2}^{0,1}}}<\dim_{\mathbf{k}}\mathfrak{m}^{n-1}E_{2}^{0,1}. (3.47)

On the other hand, let

ii0ndΣ2(i())omΣ2(i0(),𝒯).\displaystyle\mathcal{M}^{\prime}\coloneqq\bigoplus_{i\neq i_{0}}\mathop{{\mathcal{E}}nd}\nolimits_{\Sigma_{2}}(\mathcal{H}^{i}(\mathcal{E}))\oplus\mathop{{\mathcal{H}}om}\nolimits_{\Sigma_{2}}(\mathcal{H}^{i_{0}}(\mathcal{E}),\mathcal{T}). (3.48)

By Lemma 3.13 and the Serre duality, it follows that there is an isomorphism of RR-modules E22,0DfE_{2}^{2,0}\simeq Df_{\ast}\mathcal{M}^{\prime}. We can apply Corollary 3.16 to \mathcal{M}^{\prime}, to obtain the strict inequality

dim𝐤E22,0(0:𝔪n1)E22,0>dim𝐤𝔪n1E22,0.\displaystyle\dim_{\mathbf{k}}\frac{E_{2}^{2,0}}{(0:\mathfrak{m}^{n-1})_{E_{2}^{2,0}}}>\dim_{\mathbf{k}}\mathfrak{m}^{n-1}E_{2}^{2,0}. (3.49)

The strict inequalities (3.47) and (3.49) contradict the isomorphism (3.45). Hence we obtain n=1n=1, which means that i()\mathcal{H}^{i}(\mathcal{E}) is an 𝒪C\mathcal{O}_{C}-module for any ii0i\neq i_{0}. In fact, by rigidity, i()\mathcal{H}^{i}(\mathcal{E}) is a vector bundle on CC for any ii0i\neq i_{0}.

Finally, to investigate torsi0()\operatorname{tors}\mathcal{H}^{i_{0}}(\mathcal{E}), we consider the derived dual \mathcal{E}^{\vee}. As we show in Lemma 3.25 below, there is an isomorphism as follows.

i0+1()(torsi0())[1]\mathcal{H}^{-i_{0}+1}(\mathcal{E}^{\vee})\simeq(\operatorname{tors}\mathcal{H}^{i_{0}}(\mathcal{E}))^{\vee}[1]

Since i0+1i0=Lemma 3.25i0()-i_{0}+1\neq-i_{0}\stackrel{{\scriptstyle\text{Lemma~{}\ref{lm:dual exceptional object}}}}{{=}}i_{0}(\mathcal{E}^{\vee}), by applying what we have just proved to the exceptional object \mathcal{E}^{\vee}, we see that the left hand side, hence the right hand side, is a vector bundle on CC. Hence so is torsi0()\operatorname{tors}\mathcal{H}^{i_{0}}(\mathcal{E}). ∎

Remark 3.19.

Proposition 3.18 will not be used in the proof of Lemma 3.25, so that it is harmless to use Lemma 3.25 in the proof of Proposition 3.18 (as we did in the last paragraph). Actually, in the proof of Lemma 3.25, we only use Lemma 3.3 and some standard facts on homological algebra.

3.3. More on the structure of ()\mathcal{H}^{\bullet}(\mathcal{E})

In this subsection we give a structure theorem for i0()\mathcal{H}^{i_{0}}(\mathcal{E}) in Lemma 3.21. It is then used to give a structure theorem for tors()\operatorname{tors}\mathcal{H}^{\bullet}(\mathcal{E}) in Corollary 3.23.

Below is repeatedly used in this paper.

Lemma 3.20 (==[Kul97, Remark 2.3.4]).

Let 𝐃(Σ2)\mathcal{E}\in\mathbf{D}(\Sigma_{2}) be an exceptional vector bundle of rank=r\operatorname{rank}\mathcal{E}=r. Then there is an isomorphism

|C𝒪C(b)s𝒪C(b+1)rs\displaystyle\mathcal{E}|_{C}\simeq\mathcal{O}_{C}(b)^{\oplus s}\oplus\mathcal{O}_{C}(b+1)^{\oplus r-s} (3.50)

for some bb\in\mathbb{Z} and ss\in\mathbb{N} such that 1sr1\leq s\leq r.

Note that the integers bb and ss in Lemma 3.20 are uniquely determined by \mathcal{E}.

Lemma 3.21.

Let 𝐃(Σ2)\mathcal{E}\in\mathbf{D}(\Sigma_{2}) be an exceptional object. Then the unique non-torsion cohomology sheaf i0()\mathcal{H}^{i_{0}}(\mathcal{E}) decomposes as

i0()ET,\mathcal{H}^{i_{0}}(\mathcal{E})\simeq E\oplus T,

where EE is an exceptional sheaf and TT is a vector bundle on CC. Moreover, if EE is not locally free, then there is an integer aa\in\mathbb{Z} such that

  • The torsion part torsE\operatorname{tors}E is a direct sum of copies of 𝒪C(a)\mathcal{O}_{C}(a).

  • TT is a direct sum of copies of 𝒪C(a)\mathcal{O}_{C}(a) and 𝒪C(a+1)\mathcal{O}_{C}(a+1).

Definition 3.22.

For an exceptional object 𝐃(Σ2)\mathcal{E}\in\mathbf{D}(\Sigma_{2}), let E=E()E=E(\mathcal{E}) denote the exceptional sheaf EE of Lemma 3.21. Also let =()\mathcal{F}=\mathcal{F}(\mathcal{E}) denote the torsion free part of the sheaf i0()\mathcal{H}^{i_{0}}(\mathcal{E}), which is known to be an exceptional vector bundle by Lemma 3.3 (3).

Proof of Lemma 3.21.

Consider the following standard short exact sequence.

0𝒯torsi0()i0()i0()/𝒯0\displaystyle 0\to\mathcal{T}\coloneqq\operatorname{tors}\mathcal{H}^{i_{0}}(\mathcal{E})\to\mathcal{H}^{i_{0}}(\mathcal{E})\to\mathcal{F}\coloneqq\mathcal{H}^{i_{0}}(\mathcal{E})/\mathcal{T}\to 0 (3.51)

Lemma 3.3 (3) asserts that \mathcal{F} is an exceptional vector bundle. Hence we assume that 𝒯0\mathcal{T}\neq 0, since otherwise there is nothing to prove.

𝒯\mathcal{T} is a rigid sheaf again by Lemma 3.3 (3). Combined with Proposition 3.18, this implies that there are a,s>0,t0a\in\mathbb{Z},s>0,t\geq 0 such that

𝒯𝒪C(a)s𝒪C(a+1)t.\displaystyle\mathcal{T}\simeq\mathcal{O}_{C}(a)^{\oplus s}\oplus\mathcal{O}_{C}(a+1)^{\oplus t}. (3.52)

By Lemma 3.20, |C\mathcal{F}|_{C} is also rigid and hence there are b,s>0,t0b\in\mathbb{Z},s^{\prime}>0,t^{\prime}\geq 0 such that

|C𝒪C(b)s𝒪C(b+1)t.\mathcal{F}|_{C}\simeq\mathcal{O}_{C}(b)^{\oplus s^{\prime}}\oplus\mathcal{O}_{C}(b+1)^{\oplus t^{\prime}}.

If ExtΣ21(,𝒯)=0\operatorname{Ext}_{\Sigma_{2}}^{1}(\mathcal{F},\mathcal{T})=0, then the short exact sequence (3.51) splits and we are done. Therefore we assume ExtΣ21(,𝒯)0\operatorname{Ext}_{\Sigma_{2}}^{1}(\mathcal{F},\mathcal{T})\neq 0, which implies

ab1\displaystyle a\leq b-1 (3.53)

In order to make the argument conceptual, fix a 𝐤\mathbf{k}-vector space VV of dimension ss and replace 𝒪C(a)s\mathcal{O}_{C}(a)^{\oplus s} with V𝐤𝒪C(a)V\otimes_{\mathbf{k}}\mathcal{O}_{C}(a). Let

i0()/𝒪C(a+1)t\displaystyle\mathcal{H}^{\prime}\coloneqq\mathcal{H}^{i_{0}}(\mathcal{E})/\mathcal{O}_{C}(a+1)^{\oplus t} (3.54)

be the quotient by the subsheaf 𝒪C(a+1)t𝒯i0()\mathcal{O}_{C}(a+1)^{\oplus t}\subset\mathcal{T}\subset\mathcal{H}^{i_{0}}(\mathcal{E}), which fits in the following short exact sequence.

0V𝐤𝒪C(a)0\displaystyle 0\to V\otimes_{\mathbf{k}}\mathcal{O}_{C}(a)\to\mathcal{H}^{\prime}\to\mathcal{F}\to 0 (3.55)

From this we see

HomΣ2(𝒪(a+1)t,)=0,\operatorname{Hom}_{\Sigma_{2}}(\mathcal{O}(a+1)^{\oplus t},\mathcal{H}^{\prime})=0,

which implies that \mathcal{H}^{\prime} is rigid by Mukai’s lemma [Kul97, Lemma 2.1.4. 2.(a)], (which is obtained from the spectral sequence of the form (3.16)).

Let

[f:ExtΣ21(,𝒪C(a))V]Hom𝐤(ExtΣ21(,𝒪C(a)),V)ExtΣ21(,V𝐤𝒪C(a))\displaystyle[f\colon\operatorname{Ext}_{\Sigma_{2}}^{1}(\mathcal{F},\mathcal{O}_{C}(a))^{\vee}\to V]\in\operatorname{Hom}_{\mathbf{k}}(\operatorname{Ext}_{\Sigma_{2}}^{1}(\mathcal{F},\mathcal{O}_{C}(a))^{\vee},V)\simeq\operatorname{Ext}_{\Sigma_{2}}^{1}(\mathcal{F},V\otimes_{\mathbf{k}}\mathcal{O}_{C}(a)) (3.56)

correspond to the extension (3.55). We will show that ff is injective. In the long exact sequence obtained by applying HomΣ2(,)\operatorname{Hom}_{\Sigma_{2}}(\mathcal{H}^{\prime},-) to (3.55),

  • the map HomΣ2(,)HomΣ2(,)\operatorname{Hom}_{\Sigma_{2}}(\mathcal{H}^{\prime},\mathcal{H}^{\prime})\to\operatorname{Hom}_{\Sigma_{2}}(\mathcal{H}^{\prime},\mathcal{F}) is surjective since HomΣ2(,)HomΣ2(,)=𝐤id\operatorname{Hom}_{\Sigma_{2}}(\mathcal{H}^{\prime},\mathcal{F})\cong\operatorname{Hom}_{\Sigma_{2}}(\mathcal{F},\mathcal{F})=\mathbf{k}\cdot\operatorname{id}_{\mathcal{F}} by the exceptionality of \mathcal{F} and

  • Ext1(,)=0\operatorname{Ext}^{1}(\mathcal{H}^{\prime},\mathcal{H}^{\prime})=0 by the rigidity of \mathcal{H}^{\prime}.

Hence we obtain

ExtΣ21(,𝒪C(a))=0.\operatorname{Ext}^{1}_{\Sigma_{2}}(\mathcal{H}^{\prime},\mathcal{O}_{C}(a))=0.

Next we apply HomΣ2(,𝒪C(a))\operatorname{Hom}_{\Sigma_{2}}(-,\mathcal{O}_{C}(a)) to (3.55) to obtain a surjective map

VHomΣ2(V𝒪C(a),𝒪C(a))ExtΣ21(,𝒪C(a)).V^{\vee}\cong\operatorname{Hom}_{\Sigma_{2}}(V\otimes\mathcal{O}_{C}(a),\mathcal{O}_{C}(a))\to\operatorname{Ext}^{1}_{\Sigma_{2}}(\mathcal{F},\mathcal{O}_{C}(a)). (3.57)

For any φV\varphi\in V^{\vee}, the map (3.57) sends φ\varphi to (φid𝒪C(a))e(\varphi\otimes\operatorname{id}_{\mathcal{O}_{C}(a)})\circ e, where eExtΣ21(,V𝐤𝒪C(a))e\in\operatorname{Ext}_{\Sigma_{2}}^{1}(\mathcal{F},V\otimes_{\mathbf{k}}\mathcal{O}_{C}(a)) denotes the extension class (3.55). Take a basis {v1,,vs}\{v_{1},\dots,v_{s}\} of VV and decompose ee as ieivi\sum_{i}e_{i}\otimes v_{i} under the isomorphism ExtΣ21(,V𝐤𝒪C(a))ExtΣ21(,𝒪C(a))𝐤V\operatorname{Ext}_{\Sigma_{2}}^{1}(\mathcal{F},V\otimes_{\mathbf{k}}\mathcal{O}_{C}(a))\simeq\operatorname{Ext}_{\Sigma_{2}}^{1}\left(\mathcal{F},\mathcal{O}_{C}(a)\right)\otimes_{\mathbf{k}}V. Then one can confirm that (3.57) sends φ\varphi to iφ(vi)ei\sum_{i}\varphi(v_{i})e_{i} and that ff sends a linear form ξExtΣ21(,𝒪C(a))\xi\in\operatorname{Ext}_{\Sigma_{2}}^{1}(\mathcal{F},\mathcal{O}_{C}(a))^{\vee} to iξ(ei)vi\sum_{i}\xi(e_{i})v_{i}. Hence the surjectivity of the map (3.57) implies that e1,,ese_{1},\dots,e_{s} generates ExtΣ21(,𝒪C(a))\operatorname{Ext}_{\Sigma_{2}}^{1}\left(\mathcal{F},\mathcal{O}_{C}(a)\right), which in turn is equivalent to the injectivity of ff.

Now consider the universal extension of \mathcal{F} by 𝒪C(a)\mathcal{O}_{C}(a).

0ExtΣ21(,𝒪C(a))𝐤𝒪C(a)E0\displaystyle 0\to\operatorname{Ext}_{\Sigma_{2}}^{1}(\mathcal{F},\mathcal{O}_{C}(a))^{\vee}\otimes_{\mathbf{k}}\mathcal{O}_{C}(a)\to E\to\mathcal{F}\to 0 (3.58)

The inequality (3.53) implies HomΣ2(,𝒪C(a))ExtΣ21(,𝒪C(a))[1]\mathop{\mathbb{R}\mathrm{Hom}}\nolimits_{\Sigma_{2}}(\mathcal{F},\mathcal{O}_{C}(a))\simeq\operatorname{Ext}_{\Sigma_{2}}^{1}(\mathcal{F},\mathcal{O}_{C}(a))[-1], so that the distinguished triangle which (3.58) yields is isomorphic to the defining distinguished triangle (2.17) for the inverse spherical twist Ta()T^{\prime}_{a}(\mathcal{F}). In particular, there is an isomorphism ETa()E\simeq T^{\prime}_{a}(\mathcal{F}) and hence EE is an exceptional sheaf. Moreover, the injectivity of ff and the basic properties of universal extensions imply that there is an isomorphism

E(cokerf𝐤𝒪C(a)).\displaystyle\mathcal{H}^{\prime}\simeq E\oplus\left(\operatorname{coker}f\otimes_{\mathbf{k}}\mathcal{O}_{C}(a)\right). (3.59)

In what follows let ddim𝐤ExtΣ21(,𝒪C(a))d\coloneqq\dim_{\mathbf{k}}\operatorname{Ext}_{\Sigma_{2}}^{1}(\mathcal{F},\mathcal{O}_{C}(a)), so that

E𝒪C(a)sd.\mathcal{H}^{\prime}\simeq E\oplus\mathcal{O}_{C}(a)^{\oplus s-d}.

If d=0d=0, then EE\simeq\mathcal{F} is a vector bundle and we are done. So, in the rest of the proof, we assume d>0d>0; i.e., we assume that torsE0\operatorname{tors}E\neq 0.

Let us prove

ExtΣ21(,𝒪C(a+1))=0,\displaystyle\operatorname{Ext}_{\Sigma_{2}}^{1}(\mathcal{H}^{\prime},\mathcal{O}_{C}(a+1))=0, (3.60)

which together with (3.54) implies

i0()𝒪C(a+1)tE𝒪C(a)sd𝒪C(a+1)t.\displaystyle\mathcal{H}^{i_{0}}(\mathcal{E})\simeq\mathcal{H}^{\prime}\oplus\mathcal{O}_{C}(a+1)^{\oplus t}\simeq E\oplus\mathcal{O}_{C}(a)^{\oplus s-d}\oplus\mathcal{O}_{C}(a+1)^{\oplus t}. (3.61)

Since torsE𝒪C(a)d\operatorname{tors}E\simeq\mathcal{O}_{C}(a)^{\oplus d}, this is the desired conclusion.

(3.60) follows from the local-to-global spectral sequence and the following vanishings.

H1(Σ2,omΣ2(,𝒪C(a+1)))\displaystyle H^{1}(\Sigma_{2},\mathop{{\mathcal{H}}om}\nolimits_{\Sigma_{2}}(\mathcal{H}^{\prime},\mathcal{O}_{C}(a+1))) =0\displaystyle=0 (3.62)
H0(Σ2,xtΣ21(,𝒪C(a+1)))\displaystyle H^{0}(\Sigma_{2},\mathop{{\mathcal{E}}xt}\nolimits_{\Sigma_{2}}^{1}(\mathcal{H}^{\prime},\mathcal{O}_{C}(a+1))) =0\displaystyle=0 (3.63)

(3.62), in turn, follows from =E𝒪C(a)sd\mathcal{H}^{\prime}=E\oplus\mathcal{O}_{C}(a)^{\oplus s-d} and [OU15, Theorem 1.4(1)], which says that an exceptional sheaf EE whose torsion part is a non-zero direct sum of copies of 𝒪C(a)\mathcal{O}_{C}(a) satisfies

E|C𝒪C(a+1)r\displaystyle E|_{C}\simeq\mathcal{O}_{C}(a+1)^{\oplus r} (3.64)

for some rr.

Finally, (3.63) follows from the following isomorphisms.

xtΣ21(,𝒪C(a+1))(3.55)xtΣ21(V𝐤𝒪C(a),𝒪C(a+1))V𝒪C(1)\mathop{{\mathcal{E}}xt}\nolimits_{\Sigma_{2}}^{1}(\mathcal{H}^{\prime},\mathcal{O}_{C}(a+1))\stackrel{{\scriptstyle\eqref{eq:cH'-extension}}}{{\simeq}}\mathop{{\mathcal{E}}xt}\nolimits_{\Sigma_{2}}^{1}(V\otimes_{\mathbf{k}}\mathcal{O}_{C}(a),\mathcal{O}_{C}(a+1))\simeq V^{\vee}\otimes\mathcal{O}_{C}(-1)

From Lemma 3.21 we immediately obtain

Corollary 3.23.

Suppose that \mathcal{E} is an exceptional object with torsE0\operatorname{tors}E\neq 0. Then, with the notation of Lemma 3.21, it holds that

tors()=ii0i()TtorsE𝒪C(a)s𝒪C(a+1)()s\displaystyle\operatorname{tors}\mathcal{H}^{\bullet}(\mathcal{E})=\bigoplus_{i\neq i_{0}}\mathcal{H}^{i}(\mathcal{E})\oplus T\oplus\operatorname{tors}E\simeq\mathcal{O}_{C}(a)^{\oplus s}\oplus\mathcal{O}_{C}(a+1)^{\oplus\ell(\mathcal{E})-s} (3.65)

for some 1s()1\leq s\leq\ell(\mathcal{E}).

Remark 3.24.

This is analogous to [IU05, Corollary 4.10] (see also [IU05, Section 5])

Proof.

If ii0i\neq i_{0}, then Suppi()=C\operatorname{Supp}\mathcal{H}^{i}(\mathcal{E})=C and hence i()ωΣ2i()\mathcal{H}^{i}(\mathcal{E})\otimes\omega_{\Sigma_{2}}\simeq\mathcal{H}^{i}(\mathcal{E}). Thus we see

ExtΣ21(i0(),i())Serre dualityExtΣ21(i(),i0())=Lemma 3.3 (1)0.\operatorname{Ext}_{\Sigma_{2}}^{1}(\mathcal{H}^{i_{0}}(\mathcal{E}),\mathcal{H}^{i}(\mathcal{E}))\stackrel{{\scriptstyle\text{Serre duality}}}{{\simeq}}\operatorname{Ext}_{\Sigma_{2}}^{1}(\mathcal{H}^{i}(\mathcal{E}),\mathcal{H}^{i_{0}}(\mathcal{E}))^{\vee}\stackrel{{\scriptstyle\text{Lemma~{}\ref{lm:properties of cohomology sheaves} \eqref{it:cohomology(E) is rigid}}}}{{=}}0.

Since EE is a direct summand of i0()\mathcal{H}^{i_{0}}(\mathcal{E}), this implies ExtΣ21(E,i())=0,ExtΣ21(i(),E)=0\operatorname{Ext}_{\Sigma_{2}}^{1}(E,\mathcal{H}^{i}(\mathcal{E}))=0,\operatorname{Ext}_{\Sigma_{2}}^{1}(\mathcal{H}^{i}(\mathcal{E}),E)=0. Moreover, since dimSuppi()=1\dim\operatorname{Supp}\mathcal{H}^{i}(\mathcal{E})=1, the local to global spectral sequence for Ext groups implies

H1(Σ2,omΣ2(E,i()))=0,H1(Σ2,omΣ2(i(),E))=0.\displaystyle\begin{aligned} H^{1}(\Sigma_{2},\mathop{{\mathcal{H}}om}\nolimits_{\Sigma_{2}}(E,\mathcal{H}^{i}(\mathcal{E})))=0,\\ H^{1}(\Sigma_{2},\mathop{{\mathcal{H}}om}\nolimits_{\Sigma_{2}}(\mathcal{H}^{i}(\mathcal{E}),E))=0.\end{aligned} (3.66)

On the other hand, by Proposition 3.18 and the rigidity, there is a vector bundle 𝒱\mathcal{V} on CC such that i()ι𝒱\mathcal{H}^{i}(\mathcal{E})\simeq\iota_{\ast}\mathcal{V}. Hence there are isomorphisms as follows.

omΣ2(E,i())ιomC(E|C,𝒱)(3.64)ιomC(𝒪C(a+1),𝒱)romΣ2(i(),E)(3.58)omΣ2(i(),𝒪C(a))dιomC(𝒱,𝒪C(a))d\displaystyle\begin{aligned} \mathop{{\mathcal{H}}om}\nolimits_{\Sigma_{2}}(E,\mathcal{H}^{i}(\mathcal{E}))\simeq\iota_{\ast}\mathop{{\mathcal{H}}om}\nolimits_{C}(E|_{C},\mathcal{V})\stackrel{{\scriptstyle\eqref{eq:E|C = OC(a+1) oplus ell}}}{{\simeq}}\iota_{\ast}\mathop{{\mathcal{H}}om}\nolimits_{C}(\mathcal{O}_{C}(a+1),\mathcal{V})^{\oplus r}\\ \mathop{{\mathcal{H}}om}\nolimits_{\Sigma_{2}}(\mathcal{H}^{i}(\mathcal{E}),E)\stackrel{{\scriptstyle\eqref{eq:universal extension}}}{{\simeq}}\mathop{{\mathcal{H}}om}\nolimits_{\Sigma_{2}}(\mathcal{H}^{i}(\mathcal{E}),\mathcal{O}_{C}(a))^{\oplus d}\simeq\iota_{\ast}\mathop{{\mathcal{H}}om}\nolimits_{C}(\mathcal{V},\mathcal{O}_{C}(a))^{\oplus d}\end{aligned} (3.67)

Combining these isomorphisms with (3.66), we obtain the following vanishings.

H1(C,omC(𝒪C(a+1),𝒱)=0H1(C,omC(𝒱,𝒪C(a))=0\displaystyle\begin{aligned} H^{1}(C,\mathop{{\mathcal{H}}om}\nolimits_{C}(\mathcal{O}_{C}(a+1),\mathcal{V})=0\\ H^{1}(C,\mathop{{\mathcal{H}}om}\nolimits_{C}(\mathcal{V},\mathcal{O}_{C}(a))=0\end{aligned} (3.68)

From this we deduce that 𝒱\mathcal{V} is of the form

𝒱𝒪C(a)si𝒪C(a+1)ti\mathcal{V}\simeq\mathcal{O}_{C}(a)^{\oplus s_{i}}\oplus\mathcal{O}_{C}(a+1)^{\oplus t_{i}}

for some sis_{i} and tit_{i}, concluding the proof. ∎

3.4. Derived dual of exceptional objects

Let \mathcal{E} be an exceptional object which is not isomorphic to a shift of a vector bundle, and let E=E()E=E(\mathcal{E}) be the exceptional sheaf in Definition 3.22. In what follows we will mainly discuss the case where torsE0\operatorname{tors}E\neq 0. If torsE=0\operatorname{tors}E=0 (\iff EE is a vector bundle), we will replace \mathcal{E} with its derived dual \mathcal{E}^{\vee} and reduce the problem to the main case. What we mean by this will be made precise by Corollary 3.27.

Lemma 3.25.

For an exceptional object \mathcal{E} on Σ2\Sigma_{2}, the cohomology sheaves of the derived dual \mathcal{E}^{\vee} are related to those of \mathcal{E} as follows:

  • i0()=i0i_{0}(\mathcal{E}^{\vee})=-i_{0}.

  • If ii0i\neq-i_{0}, then i()xtΣ21(i+1(),𝒪Σ2)\mathcal{H}^{i}(\mathcal{E}^{\vee})\simeq\mathop{{\mathcal{E}}xt}\nolimits_{\Sigma_{2}}^{1}(\mathcal{H}^{-i+1}(\mathcal{E}),\mathcal{O}_{\Sigma_{2}}).

  • For i=i0i=-i_{0}, i0()\mathcal{H}^{-i_{0}}(\mathcal{E}^{\vee}) fits into an exact sequence

    0xtΣ21(i0+1(),𝒪Σ2)i0()omΣ2(i0(),𝒪Σ2)0.0\to\mathop{{\mathcal{E}}xt}\nolimits_{\Sigma_{2}}^{1}(\mathcal{H}^{i_{0}+1}(\mathcal{E}),\mathcal{O}_{\Sigma_{2}})\to\mathcal{H}^{-i_{0}}(\mathcal{E}^{\vee})\to\mathop{{\mathcal{H}}om}\nolimits_{\Sigma_{2}}(\mathcal{H}^{i_{0}}(\mathcal{E}),\mathcal{O}_{\Sigma_{2}})\to 0.
  • For i=i0+1i=-i_{0}+1, the cohomology sheaf can also be written as

    i0+1()xtΣ21(torsi0(),𝒪Σ2)(torsi0())[1].\mathcal{H}^{-i_{0}+1}(\mathcal{E}^{\vee})\simeq\mathop{{\mathcal{E}}xt}\nolimits_{\Sigma_{2}}^{1}(\operatorname{tors}\mathcal{H}^{i_{0}}(\mathcal{E}),\mathcal{O}_{\Sigma_{2}})\simeq\left(\operatorname{tors}\mathcal{H}^{i_{0}}(\mathcal{E})\right)^{\vee}[1].
Proof.

Consider the following spectral sequence.

E2p,q=xtΣ2p(q(),𝒪Σ2)p+q()E_{2}^{p,q}=\mathop{{\mathcal{E}}xt}\nolimits_{\Sigma_{2}}^{p}(\mathcal{H}^{-q}(\mathcal{E}),\mathcal{O}_{\Sigma_{2}})\Rightarrow\mathcal{H}^{p+q}(\mathcal{E}^{\vee})

Since Σ2\Sigma_{2} is a smooth projective surface, E2p,q=0E_{2}^{p,q}=0 if p<0p<0 or p>2p>2. Since i()\mathcal{H}^{i}(\mathcal{E}) is torsion for ii0i\neq i_{0} by Lemma 3.3 (2), E20,q=0E_{2}^{0,q}=0 for qi0q\neq-i_{0}. Moreover, for any ii0i\neq i_{0}, i()\mathcal{H}^{i}(\mathcal{E}) is pure by Lemma 3.3 (2). Furthermore, torsi0()\operatorname{tors}\mathcal{H}^{i_{0}}(\mathcal{E}) is pure again by Lemma 3.3 (2) and i0()/torsi0()\mathcal{H}^{i_{0}}(\mathcal{E})/\operatorname{tors}\mathcal{H}^{i_{0}}(\mathcal{E}) is locally free by Lemma 3.3 (3). Hence by [HL10, Theorem 1.1.10, 1)2)1)\Rightarrow 2)], E22,q=0E_{2}^{2,q}=0 for any qq\in\mathbb{Z}.

Summing up, we see that E2p,q0E_{2}^{p,q}\neq 0 only if (p,q)=(0,i0)(p,q)=(0,-i_{0}) or p=1p=1. In particular, this spectral sequence is E2E_{2}-degenerate. All assertions follow from these observations. ∎

Lemma 3.26.

Let \mathcal{E} be an exceptional object such that both E=E()E=E(\mathcal{E}) and E()E(\mathcal{E}^{\vee}) are vector bundles. Then E[i0]\mathcal{E}\simeq E[-i_{0}].

Proof.

We consider the spectral sequence

E2p,q=ExtΣ2p(E,q())Homp+q(E,).E_{2}^{p,q}=\operatorname{Ext}_{\Sigma_{2}}^{p}(E,\mathcal{H}^{q}(\mathcal{E}))\Rightarrow\operatorname{Hom}^{p+q}(E,\mathcal{E}).

As Σ2\Sigma_{2} is a smooth surface, E2p,q0E_{2}^{p,q}\neq 0 only if 0p20\leq p\leq 2. In particular, it is E3E_{3}-degenerate. Since E22,i01HomΣ2(i01(),E)=0E_{2}^{2,i_{0}-1}\simeq\operatorname{Hom}_{\Sigma_{2}}(\mathcal{H}^{i_{0}-1}(\mathcal{E}),E)^{\vee}=0 for i01()\mathcal{H}^{i_{0}-1}(\mathcal{E}) being torsion and EE being torsion free, it follows that d20,i0=0d_{2}^{0,i_{0}}=0 and hence

E20,i0=kerd20,i0E30,i0E0,i0.\displaystyle E_{2}^{0,i_{0}}=\ker d_{2}^{0,i_{0}}\simeq E_{3}^{0,i_{0}}\simeq E_{\infty}^{0,i_{0}}. (3.69)

Take any

[φ:E[i0]]Exti0(E,)[\varphi\colon E[-i_{0}]\to\mathcal{E}]\in\operatorname{Ext}^{i_{0}}(E,\mathcal{E})

whose image under the surjection Exti0(E,)E0,i0\operatorname{Ext}^{i_{0}}(E,\mathcal{E})\twoheadrightarrow E_{\infty}^{0,i_{0}} of the spectral sequence, which in fact is i0(φ)\mathcal{H}^{i_{0}}(\varphi), corresponds to the natural inclusion Ei0()E\hookrightarrow\mathcal{H}^{i_{0}}(\mathcal{E}) under the isomorphisms (3.69). Then the derived dual

φ:E[i0]\varphi^{\vee}\colon\mathcal{E}^{\vee}\to E^{\vee}[i_{0}]

induces the surjection

i0()omΣ2(i0(),𝒪Σ2)E\mathcal{H}^{-i_{0}}(\mathcal{E}^{\vee})\to\mathop{{\mathcal{H}}om}\nolimits_{\Sigma_{2}}(\mathcal{H}^{i_{0}}(\mathcal{E}),\mathcal{O}_{\Sigma_{2}})\simeq E^{\vee}

in Lemma 3.25. As we assumed that E()EE(\mathcal{E}^{\vee})\simeq E^{\vee} is torsion free, by applying what we have just confirmed to \mathcal{E}^{\vee}, we similarly obtain a morphism

ψ:E[i0]\psi\colon E^{\vee}[i_{0}]\to\mathcal{E}^{\vee}

such that i0(ψ)\mathcal{H}^{-i_{0}}(\psi) is the inclusion EE()i0()E^{\vee}\simeq E(\mathcal{E}^{\vee})\hookrightarrow\mathcal{H}^{-i_{0}}(\mathcal{E}^{\vee}) as a direct summand. Then it follows that the composite φψ\varphi^{\vee}\circ\psi is an automorphism of E[i0]E^{\vee}[i_{0}] and therefore \mathcal{E}^{\vee} splits as a direct sum of E[i0]E^{\vee}[i_{0}] and Coneψ\operatorname{Cone}\psi. Since the exceptional object \mathcal{E}^{\vee} is indecomposable, this implies Coneψ=0\operatorname{Cone}\psi=0 and hence E[i0]\mathcal{E}^{\vee}\simeq E^{\vee}[i_{0}], which is equivalent to E[i0]\mathcal{E}\simeq E[-i_{0}]. ∎

Thus we immediately obtain the following

Corollary 3.27.

Let \mathcal{E} be an exceptional object which is not isomorphic to a shift of a vector bundle. If E=E()E=E(\mathcal{E}) is torsion free, then torsE()0\operatorname{tors}E(\mathcal{E}^{\vee})\neq 0.

3.5. Length of the torsion part

Now we introduce the notion of length, which measures for an exceptional object the distance from a shift of a vector bundle. The proof of Theorem 3.1 is reduced to the assertion Theorem 3.32 that one can always reduce the length by an appropriate spherical twist.

Definition 3.28.

Let γ\gamma be the generic point of CC and 𝒪Σ2,γ\mathcal{O}_{\Sigma_{2},\gamma} the local ring of Σ2\Sigma_{2} at γ\gamma. For a coherent sheaf \mathcal{H}, put

()=length𝒪Σ2,γtorsγ\ell(\mathcal{H})=\operatorname{length}_{\mathcal{O}_{\Sigma_{2},\gamma}}\operatorname{tors}\mathcal{H}_{\gamma}

where γ\mathcal{H}_{\gamma} is the stalk of \mathcal{H} at γ\gamma and length𝒪γtorsγ\operatorname{length}_{\mathcal{O}_{\gamma}}\operatorname{tors}\mathcal{H}_{\gamma} is the length of its torsion part.

For an object D(Σ2)\mathcal{E}\in D(\Sigma_{2}), we define

()i(i()).\ell(\mathcal{E})\coloneqq\sum_{i}\ell(\mathcal{H}^{i}(\mathcal{E})).

Let us give a bit more concrete description for ()\ell(\mathcal{E}). Let γ𝐃(𝒪Σ2,γ)\mathcal{E}_{\gamma}\in\mathbf{D}(\mathcal{O}_{\Sigma_{2},\gamma}) be the pull-back of \mathcal{E} by the flat morphism Spec𝒪Σ2,γΣ2\operatorname{Spec}\mathcal{O}_{\Sigma_{2},\gamma}\to\Sigma_{2}. The length function \ell is defined for objects in 𝐃(𝒪Σ2,γ)\mathbf{D}(\mathcal{O}_{\Sigma_{2},\gamma}) as well, and it immediately follows from the exactness of the (underived) pull-back functor that (γ)=()\ell(\mathcal{E}_{\gamma})=\ell(\mathcal{E}). On the other hand, since 𝒪Σ2,γ\mathcal{O}_{\Sigma_{2},\gamma} is a DVR, for 𝐃(𝒪Σ2,γ)\mathcal{M}\in\mathbf{D}(\mathcal{O}_{\Sigma_{2},\gamma}) there is an isomorphism ii()[i]\mathcal{M}\simeq\bigoplus_{i\in\mathbb{Z}}\mathcal{H}^{i}(\mathcal{M})[-i] and hence the equality ()=i(i())\ell(\mathcal{M})=\sum_{i\in\mathbb{Z}}\ell(\mathcal{H}^{i}(\mathcal{M})). By the structure theorem for finitely generated modules over a DVR, i()\mathcal{H}^{i}(\mathcal{M}) is a direct sum of finite copies of 𝒪Σ2,γ\mathcal{O}_{\Sigma_{2},\gamma} and 𝒪Σ2,γ/(tp)\mathcal{O}_{\Sigma_{2},\gamma}/(t^{p}) for various p1p\geq 1, where tt is a generator of the maximal ideal. Based on this, we obtain the following invariance.

Lemma 3.29.

For any object 𝐃(Σ2)\mathcal{E}\in\mathbf{D}(\Sigma_{2}), ()=()\ell(\mathcal{E}^{\vee})=\ell(\mathcal{E}).

Proof.

Note first that ()=(()γ)=((γ))\ell(\mathcal{E}^{\vee})=\ell((\mathcal{E}^{\vee})_{\gamma})=\ell((\mathcal{E}_{\gamma})^{\vee}). Hence it is enough to show ()=()\ell(\mathcal{M}^{\vee})=\ell(\mathcal{M}) for all 𝐃(𝒪Σ2,γ)\mathcal{M}\in\mathbf{D}(\mathcal{O}_{\Sigma_{2},\gamma}). By the explicit descriptions of \mathcal{M} we gave above, it is enough to show this for =𝒪Σ2,γ/(tp)\mathcal{M}=\mathcal{O}_{\Sigma_{2},\gamma}/(t^{p}). In this case one can easily confirm [1]\mathcal{M}^{\vee}\simeq\mathcal{M}[-1], so we are done. ∎

If \mathcal{E} is an exceptional object with Supp(i0)=Σ2\operatorname{Supp}(\mathcal{H}^{i_{0}})=\Sigma_{2}, then Proposition 3.18 implies

()=irankCtors(i())=ii0rankCi()+rankCtorsi0(),\displaystyle\begin{aligned} \ell(\mathcal{E})&=\sum_{i}\operatorname{rank}_{C}\operatorname{tors}(\mathcal{H}^{i}(\mathcal{E}))\\ &=\sum_{i\neq i_{0}}\operatorname{rank}_{C}\mathcal{H}^{i}(\mathcal{E})+\operatorname{rank}_{C}\operatorname{tors}\mathcal{H}^{i_{0}}(\mathcal{E}),\end{aligned} (3.70)

where rankC\operatorname{rank}_{C} denotes the rank of a coherent sheaf on CC. Note that thus defined ()\ell(\mathcal{E}) is the same as ()\ell(\mathcal{E}) in Corollary 3.23. From this we immediately obtain the following characterization of exceptional vector bundles among exceptional objects.

Lemma 3.30.

An exceptional object \mathcal{E} is isomorphic to a shift of a vector bundle if and only if ()=0\ell(\mathcal{E})=0.

The following (sub)additivity of the length function with respect to short exact sequences will be useful later.

Lemma 3.31.

For an exact sequence

012300\to\mathcal{H}_{1}\to\mathcal{H}_{2}\to\mathcal{H}_{3}\to 0

in cohΣ2\operatorname{coh}\Sigma_{2}, an inequality

(2)(1)+(3)\ell(\mathcal{H}_{2})\leq\ell(\mathcal{H}_{1})+\ell(\mathcal{H}_{3})

holds. This is an equality if 1\mathcal{H}_{1} is a torsion sheaf.

Proof.

By taking the stalks at γ\gamma, it is enough to show the similar statements for coh𝒪Σ2,γ\operatorname{coh}\mathcal{O}_{\Sigma_{2},\gamma}. It follows from the standard arguments on local cohomology that

0tors1,γtors2,γtors3,γ\displaystyle 0\to\operatorname{tors}\mathcal{H}_{1,\gamma}\to\operatorname{tors}\mathcal{H}_{2,\gamma}\to\operatorname{tors}\mathcal{H}_{3,\gamma} (3.71)

is exact and that the last map is a surjection if 1\mathcal{H}_{1} is itself torsion (note that tors\operatorname{tors} is isomorphic to the 0-th local cohomology functor Γ{0}\Gamma_{\{0\}} at the closed point 0Spec𝒪Σ2,γ0\in\operatorname{Spec}\mathcal{O}_{\Sigma_{2},\gamma}). The assertions immediately follow from this and the additivity of the length of modules under short exact sequences. ∎

3.6. Proof of Theorem 3.1

Let us complete the proof of Theorem 3.1. Recall again the decomposition

i0()ET\mathcal{H}^{i_{0}}(\mathcal{E})\simeq E\oplus T

from Lemma 3.21, where EE is an exceptional sheaf and TT is a torsion sheaf. There are two possibilities as follows.

  1. (1)

    EE is not torsion free. Then by Lemma 3.21, there are a,d>0a\in\mathbb{Z},d\in\mathbb{Z}_{>0} such that

    torsE𝒪C(a)d.\displaystyle\operatorname{tors}E\simeq\mathcal{O}_{C}(a)^{\oplus d}. (3.72)
  2. (2)

    EE is torsion free. Then by Corollary 3.27, \mathcal{E}^{\vee} fits into the case (1). Namely, the exceptional sheaf EE()E^{\prime}\coloneqq E(\mathcal{E}^{\vee}) has a non-trivial torsion and there is a torsion sheaf TT^{\prime}, aa^{\prime}\in\mathbb{Z}, and d>0d^{\prime}>0 such that i0()ET\mathcal{H}^{-i_{0}}(\mathcal{E}^{\vee})\simeq E^{\prime}\oplus T^{\prime} and torsE𝒪C(a)d\operatorname{tors}E^{\prime}\simeq\mathcal{O}_{C}(a^{\prime})^{\oplus d^{\prime}}.

Theorem 5.4 obviously follows from the following theorem, which asserts that it is always possible to decrease the length of \mathcal{E} by an appropriate spherical twist.

Theorem 3.32.

Under the notation of the previous paragraphs, the following holds.

  • In the case (1), i.e., if EE is not torsion free, then (Ta)<()\ell(T_{a}\mathcal{E})<\ell(\mathcal{E}).

  • In the case (2), i.e., if EE is torsion free, then (Ta3)<()\ell(T_{-a^{\prime}-3}\mathcal{E})<\ell(\mathcal{E}).

Proof.

In this proof, to make life easy, we assume i0=i0()=0i_{0}=i_{0}(\mathcal{E})=0. The general case is easily reduced to this just by replacing \mathcal{E} with [i0]\mathcal{E}[i_{0}].

Consider first the case (1). For each i0i\neq 0, by Corollary 3.23 there are si,ti0s_{i},t_{i}\geq 0 such that i()𝒪C(a)si𝒪C(a+1)ti\mathcal{H}^{i}(\mathcal{E})\simeq\mathcal{O}_{C}(a)^{\oplus s_{i}}\oplus\mathcal{O}_{C}(a+1)^{\oplus t_{i}}. Putting s0sds_{0}\coloneqq s-d and t0tt_{0}\coloneqq t in (3.61), we can also write 0()E𝒪C(a)s0𝒪C(a+1)t0\mathcal{H}^{0}(\mathcal{E})\simeq E\oplus\mathcal{O}_{C}(a)^{\oplus s_{0}}\oplus\mathcal{O}_{C}(a+1)^{\oplus t_{0}}. Consider the following spectral sequence.

E2p,q=p(Ta(q()))p+q(Ta)E_{2}^{p,q}=\mathcal{H}^{p}(T_{a}(\mathcal{H}^{q}(\mathcal{E})))\Rightarrow\mathcal{H}^{p+q}(T_{a}\mathcal{E})

By direct computations we easily see that E2p,qE_{2}^{p,q} are as follows.

E2p,q{𝒪C(a1)tqp=1(=E/torsE)(p,q)=(0,0)𝒪C(a)sqp=10otherwise\displaystyle E_{2}^{p,q}\simeq\begin{cases}\mathcal{O}_{C}(a-1)^{\oplus t_{q}}&p=-1\\ \mathcal{F}(=E/\operatorname{tors}E)&(p,q)=(0,0)\\ \mathcal{O}_{C}(a)^{\oplus s_{q}}&p=1\\ 0&\text{otherwise}\end{cases} (3.73)

Thus we see

p,q(E2p,q)=()(torsE)<().\displaystyle\sum_{p,q}\ell(E_{2}^{p,q})=\ell(\mathcal{E})-\ell(\operatorname{tors}E)<\ell(\mathcal{E}). (3.74)

Noting that the differential map d2p,qd_{2}^{p,q} is non-zero only if p=1p=-1, we easily see that E3p,qE_{3}^{p,q} are as follows.

E3p,q{kerd21,qp=1cokerd21,q+1p=1E2p,qotherwiseE_{3}^{p,q}\simeq\begin{cases}\ker d_{2}^{-1,q}&p=-1\\ \operatorname{coker}d_{2}^{-1,q+1}&p=1\\ E_{2}^{p,q}&\text{otherwise}\end{cases}

For each qq\in\mathbb{Z}, since E21,qE_{2}^{-1,q} is torsion, it follows from Lemma 3.31 that (cokerd21,q+1)(E21,q1)\ell(\operatorname{coker}d_{2}^{-1,q+1})\leq\ell(E_{2}^{1,q-1}). Also, the inclusion kerd21,qE21,q\ker d_{2}^{-1,q}\hookrightarrow E_{2}^{-1,q} implies (kerd21,q)(E21,q)\ell(\ker d_{2}^{-1,q})\leq\ell(E_{2}^{-1,q}). Thus we have confirmed the inequality

(E3p,q)(E2p,q)(p,q).\displaystyle\ell(E_{3}^{p,q})\leq\ell(E_{2}^{p,q})\quad\forall(p,q). (3.75)

Finally, the spectral sequence degenerates at E3E_{3} and hence Lemma 3.31 implies

(Ta)p,q(Ep,q)=p,q(E3p,q).\displaystyle\ell(T_{a}\mathcal{E})\leq\sum_{p,q}\ell(E_{\infty}^{p,q})=\sum_{p,q}\ell(E_{3}^{p,q}). (3.76)

Combining (3.74), (3.75), and (3.76), we conclude (Ta)<()\ell(T_{a}\mathcal{E})<\ell(\mathcal{E}).

Consider the other case (2). By Corollary 3.27, \mathcal{E}^{\vee} is then as in the case (1). By applying our conclusion for the case (1) to \mathcal{E}^{\vee}, we obtain the inequality

(Ta())<().\displaystyle\ell(T_{a^{\prime}}(\mathcal{E}^{\vee}))<\ell(\mathcal{E}^{\vee}). (3.77)

On the other hand, the left hand side is computed as follows.

(Ta())=(𝒪Σ2(C)Ta())=(2.27)(Ta+1())=(2.21)((Ta3))=Lemma 3.29(Ta3).\displaystyle\begin{aligned} \ell(T_{a^{\prime}}\left(\mathcal{E}^{\vee}\right))=\ell(\mathcal{O}_{\Sigma_{2}}(-C)\otimes T_{a^{\prime}}\left(\mathcal{E}^{\vee}\right))\stackrel{{\scriptstyle\eqref{eq:Ta Ta+1 = O(C)}}}{{=}}\ell(T^{\prime}_{a^{\prime}+1}\left(\mathcal{E}^{\vee}\right))\\ \stackrel{{\scriptstyle\eqref{eq:dual and spherical twists}}}{{=}}\ell\left(\left(T_{-a^{\prime}-3}\mathcal{E}\right)^{\vee}\right)\stackrel{{\scriptstyle\text{Lemma~{}\ref{lm:length is invariant under dual}}}}{{=}}\ell\left(T_{-a^{\prime}-3}\mathcal{E}\right).\end{aligned} (3.78)

Thus we see that (Ta3)=(3.78)(Ta())<(3.77)()=Lemma 3.29()\ell(T_{-a^{\prime}-3}\mathcal{E})\stackrel{{\scriptstyle\eqref{eq:computation of length}}}{{=}}\ell(T_{a^{\prime}}\left(\mathcal{E}^{\vee}\right))\stackrel{{\scriptstyle\eqref{eq: Ta' decreases the length of cEvee}}}{{<}}\ell(\mathcal{E}^{\vee})\stackrel{{\scriptstyle\text{Lemma~{}\ref{lm:length is invariant under dual}}}}{{=}}\ell(\mathcal{E}). ∎

Remark 3.33.

Suppose both torsE()0\operatorname{tors}E(\mathcal{E})\neq 0 and torsE()0\operatorname{tors}E(\mathcal{E}^{\vee})\neq 0. From the third item of Lemma 3.25 and (5.4) in the proof of Lemma 5.1 below, it actually follows that a=a3a=-a^{\prime}-3. This does not seem to be a mere coincidence, but the authors do not have a good account of this.

4. Exceptional objects sharing the same class in K0(Σ2)\operatorname{K_{0}}\left(\Sigma_{2}\right)

The goal of this section is to prove Corollary 4.4 on the set of exceptional objects sharing the same class in K0(Σ2)\operatorname{K_{0}}\left(\Sigma_{2}\right).

Let 𝐃(Σ2)\mathcal{E}\in\mathbf{D}(\Sigma_{2}) be an exceptional vector bundle of rank rr. Recall from (3.50) of Lemma 3.20 that there is an isomorphism

|C𝒪C(b)s𝒪C(b+1)rs\displaystyle\mathcal{E}|_{C}\simeq\mathcal{O}_{C}(b)^{\oplus s}\oplus\mathcal{O}_{C}(b+1)^{\oplus r-s} (4.1)

for some bb\in\mathbb{Z} and ss\in\mathbb{N} such that 1sr1\leq s\leq r. We freely use this result, especially the symbols r,sr,s, and bb, throughout this section.

Lemma 4.1.

Let 𝐃(Σ2)\mathcal{E}\in\mathbf{D}(\Sigma_{2}) be an exceptional vector bundle. Then Tb1T_{b-1}\mathcal{E} and TbT^{\prime}_{b}\mathcal{E} are exceptional vector bundles.

Proof.

From the defining distinguished triangle of spherical twists (2.16), it immediately follows that Tb1T_{b-1}\mathcal{E} is an exceptional sheaf. Note that it is isomorphic to (Tb)(C)\left(T^{\prime}_{b}\mathcal{E}\right)(C). It then follows from the defining distinguished triangle for the inverse spherical twist (2.17) that TbT^{\prime}_{b}\mathcal{E} is torsion free. Thus we see that Tb1T_{b-1}\mathcal{E} and TbT^{\prime}_{b}\mathcal{E} are both exceptional vector bundles. ∎

Lemma 4.2.

Let 𝐃(Σ2)\mathcal{E}\in\mathbf{D}(\Sigma_{2}) be an exceptional vector bundle such that s=rs=r holds in (3.50). Then Tb1T_{b-1}\mathcal{E}\simeq\mathcal{E} and Tb2(C)T_{b-2}\mathcal{E}\simeq\mathcal{E}(C).

Proof.

The first assertion immediately follows from the vanishing HomΣ2(𝒪C(b1),)=0\mathop{\mathbb{R}\mathrm{Hom}}\nolimits_{\Sigma_{2}}(\mathcal{O}_{C}(b-1),\mathcal{E})=0. The second assertion follows from Tb1\mathcal{E}\simeq T^{\prime}_{b-1}\mathcal{E}, which is nothing but the first assertion, and the isomorphism Tb2𝒪Σ2(C)𝒪Σ2Tb1T_{b-2}\mathcal{E}\simeq\mathcal{O}_{\Sigma_{2}}(C)\otimes_{\mathcal{O}_{\Sigma_{2}}}T^{\prime}_{b-1}\mathcal{E} (which follows from (2.28) and (2.19)). ∎

Theorem 4.3.

Let 𝐃(Σ2)\mathcal{E}\in\mathbf{D}(\Sigma_{2}) be an exceptional object with rank>0\operatorname{rank}\mathcal{E}>0. Then there exists bBK0𝗍𝗋𝗂𝗏b\in B^{K_{0}-\mathsf{triv}} and mm^{\prime}\in\mathbb{Z} such that b()[2m]b(\mathcal{E})[2m^{\prime}] is an exceptional vector bundle.

Proof.

Let us choose an isomorphism as in (3.1). If nn happens to be an odd number, noting that Tb(Tb)\mathcal{F}\simeq T_{b}\left(T^{\prime}_{b}\mathcal{F}\right) and TbT^{\prime}_{b}\mathcal{F} is an exceptional vector bundle for suitable bb\in\mathbb{Z} by Lemma 4.1, we may assume without loss of generality that nn is even. Now the assertion is an immediate consequence of Proposition 2.8, since Ta±2BK0𝗍𝗋𝗂𝗏T_{a}^{\pm 2}\in B^{K_{0}-\mathsf{triv}} for any aa\in\mathbb{Z}. ∎

Corollary 4.4.

Let 𝐃(Σ2)\mathcal{E}\in\mathbf{D}(\Sigma_{2}) be an exceptional object. Then

  1. (1)

    There exists a unique exceptional vector bundle \mathcal{F} on Σ2\Sigma_{2} such that

    []={[]K0(Σ2)if rank>0[]K0(Σ2)if rank<0\displaystyle[\mathcal{E}]=\begin{cases}[\mathcal{F}]\in\operatorname{K_{0}}\left(\Sigma_{2}\right)&\text{if }\operatorname{rank}\mathcal{E}>0\\ -[\mathcal{F}]\in\operatorname{K_{0}}\left(\Sigma_{2}\right)&\text{if }\operatorname{rank}\mathcal{E}<0\end{cases} (4.2)

    (recall that rank0\operatorname{rank}\mathcal{E}\neq 0 by Corollary 3.8).

  2. (2)

    The action of the group BK0𝗍𝗋𝗂𝗏×2B^{K_{0}-\mathsf{triv}}\times 2\mathbb{Z} on the following set is transitive.

    {𝐃(Σ2)exceptional object such that []=[]K0(Σ2)}\displaystyle\left\{\mathcal{E}^{\prime}\in\mathbf{D}(\Sigma_{2})\mid\text{exceptional object such that }[\mathcal{E}^{\prime}]=[\mathcal{E}]\in\operatorname{K_{0}}\left(\Sigma_{2}\right)\right\} (4.3)
Proof.

The existence of \mathcal{F} as in (1) is a direct consequence of Theorem 4.3. The uniqueness of such \mathcal{F} is Lemma 2.27. One can prove (2) again by Theorem 4.3, by showing that any \mathcal{E}^{\prime} is in the same orbit of \mathcal{F} or [1]\mathcal{F}[1], depending on rank>0\operatorname{rank}\mathcal{E}>0 or rank<0\operatorname{rank}\mathcal{E}<0. ∎

5. Constructibility of exceptional collections

The aim of this section is Corollary 5.6, which asserts that any exceptional collection on Σ2\Sigma_{2} is extendable to a full exceptional collection.

We first show that any exceptional collection on Σ2\Sigma_{2} is sent to an exceptional collection consisting of (shifts of) vector bundles by a sequence of spherical twists.

Lemma 5.1.

Let (,)𝖤𝖢2(Σ2)(\mathcal{B},\mathcal{E})\in\mathsf{EC}_{2}(\Sigma_{2}) be an exceptional pair such that \mathcal{B} is a vector bundle and \mathcal{E} is not isomorphic to a shift of a sheaf. Suppose also that E=E()E=E(\mathcal{E}) defined in Definition 3.22 is not torsion free, so that torsE\operatorname{tors}E is a direct sum of copies of 𝒪C(a)\mathcal{O}_{C}(a) for some aa\in\mathbb{Z}. Then TaT_{a}\mathcal{B} is isomorphic to either \mathcal{B} or (C)\mathcal{B}(C); in particular, it remains to be a vector bundle.

Proof.

Recall the decomposition i0()=ET\mathcal{H}^{i_{0}}(\mathcal{E})=E\oplus T from Lemma 3.21. Set torsE𝒪C(a)d\operatorname{tors}E\simeq\mathcal{O}_{C}(a)^{\oplus d}, where d>0d>0 by the assumption. By Corollary 3.23, there is an isomorphism

tors()ii0i()TtorsE𝒪C(a)s𝒪C(a+1)()s\displaystyle\operatorname{tors}\mathcal{H}^{\bullet}(\mathcal{E})\simeq\bigoplus_{i\neq i_{0}}\mathcal{H}^{i}(\mathcal{E})\oplus T\oplus\operatorname{tors}E\simeq\mathcal{O}_{C}(a)^{\oplus s}\oplus\mathcal{O}_{C}(a+1)^{\oplus\ell(\mathcal{E})-s} (5.1)

for some 1s()1\leq s\leq\ell(\mathcal{E}). Also, again by Lemma 3.20, there is an isomorphism

|C𝒪C(b)s𝒪C(b+1)rs\mathcal{B}|_{C}\simeq\mathcal{O}_{C}(b)^{\oplus s^{\prime}}\oplus\mathcal{O}_{C}(b+1)^{\oplus r^{\prime}-s^{\prime}}

for some bb\in\mathbb{Z} and 1sr1\leq s^{\prime}\leq r^{\prime}.

Consider the following spectral sequence.

E2p,q=ExtΣ2p(q(),)ExtΣ2p+q(,)\displaystyle E_{2}^{p,q}=\operatorname{Ext}_{\Sigma_{2}}^{p}(\mathcal{H}^{-q}(\mathcal{E}),\mathcal{B})\Rightarrow\operatorname{Ext}_{\Sigma_{2}}^{p+q}(\mathcal{E},\mathcal{B}) (5.2)

The assumption HomΣ2(,)=0\mathop{\mathbb{R}\mathrm{Hom}}\nolimits_{\Sigma_{2}}(\mathcal{E},\mathcal{B})=0 implies the vanishing of the limit ExtΣ2n(,)=0\operatorname{Ext}_{\Sigma_{2}}^{n}(\mathcal{E},\mathcal{B})=0 for all nn\in\mathbb{Z}. Also, since Σ2\Sigma_{2} is a smooth projective surface, E2p,q0E_{2}^{p,q}\neq 0 only if 0p20\leq p\leq 2 and hence (5.2) is E3E_{3}-degenerate everywhere and E2E_{2}-degenerate at p=1p=1. These imply that

  • E20,q=0E_{2}^{0,q}=0 for all qi0q\neq-i_{0}, for q()\mathcal{H}^{-q}(\mathcal{E}) being torsion and \mathcal{B} being torsion free. Hence (5.2) is E2E_{2}-degenerate at (2,q)(2,q) for each qi01q\neq-i_{0}-1, so that E22,qE2,q=0E_{2}^{2,q}\simeq E_{\infty}^{2,q}=0.

  • E21,q=0E_{2}^{1,q}=0 for all qq.

Thus we have confirmed that

  • E2p,q0E_{2}^{p,q}\neq 0 only if (p,q)=(0,i0)(p,q)=(0,-i_{0}) or (2,i01)(2,-i_{0}-1). Moreover,

  • d20,i0:E20,i0E22,i01d_{2}^{0,-i_{0}}\colon E_{2}^{0,-i_{0}}\to E_{2}^{2,-i_{0}-1} is an isomorphism by the E3E_{3}-degeneracy of (5.2) and the vanishing of the limit.

Note that 0=E22,i0ExtΣ22(ET,)0=E_{2}^{2,-i_{0}}\simeq\operatorname{Ext}_{\Sigma_{2}}^{2}(E\oplus T,\mathcal{B}) and the Serre duality imply HomΣ2(,torsE)=0\operatorname{Hom}_{\Sigma_{2}}(\mathcal{B},\operatorname{tors}E)=0. Thus we see

ab1.\displaystyle a\leq b-1. (5.3)

Take any qi0,i01q\neq-i_{0},-i_{0}-1. We know that q()\mathcal{H}^{-q}(\mathcal{E}) is a direct sum of invertible sheaves on C1C\simeq\mathbb{P}^{1}, so the vanishings E21,q=0=E22,qE_{2}^{1,q}=0=E_{2}^{2,q} imply q()=0\mathcal{H}^{-q}(\mathcal{E})=0 if s<rs^{\prime}<r^{\prime}, and that q()\mathcal{H}^{-q}(\mathcal{E}) is a direct sum of (possibly 0) copies of 𝒪C(b1)\mathcal{O}_{C}(b-1) if s=rs^{\prime}=r^{\prime}.

Let us first settle the case s<rs^{\prime}<r^{\prime}. As we mentioned in the previous paragraph, in this case i()0\mathcal{H}^{i}(\mathcal{E})\neq 0 only if i=i0i=i_{0} or i0+1i_{0}+1. Note in fact that i0+1()0\mathcal{H}^{i_{0}+1}(\mathcal{E})\neq 0, since it is assumed in the statement that \mathcal{E} is not isomorphic to a shift of a sheaf.

It follows from the vanishing 0=E20,1=HomΣ2(i0+1(),i0())0=E_{2}^{0,-1}=\operatorname{Hom}_{\Sigma_{2}}(\mathcal{H}^{i_{0}+1}(\mathcal{E}),\mathcal{H}^{i_{0}}(\mathcal{E})) in the spectral sequence (3.4) that 0=HomΣ2(i0+1(),torsE)0=\operatorname{Hom}_{\Sigma_{2}}(\mathcal{H}^{i_{0}+1}(\mathcal{E}),\operatorname{tors}E) and thus

i0+1()𝒪C(a+1)t\displaystyle\mathcal{H}^{i_{0}+1}(\mathcal{E})\simeq\mathcal{O}_{C}(a+1)^{\oplus t} (5.4)

for some t>0t>0. On the other hand, we know that

0=E21,i01=ExtΣ21(i0+1(),),0=E_{2}^{1,-i_{0}-1}=\operatorname{Ext}_{\Sigma_{2}}^{1}(\mathcal{H}^{i_{0}+1}(\mathcal{E}),\mathcal{B}),

so that a+1ba+1\geq b (recall rs>0r^{\prime}-s^{\prime}>0). Thus we see that a=b1a=b-1. Hence Ta()T_{a}(\mathcal{B})\simeq\mathcal{B} by Lemma 4.1.

Next consider the case s=rs^{\prime}=r^{\prime}. Note that at least one of the sheaves 𝒪C(a)\mathcal{O}_{C}(a) or 𝒪C(a+1)\mathcal{O}_{C}(a+1) appears as a direct summand of ii0i()\bigoplus_{i\neq i_{0}}\mathcal{H}^{i}(\mathcal{E}), since \mathcal{E} is not a shift of a sheaf. Then the vanishing 0=E21,q=Ext1(q(),)0=E_{2}^{1,q}=\operatorname{Ext}^{1}(\mathcal{H}^{-q}(\mathcal{E}),\mathcal{B}) for all qi0q\neq-i_{0} in (5.2) implies that either a=b1a=b-1 or b2b-2. Then by Lemma 4.2, Ta()T_{a}(\mathcal{B}) is isomorphic to \mathcal{B} and (C)\mathcal{B}(C), respectively. ∎

Lemma 5.2.

Let (,)𝖤𝖢2(Σ2)(\mathcal{B},\mathcal{E})\in\mathsf{EC}_{2}(\Sigma_{2}) be an exceptional pair. Suppose that \mathcal{B} is a vector bundle and \mathcal{E} is a sheaf such that

𝒯tors=torsE𝒪C(a)d0.\mathcal{T}\coloneqq\operatorname{tors}\mathcal{E}=\operatorname{tors}E\simeq\mathcal{O}_{C}(a)^{\oplus d}\neq 0.

Then TaT_{a}\mathcal{B} is isomorphic to either \mathcal{B} or (C)\mathcal{B}(C) and it holds that (Ta,Ta)𝖤𝖢𝖵𝖡2(Σ2)(T_{a}\mathcal{B},T_{a}\mathcal{E})\in\mathsf{ECVB}_{2}(\Sigma_{2}).

Proof.

Consider the short exact sequence as follows.

0𝒯0\displaystyle 0\to\mathcal{T}\to\mathcal{E}\to\mathcal{F}\to 0 (5.5)

By [OU15, Theorem 1.4 (1)], we know that d=homΣ2(𝒪C(a),)d=\hom_{\Sigma_{2}}(\mathcal{O}_{C}(a),\mathcal{E}) and TaT_{a}\mathcal{E}\simeq\mathcal{F} is an exceptional vector bundle. Hence all we have to show is that TaT_{a}\mathcal{B} is a vector bundle.

The vanishing Hom(,)=0\mathop{\mathbb{R}\mathrm{Hom}}\nolimits(\mathcal{E},\mathcal{B})=0 implies ExtΣ2i(𝒯,)ExtΣ2i+1(,)\operatorname{Ext}_{\Sigma_{2}}^{i}(\mathcal{T},\mathcal{B})\simeq\operatorname{Ext}_{\Sigma_{2}}^{i+1}(\mathcal{F},\mathcal{B}). Especially one has

ExtΣ21(𝒯,)\displaystyle\operatorname{Ext}_{\Sigma_{2}}^{1}(\mathcal{T},\mathcal{B}) ExtΣ22(,),\displaystyle\simeq\operatorname{Ext}_{\Sigma_{2}}^{2}(\mathcal{F},\mathcal{B}), (5.6)
ExtΣ22(𝒯,)\displaystyle\operatorname{Ext}_{\Sigma_{2}}^{2}(\mathcal{T},\mathcal{B}) =0.\displaystyle=0. (5.7)

If ExtΣ21(𝒯,)=0\operatorname{Ext}_{\Sigma_{2}}^{1}(\mathcal{T},\mathcal{B})=0, then it follows that Hom(𝒪C(a),)=0\mathop{\mathbb{R}\mathrm{Hom}}\nolimits(\mathcal{O}_{C}(a),\mathcal{B})=0 and hence Ta=T_{a}\mathcal{B}=\mathcal{B}. Therefore we may assume ExtΣ21(𝒯,)0\operatorname{Ext}_{\Sigma_{2}}^{1}(\mathcal{T},\mathcal{B})\neq 0, or by taking the dual,

ExtΣ21(,𝒯KΣ2)0.\displaystyle\operatorname{Ext}_{\Sigma_{2}}^{1}(\mathcal{B},\mathcal{T}\otimes K_{\Sigma_{2}})\neq 0. (5.8)

The Serre dual of (5.6) and its restriction to CC yield the commutative square as follows.

HomΣ2(,KΣ2){\operatorname{Hom}_{\Sigma_{2}}(\mathcal{B},\mathcal{F}\otimes K_{\Sigma_{2}})}ExtΣ21(,𝒯KΣ2){\operatorname{Ext}_{\Sigma_{2}}^{1}(\mathcal{B},\mathcal{T}\otimes K_{\Sigma_{2}})}HomC(|C,|C){\operatorname{Hom}_{C}(\mathcal{B}|_{C},\mathcal{F}|_{C})}ExtC1(|C,𝒯){\operatorname{Ext}_{C}^{1}(\mathcal{B}|_{C},\mathcal{T})}\scriptstyle{\simeq}\scriptstyle{\simeq} (5.9)

In the second row of the diagram (5.9), we omit KΣ2\otimes K_{\Sigma_{2}} by fixing an isomorphism KΣ2𝒪C𝒪CK_{\Sigma_{2}}\otimes\mathcal{O}_{C}\simeq\mathcal{O}_{C} and regard 𝒯\mathcal{T} as a sheaf on CC. By restricting the locally split short exact sequence (5.5) to CC, we obtain the following short exact sequence.

0𝒯|C|C0\displaystyle 0\to\mathcal{T}\to\mathcal{E}|_{C}\to\mathcal{F}|_{C}\to 0 (5.10)

From (5.10) and the surjectivity of the second row in (5.9), we obtain the following isomorphism.

ExtC1(|C,|C)ExtC1(|C,|C)\displaystyle\operatorname{Ext}_{C}^{1}(\mathcal{B}|_{C},\mathcal{E}|_{C})\simeq\operatorname{Ext}_{C}^{1}(\mathcal{B}|_{C},\mathcal{F}|_{C}) (5.11)

As in the proof of Lemma 5.1, put

|C𝒪C(b)s𝒪C(b+1)rs\mathcal{B}|_{C}\simeq\mathcal{O}_{C}(b)^{\oplus s^{\prime}}\oplus\mathcal{O}_{C}(b+1)^{\oplus r^{\prime}-s^{\prime}}

for bb\in\mathbb{Z} and 1sr1\leq s^{\prime}\leq r^{\prime} where r=rankr^{\prime}=\operatorname{rank}\mathcal{B}. Then (5.7) implies

ab1.a\leq b-1.

To determine the value of aa, let us compute the dimensions of the both sides of (5.11).

In order to compute the dimension of the left hand side, recall that |C𝒪C(a+1)e\mathcal{E}|_{C}\simeq\mathcal{O}_{C}(a+1)^{\oplus e} by [OU15, Theorem 1.4(1)]. We know moreover that e=d+re=d+r by (5.10), where r=rankr=\operatorname{rank}\mathcal{F}. Thus we obtain the following descriptions.

extC1(|C,|C)={0if a=b1e((rs)(ba1)+s(ba2))if ab2\displaystyle\operatorname{ext}_{C}^{1}(\mathcal{B}|_{C},\mathcal{E}|_{C})=\begin{cases}0&\text{if }a=b-1\\ e\left((r^{\prime}-s^{\prime})(b-a-1)+s^{\prime}(b-a-2)\right)&\text{if }a\leq b-2\end{cases} (5.12)

Let us compute the dimension of the right hand side of (5.11). Since \mathcal{F} is an exceptional vector bundle, there are ff\in\mathbb{Z} and 1sr1\leq s\leq r such that

|C𝒪C(f)s𝒪C(f+1)rs.\mathcal{F}|_{C}\simeq\mathcal{O}_{C}(f)^{\oplus s}\oplus\mathcal{O}_{C}(f+1)^{\oplus r-s}.

Note that by (5.8) and (5.9) we have

HomC(|C,|C)0,\operatorname{Hom}_{C}(\mathcal{B}|_{C},\mathcal{F}|_{C})\neq 0,

which implies fb1f\geq b-1.

Suppose for a contradiction that extC1(|C,|C)0\operatorname{ext}_{C}^{1}(\mathcal{B}|_{C},\mathcal{F}|_{C})\neq 0. Then we obtain f=b1f=b-1 and

0extC1(|C,|C)=s(rs).\displaystyle 0\neq\operatorname{ext}_{C}^{1}(\mathcal{B}|_{C},\mathcal{F}|_{C})=s(r^{\prime}-s^{\prime}). (5.13)

Using sr<r+d=es\leq r<r+d=e, we immediately see that (5.13) is strictly smaller than the second line of the right hand side of (5.12). This contradicts the isomorphism (5.11). Thus we have confirmed

extC1(|C,|C)=(5.11)extC1(|C,|C)=0.\operatorname{ext}_{C}^{1}(\mathcal{B}|_{C},\mathcal{E}|_{C})\stackrel{{\scriptstyle\eqref{eq:isom of Ext^1}}}{{=}}\operatorname{ext}_{C}^{1}(\mathcal{B}|_{C},\mathcal{F}|_{C})=0.

Then (5.12) implies either a=b1a=b-1 or (a,rs)=(b2,0)(a,r^{\prime}-s^{\prime})=(b-2,0). Thus we see that TaT_{a}\mathcal{B} is isomorphic to \mathcal{B} or (C)\mathcal{B}(C), respectively, by Lemma 4.1 and Lemma 4.2. Thus we conclude the proof. ∎

Corollary 5.3.

Let (,)𝖤𝖢2(Σ2)(\mathcal{B},\mathcal{E})\in\mathsf{EC}_{2}(\Sigma_{2}) be an exceptional pair such that \mathcal{B} is a vector bundle and ()>0\ell(\mathcal{E})>0.

  1. (1)

    Suppose that torsE()\operatorname{tors}E(\mathcal{E}) is non-zero and is a direct sum of copies of 𝒪C(a)\mathcal{O}_{C}(a). Then (Ta())<()\ell(T_{a}(\mathcal{E}))<\ell(\mathcal{E}) and Ta()T_{a}(\mathcal{B}) is isomorphic to either \mathcal{B} or (C)\mathcal{B}(C); in particular, it is a vector bundle.

  2. (2)

    Suppose that torsE()=0\operatorname{tors}E(\mathcal{E})=0, so that torsE()\operatorname{tors}E(\mathcal{E}^{\vee}) is non-zero by Lemma 3.26. Suppose that it is a direct sum of copies of 𝒪C(a)\mathcal{O}_{C}(a^{\prime}). Then (Ta3())<()\ell(T_{-a^{\prime}-3}(\mathcal{E}))<\ell(\mathcal{E}) and Ta3()T_{-a^{\prime}-3}(\mathcal{B}) is isomorphic to either \mathcal{B} or (C)\mathcal{B}(C); in particular, it is a vector bundle.

In particular, there is c=c()c=c(\mathcal{E})\in\mathbb{Z} which depends only on \mathcal{E}, independent of \mathcal{B} in particular, such that (Tc)<()\ell(T_{c}\mathcal{E})<\ell(\mathcal{E}) and TcT_{c}\mathcal{B} is isomorphic to either \mathcal{B} or (C)\mathcal{B}(C).

Proof.

The assertions on the lengths are already proven in Theorem 3.32.

Let us first assume torsE()0\operatorname{tors}E(\mathcal{E})\neq 0. The case where \mathcal{E} is not isomorphic to a shift of a sheaf is settled in Lemma 5.1. The case \mathcal{E} is isomorphic to a shift of a sheaf is settled in Lemma 5.2 (we may assume i0=0i_{0}=0, without loss of generality).

If torsE()=0\operatorname{tors}E(\mathcal{E})=0, we reduce the proof to the first case. In fact, it follows from Lemma 3.26 that torsE()0\operatorname{tors}E(\mathcal{E}^{\vee})\neq 0. Note that

((KΣ2),)\displaystyle\left(\mathcal{B}^{\vee}(K_{\Sigma_{2}}),\mathcal{E}^{\vee}\right) (5.14)

is also an exceptional pair such that the first component is a vector bundle and ()=Lemma 3.29()>0\ell(\mathcal{E}^{\vee})\stackrel{{\scriptstyle\text{Lemma~{}\ref{lm:length is invariant under dual}}}}{{=}}\ell(\mathcal{E})>0. Suppose that torsE()\operatorname{tors}E(\mathcal{E}^{\vee}) is a direct sum of copies of 𝒪C(a)\mathcal{O}_{C}(a^{\prime}). Then, by applying the conclusion of the previous paragraph to the pair (5.14), it follows that Ta((KΣ2))T_{a^{\prime}}(\mathcal{B}^{\vee}(K_{\Sigma_{2}})) is isomorphic to either (KΣ2)\mathcal{B}^{\vee}(K_{\Sigma_{2}}) or (KΣ2+C)\mathcal{B}^{\vee}(K_{\Sigma_{2}}+C). Then from the following computation we see that T3a()T_{-3-a^{\prime}}(\mathcal{B}) is isomorphic to either \mathcal{B} or (C)\mathcal{B}(C).

Ta((KΣ2))Lemma 2.3(Ta())(KΣ2)Lemma 2.4 (2.21)(T2a())(KΣ2)\displaystyle T_{a^{\prime}}(\mathcal{B}^{\vee}(K_{\Sigma_{2}}))\stackrel{{\scriptstyle\text{Lemma~{}\ref{lm:conjugation of spherical twist}}}}{{\simeq}}\left(T_{a^{\prime}}(\mathcal{B}^{\vee})\right)(K_{\Sigma_{2}})\stackrel{{\scriptstyle\text{Lemma~{}\ref{lm:spherical twist and dual} \eqref{eq:dual and spherical twists}}}}{{\simeq}}\left(T^{\prime}_{-2-a^{\prime}}(\mathcal{B})\right)^{\vee}(K_{\Sigma_{2}}) (5.15)
(2.28)(T1a((C)))(KΣ2)Lemma 2.3(T3a())(KΣ2+C)\displaystyle\stackrel{{\scriptstyle\eqref{eq:square of inverse twists as square of twists and O(-2C)}}}{{\simeq}}\left(T_{-1-a^{\prime}}\left(\mathcal{B}(-C)\right)\right)^{\vee}(K_{\Sigma_{2}})\stackrel{{\scriptstyle\text{Lemma~{}\ref{lm:conjugation of spherical twist}}}}{{\simeq}}\left(T_{-3-a^{\prime}}(\mathcal{B})\right)^{\vee}(K_{\Sigma_{2}}+C) (5.16)

Theorem 5.4.

For any NN\in\mathbb{Z} with 1N41\leq N\leq 4 and any exceptional collection ¯=(1,2,,N)𝖤𝖢N(Σ2)\underline{\mathcal{E}}=(\mathcal{E}_{1},\mathcal{E}_{2},\dots,\mathcal{E}_{N})\in\mathsf{EC}_{N}(\Sigma_{2}), there exists a product of spherical twists of the form TaT_{a} for some aa\in\mathbb{Z}, denoted by bb such that b(¯)N𝖤𝖢𝖵𝖡N(Σ2)b(\underline{\mathcal{E}})\in\mathbb{Z}^{N}\cdot\mathsf{ECVB}_{N}(\Sigma_{2}).

Proof.

Without loss of generality, we may and will assume i0(i)=0i_{0}(\mathcal{E}_{i})=0 for all i=1,,Ni=1,\dots,N. We prove the assertion by an induction on NN.

The case when N=1N=1 is nothing but Theorem 3.1. Consider the case when N>1N>1. By applying the induction hypothesis to the subcollection (1,,N1)\left(\mathcal{E}_{1},\dots,\mathcal{E}_{N-1}\right), we may and will assume that these are already vector bundles, say, (F1,,FN1)\left(F_{1},\dots,F_{N-1}\right). Suppose N\mathcal{E}_{N} is not a vector bundle; i.e., (N)>0\ell(\mathcal{E}_{N})>0. Otherwise there is nothing to show. In this case, since (Fi,N)𝖤𝖢2(Σ2)(F_{i},\mathcal{E}_{N})\in\mathsf{EC}_{2}(\Sigma_{2}) for each 1iN11\leq i\leq N-1, if we take c=c(N)c=c(\mathcal{E}_{N}) as in Corollary 5.3, then TcFiT_{c}F_{i} remains to be a vector bundle for all i=1,,N1i=1,\dots,N-1 and it holds that (TcN)<(N)\ell(T_{c}\mathcal{E}_{N})<\ell(\mathcal{E}_{N}). By repeating this process until (N)\ell(\mathcal{E}_{N}) reaches 0, we achieve our goal. ∎

The constructibility for exceptional collections consisting of vector bundles is shown in [Kul97, Theorem 3.1.8.2] for a class of weak del Pezzo surfaces. Though Σ2\Sigma_{2} is not contained in the class, we can deduce the same assertion for Σ2\Sigma_{2} from it:

Theorem 5.5.

Any exceptional collection on Σ2\Sigma_{2} consisting of vector bundles can be extended to a full exceptional collection.

Proof.

Let

π:YΣ2\displaystyle\pi\colon Y\to\Sigma_{2} (5.17)

be the blowup of Σ2\Sigma_{2} in a point outside of the curve CC. Let EYE\subset Y be the exceptional curve.

Take ¯(1,,N)𝖤𝖢𝖵𝖡N(Σ2)\underline{\mathcal{E}}\coloneqq\left(\mathcal{E}_{1},\dots,\mathcal{E}_{N}\right)\in\mathsf{ECVB}_{N}(\Sigma_{2}), so that (𝒪E(1),π¯)𝖤𝖢N+1(Y).\left(\mathcal{O}_{E}(-1),\pi^{\ast}\underline{\mathcal{E}}\right)\in\mathsf{EC}_{N+1}(Y). Write Fπ1F\coloneqq\pi^{\ast}\mathcal{E}_{1}, and consider the right mutation RF𝒪E(1)R_{F}\mathcal{O}_{E}(-1). We claim that this is an exceptional vector bundle.

To see this, note first that the semiorthogonality HomY(F,𝒪E(1))=0\mathop{\mathbb{R}\mathrm{Hom}}\nolimits_{Y}(F,\mathcal{O}_{E}(-1))=0 implies F|E𝒪ErF|_{E}\simeq\mathcal{O}_{E}^{\oplus r}, where r=rankFr=\operatorname{rank}F. Thus we obtain the following short exact sequence.

0FrRF𝒪E(1)𝒪E(1)0.\displaystyle 0\to F^{\oplus r}\to R_{F}\mathcal{O}_{E}(-1)\to\mathcal{O}_{E}(-1)\to 0. (5.18)

Suppose for a contradiction that the torsion part of the exceptional sheaf RF𝒪E(1)R_{F}\mathcal{O}_{E}(-1) is nontrivial. Then it should map injectively to 𝒪E(1)\mathcal{O}_{E}(-1). It then implies that the canonical morphism from the locally free sheaf FrF^{\oplus r} to the torsion free part of RF𝒪E(1)R_{F}\mathcal{O}_{E}(-1) is both injective and surjective in codimension 1, which hence is an isomorphism. This contradicts the indecomposability of RF𝒪E(1)R_{F}\mathcal{O}_{E}(-1).

Thus we have obtained (F,RF𝒪E(1),π2,,πN)𝖤𝖢𝖵𝖡N+1(Y)\left(F,R_{F}\mathcal{O}_{E}(-1),\pi^{\ast}\mathcal{E}_{2},\dots,\pi^{\ast}\mathcal{E}_{N}\right)\in\mathsf{ECVB}_{N+1}(Y). By [Kul97, Theorem 3.1.8.2], it extends to a full exceptional collection on YY (note that YY is a weak del Pezzo surface obtained by blowing up 2\mathbb{P}^{2} first in a point and then in a point on the (1)(-1)-curve, (hence) that |KY||-K_{Y}| is base point free and KY2=7>1K_{Y}^{2}=7>1). By applying the left mutation again, we obtain a full exceptional collection (𝒪E(1),π1,,πN,N+1,,4)𝖥𝖤𝖢(Y)\left(\mathcal{O}_{E}(-1),\pi^{\ast}\mathcal{E}_{1},\dots,\pi^{\ast}\mathcal{E}_{N},\mathcal{E}^{\prime}_{N+1},\dots,\mathcal{E}^{\prime}_{4}\right)\in\mathsf{FEC}(Y). Since N+1,,4𝒪E(1)\mathcal{E}^{\prime}_{N+1},\dots,\mathcal{E}^{\prime}_{4}\in{}^{\perp}\langle\mathcal{O}_{E}(-1)\rangle, there are some objects N+1,,4𝐃(Σ2)\mathcal{E}_{N+1},\dots,\mathcal{E}_{4}\in\mathbf{D}(\Sigma_{2}) such that N+1𝕃πN+1,,4𝕃π4\mathcal{E}^{\prime}_{N+1}\simeq\mathbb{L}\pi^{\ast}\mathcal{E}_{N+1},\dots,\mathcal{E}^{\prime}_{4}\simeq\mathbb{L}\pi^{\ast}\mathcal{E}_{4}. Since 𝕃π:𝐃(Σ2)𝒪E(1)\mathbb{L}\pi^{\ast}\colon\mathbf{D}(\Sigma_{2})\to{}^{\perp}\langle\mathcal{O}_{E}(-1)\rangle is an equivalence, we see that 1,,N,N+1,,4\mathcal{E}_{1},\dots,\mathcal{E}_{N},\mathcal{E}_{N+1},\dots,\mathcal{E}_{4} is a full exceptional collection of 𝐃(Σ2)\mathbf{D}(\Sigma_{2}). ∎

We finally obtain the following constructibility theorem for Σ2\Sigma_{2}.

Corollary 5.6.

Any exceptional collection on Σ2\Sigma_{2} can be extended to a full exceptional collection.

Proof.

For any exceptional collection ¯𝖤𝖢N(Σ2)\underline{\mathcal{E}}\in\mathsf{EC}_{N}(\Sigma_{2}), by Theorem 5.4, there exists bBb\in B such that b(¯)N𝖤𝖢𝖵𝖡N(Σ2)b\left(\underline{\mathcal{E}}\right)\in\mathbb{Z}^{N}\cdot\mathsf{ECVB}_{N}(\Sigma_{2}). Then by Theorem 5.5, b(¯)b\left(\underline{\mathcal{E}}\right) can be extended to a full exceptional collection on Σ2\Sigma_{2}. Applying b1b^{-1} to the extended collection, one obtains the desired full exceptional collection which extends ¯\underline{\mathcal{E}}. ∎

Remark 5.7.

Contrary to Theorem 4.3, an exceptional collection (consisting of objects of rank>0\operatorname{rank}>0) of length at least 22 is not necessarily numerically equivalent to an exceptional collection of vector bundles. Namely, for each N=2,3,4,N=2,3,4, there is an exceptional collection (1,,N)𝖤𝖢N(Σ2)(\mathcal{E}_{1},\dots,\mathcal{E}_{N})\in\mathsf{EC}_{N}(\Sigma_{2}) of length NN and ranki>0\operatorname{rank}\mathcal{E}_{i}>0 for i=1,,Ni=1,\dots,N for which there is no exceptional collection of vector bundles (F1,,FN)𝖤𝖢𝖵𝖡N(Σ2)(F_{1},\dots,F_{N})\in\mathsf{ECVB}_{N}(\Sigma_{2}) such that [i]=[Fi]K0(Σ2)[\mathcal{E}_{i}]=[F_{i}]\in\operatorname{K_{0}}\left(\Sigma_{2}\right) for i=1,,Ni=1,\dots,N. Example 5.8 below is such an example for N=2N=2. Examples for N=3,4N=3,4 are obtained by extending examples of length 22 by Corollary 5.6. For these exceptional collections, in particular, in Theorem 5.4 one can not take bb from BK0𝗍𝗋𝗂𝗏B^{K_{0}-\mathsf{triv}}. This is in contrast to Theorem 6.2.

Example 5.8.

Consider the following exceptional pair.

¯(1,2)T1(𝒪Σ2,𝒪Σ2(C+4f))𝖤𝖢2(Σ2).\underline{\mathcal{E}}\coloneqq(\mathcal{E}_{1},\mathcal{E}_{2})\coloneqq T_{-1}\left(\mathcal{O}_{\Sigma_{2}},\mathcal{O}_{\Sigma_{2}}(C+4f)\right)\in\mathsf{EC}_{2}(\Sigma_{2}).

Then there is no exceptional pair of vector bundles (F1,F2)𝖤𝖢𝖵𝖡2(Σ2)(F_{1},F_{2})\in\mathsf{ECVB}_{2}(\Sigma_{2}) such that [i]=[Fi]K0(Σ2)[\mathcal{E}_{i}]=[F_{i}]\in\operatorname{K_{0}}\left(\Sigma_{2}\right) for i=1,2i=1,2 for the following reason.

Note first that 1𝒪Σ2\mathcal{E}_{1}\simeq\mathcal{O}_{\Sigma_{2}} and 2\mathcal{E}_{2} is a sheaf with tors2=𝒪C(2)\operatorname{tors}\mathcal{E}_{2}=\mathcal{O}_{C}(-2) by [OU15, Theorem 1.4]. Suppose that there is a pair (F1,F2)(F_{1},F_{2}) as above. Then F1𝒪Σ2F_{1}\simeq\mathcal{O}_{\Sigma_{2}} by Lemma 2.27, and since

[F2]=[2]=[𝒪Σ2(C+4f)]χ(𝒪Σ2(C+4f),𝒪C(1))[𝒪C(1)]=[𝒪Σ2(3C+4f)],[F_{2}]=[\mathcal{E}_{2}]=[\mathcal{O}_{\Sigma_{2}}(C+4f)]-\chi(\mathcal{O}_{\Sigma_{2}}(C+4f),\mathcal{O}_{C}(-1))\left[\mathcal{O}_{C}(-1)\right]=[\mathcal{O}_{\Sigma_{2}}(3C+4f)],

it follows that F2𝒪Σ2(3C+4f)F_{2}\simeq\mathcal{O}_{\Sigma_{2}}(3C+4f) again by Lemma 2.27. This, however, leads to the following contradiction.

0=ExtΣ22(F2,F1)=ExtΣ22(𝒪Σ2(3C+4f),𝒪Σ2)0\displaystyle 0=\operatorname{Ext}_{\Sigma_{2}}^{2}(F_{2},F_{1})=\operatorname{Ext}^{2}_{\Sigma_{2}}(\mathcal{O}_{\Sigma_{2}}(3C+4f),\mathcal{O}_{\Sigma_{2}})\neq 0 (5.19)
Remark 5.9.

Given an exceptional collection ¯=(1,,N)𝖤𝖢N(Σ2)\underline{\mathcal{E}}=\left(\mathcal{E}_{1},\dots,\mathcal{E}_{N}\right)\in\mathsf{EC}_{N}(\Sigma_{2}) such that ranki>0\operatorname{rank}\mathcal{E}_{i}>0 for all ii, by Corollary 4.4 (1) there is a unique sequence of exceptional vector bundles F1,,FNF_{1},\dots,F_{N} such that [i]=[Fi]K0(Σ2)[\mathcal{E}_{i}]=[F_{i}]\in\operatorname{K_{0}}\left(\Sigma_{2}\right) for all i=1,,Ni=1,\dots,N. Example 5.8 implies that (F1,,FN)(F_{1},\dots,F_{N}) is not necessarily an exceptional collection.

It also implies that the map 𝗀𝖾𝗇|𝖤𝖢𝖵𝖡N(Σ2):𝖤𝖢𝖵𝖡N(Σ2)𝖤𝖢N(1×1)\operatorname{\mathsf{gen}}|_{\mathsf{ECVB}_{N}(\Sigma_{2})}\colon\mathsf{ECVB}_{N}(\Sigma_{2})\to\mathsf{EC}_{N}(\mathbb{P}^{1}\times\mathbb{P}^{1}) (the case N=4N=4 appears in Figure 2.66) is not surjective for N=2,3,4N=2,3,4, though it is for N=1N=1 by [OU15, Lemma 4.6] and Proposition 2.32. In fact, let ¯=(1,,N)𝖤𝖢N(Σ2)\underline{\mathcal{E}}=(\mathcal{E}_{1},\dots,\mathcal{E}_{N})\in\mathsf{EC}_{N}(\Sigma_{2}) be an exceptional collection (consisting of objects of rank>0\operatorname{rank}>0) which is not numerically equivalent to an exceptional collection of vector bundles. Then 𝗀𝖾𝗇(¯)\operatorname{\mathsf{gen}}(\underline{\mathcal{E}}) is not in the image of 𝗀𝖾𝗇|𝖤𝖢𝖵𝖡N(Σ2)\operatorname{\mathsf{gen}}|_{\mathsf{ECVB}_{N}(\Sigma_{2})}. In fact, an exceptional collection of vector bundles ¯=(F1,,FN)𝖤𝖢𝖵𝖡N(Σ2)\underline{\mathcal{F}}=(F_{1},\dots,F_{N})\in\mathsf{ECVB}_{N}(\Sigma_{2}) such that 𝗀𝖾𝗇(¯)=𝗀𝖾𝗇(¯)𝖤𝖢N(1×1)\operatorname{\mathsf{gen}}(\underline{\mathcal{F}})=\operatorname{\mathsf{gen}}(\underline{\mathcal{E}})\in\mathsf{EC}_{N}(\mathbb{P}^{1}\times\mathbb{P}^{1}) must satisfy [i]=[Fi]K0(Σ2)[\mathcal{E}_{i}]=[F_{i}]\in\operatorname{K_{0}}\left(\Sigma_{2}\right) for i=1,,Ni=1,\dots,N, which contradicts the choice of ¯\underline{\mathcal{E}}.

Despite Example 5.8, by Corollary 2.11, one can always bring an arbitrary exceptional collection (consisting of objects of rank>0\operatorname{rank}>0) to an exceptional collection of vector bundles by an element of BK0𝗍𝗋𝗂𝗏B^{K_{0}-\mathsf{triv}} up to a twist by T0T_{0}.

Theorem 5.10.

Let ¯=(1,,N)𝖤𝖢N(Σ2)\underline{\mathcal{E}}=(\mathcal{E}_{1},\dots,\mathcal{E}_{N})\in\mathsf{EC}_{N}(\Sigma_{2}) be an exceptional collection with 2N42\leq N\leq 4 and ranki>0\operatorname{rank}\mathcal{E}_{i}>0 for all i=1,,Ni=1,\dots,N. Then there exists bBK0𝗍𝗋𝗂𝗏b\in B^{K_{0}-\mathsf{triv}} such that b(¯)(2)N(𝖤𝖢𝖵𝖡N(Σ2)T0(𝖤𝖢𝖵𝖡N(Σ2)))b(\underline{\mathcal{E}})\in(2\mathbb{Z})^{N}\cdot\left(\mathsf{ECVB}_{N}(\Sigma_{2})\cup T_{0}\left(\mathsf{ECVB}_{N}(\Sigma_{2})\right)\right).

Proof.

We may assume without loss of generality that i0=0i_{0}=0 for any member of the collection ¯\underline{\mathcal{E}}. By Theorem 5.4, there is bBb\in B such that b(¯)𝖤𝖢𝖵𝖡N(Σ2)b(\underline{\mathcal{E}})\in\mathsf{ECVB}_{N}(\Sigma_{2}). Now the assertion immediately follows from the general description of elements of BB given in (2.34) and (2.35) (note that T0=T0(T02)T_{0}=T^{\prime}_{0}(T_{0}^{2})). ∎

Remark 5.11.

Here we give some speculations on the spaces of Bridgeland stability conditions and a resulting question.

To start with, it is conceivable that there is a local homeomorphism φ\varphi as in the following commutative diagram whose restriction to the subspaces of algebraic stability conditions is compatible with the generalization map

𝗀𝖾𝗇:𝖥𝖲𝖤𝖢(Σ2)𝖥𝖲𝖤𝖢(1×1),\displaystyle\operatorname{\mathsf{gen}}\colon\mathsf{FSEC}(\Sigma_{2})\to\mathsf{FSEC}(\mathbb{P}^{1}\times\mathbb{P}^{1}), (5.20)

where 𝖥𝖲𝖤𝖢()\mathsf{FSEC}(\bullet) denote the set of isomorphism classes of full strong exceptional collections on \bullet (recall that a full strong exceptional collection yields a chamber of algebraic stability conditions in Stab()\operatorname{Stab}(\bullet)). The fact that 𝗀𝖾𝗇\operatorname{\mathsf{gen}} restricts to the sets of strong exceptional collections follows from Remark 2.19 and Corollary 2.17.

Stab(Σ2){\operatorname{Stab}(\Sigma_{2})}Stab(1×1){\operatorname{Stab}(\mathbb{P}^{1}\times\mathbb{P}^{1})}Hom(K0(Σ2),){\operatorname{Hom}(\operatorname{K_{0}}\left(\Sigma_{2}\right),\mathbb{C})}Hom(K0(1×1),){\operatorname{Hom}(\operatorname{K_{0}}\left(\mathbb{P}^{1}\times\mathbb{P}^{1}\right),\mathbb{C})}φ\scriptstyle{\varphi}p\scriptstyle{p}q\scriptstyle{q}\scriptstyle{\sim}(K0(𝗀𝖾𝗇)1)\scriptstyle{(\operatorname{K_{0}}\left(\operatorname{\mathsf{gen}}\right)^{-1})^{\ast}} (5.21)

Conjecturally, the Galois group (== the group of fiber-preserving automorphisms of Stab(Σ2)\operatorname{Stab}(\Sigma_{2})) of pp coincides with BK0𝗍𝗋𝗂𝗏×2B^{K_{0}-\mathsf{triv}}\times 2\mathbb{Z}. As BK0𝗍𝗋𝗂𝗏B^{K_{0}-\mathsf{triv}} do not deform to 1×1\mathbb{P}^{1}\times\mathbb{P}^{1}, it is conceivable that BK0𝗍𝗋𝗂𝗏B^{K_{0}-\mathsf{triv}} coincides with the Galois group of φ\varphi. Therefore it seems reasonable to ask the following question, which is an analogue of Corollary 4.4 (2). Unfortunately, Theorem 5.10 is not strong enough to answer it in the affirmative.

Question 5.12.

For each ¯𝖥𝖲𝖤𝖢(Σ2)\underline{\mathcal{E}}\in\mathsf{FSEC}(\Sigma_{2}), the action of the group BK0𝗍𝗋𝗂𝗏×2B^{K_{0}-\mathsf{triv}}\times 2\mathbb{Z} on the following set is transitive.

{¯𝖥𝖲𝖤𝖢(Σ2)[¯]=[¯]K0(Σ2)4}\displaystyle\left\{\underline{\mathcal{E}}^{\prime}\in\mathsf{FSEC}(\Sigma_{2})\mid[\underline{\mathcal{E}}^{\prime}]=[\underline{\mathcal{E}}]\in\operatorname{K_{0}}\left(\Sigma_{2}\right)^{4}\right\} (5.22)

6. Braid group acts transitively on the set of full exceptional collections

This section is devoted to the proof of the following theorem.

Theorem 6.1.

The action G4𝖤𝖢4(Σ2)G_{4}\curvearrowright\mathsf{EC}_{4}(\Sigma_{2}) is transitive. Namely, Conjecture 1.2 holds true for Σ2\Sigma_{2}.

The proof is divided into 3 steps. Let ¯𝖤𝖢4(Σ2)\underline{\mathcal{E}}\in\mathsf{EC}_{4}(\Sigma_{2}) be the given exceptional collection of length 44.

Step 1.

By Corollary 2.38, there exists σG4\sigma\in G_{4} such that

[σ(¯)]=[𝗌𝗍𝖽]𝗇𝗎𝗆𝖥𝖤𝖢(Σ2),\displaystyle[\sigma(\underline{\mathcal{E}})]=[\mathcal{E}^{\mathsf{std}}]\in\mathsf{numFEC}(\Sigma_{2}), (6.1)

where 𝗌𝗍𝖽\mathcal{E}^{\mathsf{std}} is the standard full exceptional collection defined in (2.78). Recall that there might be a difference between σ(¯)\sigma(\underline{\mathcal{E}}) and 𝗌𝗍𝖽\mathcal{E}^{\mathsf{std}} which is invisible on the numerical level. The rest of the proof is devoted to killing this (possible) difference.

Step 2.

Next, we show the following theorem.

Theorem 6.2.

For any ¯𝖤𝖢4(Σ2)\underline{\mathcal{E}}\in\mathsf{EC}_{4}(\Sigma_{2}) satisfying

[¯]=[𝗌𝗍𝖽]𝗇𝗎𝗆𝖥𝖤𝖢(Σ2),\displaystyle[\underline{\mathcal{E}}]=[\mathcal{E}^{\mathsf{std}}]\in\mathsf{numFEC}(\Sigma_{2}), (6.2)

there exists bBK0𝗍𝗋𝗂𝗏b\in B^{K_{0}-\mathsf{triv}} such that b(¯)𝗌𝗍𝖽b(\underline{\mathcal{E}})\simeq\mathcal{E}^{\mathsf{std}}.

Proof of Theorem 6.2.

Write ¯=(1,2,3,4)\underline{\mathcal{E}}=\left(\mathcal{E}_{1},\mathcal{E}_{2},\mathcal{E}_{3},\mathcal{E}_{4}\right). We may and will assume that i0=0i_{0}=0 for all of the objects in the collection. By Corollary 4.4 (2), there exists bBK0𝗍𝗋𝗂𝗏b\in B^{K_{0}-\mathsf{triv}} such that b(1)𝒪Σ2b(\mathcal{E}_{1})\simeq\mathcal{O}_{\Sigma_{2}}. Hence by replacing ¯\underline{\mathcal{E}} with b(¯)b(\underline{\mathcal{E}}), we may and will assume that 1=𝒪Σ2\mathcal{E}_{1}=\mathcal{O}_{\Sigma_{2}}.

Next, by Corollary 5.3, there is bBb\in B such that b(𝒪Σ2)b(\mathcal{O}_{\Sigma_{2}}) and b(2)b(\mathcal{E}_{2}) are both vector bundles. Recall that bb is like either (2.34) or (2.35). If bb is like (2.34), then it follows that both b0(𝒪Σ2)b_{0}(\mathcal{O}_{\Sigma_{2}}) and b0(2)b_{0}(\mathcal{E}_{2}) are line bundles for some b0BK0𝗍𝗋𝗂𝗏b_{0}\in B^{K_{0}-\mathsf{triv}}. Since [b0(2)]=[2]=[𝒪Σ2(f)][b_{0}(\mathcal{E}_{2})]=[\mathcal{E}_{2}]=[\mathcal{O}_{\Sigma_{2}}(f)] and [b0(𝒪Σ2)]=[𝒪Σ2][b_{0}(\mathcal{O}_{\Sigma_{2}})]=[\mathcal{O}_{\Sigma_{2}}], it follows from Lemma 2.27 that b0(2)𝒪Σ2(f)b_{0}(\mathcal{E}_{2})\simeq\mathcal{O}_{\Sigma_{2}}(f) and b0(𝒪Σ2)𝒪Σ2b_{0}(\mathcal{O}_{\Sigma_{2}})\simeq\mathcal{O}_{\Sigma_{2}}.

If bb is like (2.35) for a0=1a_{0}=-1, then both T1b0(𝒪Σ2)T_{-1}b_{0}(\mathcal{O}_{\Sigma_{2}}) and T1b0(2)T_{-1}b_{0}(\mathcal{E}_{2}) are line bundles for some b0BK0𝗍𝗋𝗂𝗏b_{0}\in B^{K_{0}-\mathsf{triv}}. Then it follows from the following computations and Lemma 2.27 that T1b0(𝒪Σ2,2)=(𝒪Σ2,𝒪Σ2(C+f))T_{-1}b_{0}(\mathcal{O}_{\Sigma_{2}},\mathcal{E}_{2})=(\mathcal{O}_{\Sigma_{2}},\mathcal{O}_{\Sigma_{2}}(C+f)).

[T1b0(𝒪Σ2)]=[T1(𝒪Σ2)]=[𝒪Σ2],\displaystyle[T_{-1}b_{0}(\mathcal{O}_{\Sigma_{2}})]=[T_{-1}(\mathcal{O}_{\Sigma_{2}})]=[\mathcal{O}_{\Sigma_{2}}], (6.3)
[T1b0(2)]=[T1(2)]=[T1(𝒪Σ2(f))]=[𝒪Σ2(C+f)],\displaystyle[T_{-1}b_{0}(\mathcal{E}_{2})]=[T_{-1}(\mathcal{E}_{2})]=[T_{-1}(\mathcal{O}_{\Sigma_{2}}(f))]=[\mathcal{O}_{\Sigma_{2}}(C+f)], (6.4)

This immediately implies b0(𝒪Σ2,2)=(𝒪Σ2,𝒪Σ2(f))b_{0}(\mathcal{O}_{\Sigma_{2}},\mathcal{E}_{2})=(\mathcal{O}_{\Sigma_{2}},\mathcal{O}_{\Sigma_{2}}(f)). Hence we may and will assume i=i𝗌𝗍𝖽\mathcal{E}_{i}=\mathcal{E}^{\mathsf{std}}_{i} for i=1,2i=1,2.

At this point, in fact, we are done. To see this, note that both (3,4)(\mathcal{E}_{3},\mathcal{E}_{4}) and (3𝗌𝗍𝖽,4𝗌𝗍𝖽)(\mathcal{E}^{\mathsf{std}}_{3},\mathcal{E}^{\mathsf{std}}_{4}) are exceptional pairs of the triangulated subcategory 1𝗌𝗍𝖽,2𝗌𝗍𝖽𝐃(Σ2){}^{\perp}\langle\mathcal{E}^{\mathsf{std}}_{1},\mathcal{E}^{\mathsf{std}}_{2}\rangle\subset\mathbf{D}(\Sigma_{2}), which is equivalent to 𝐃(1)\mathbf{D}(\mathbb{P}^{1}), satisfying [i]=[i𝗌𝗍𝖽]K0(1𝗌𝗍𝖽,2𝗌𝗍𝖽)K0(𝐃(Σ2))[\mathcal{E}_{i}]=[\mathcal{E}^{\mathsf{std}}_{i}]\in K_{0}\left({}^{\perp}\langle\mathcal{E}^{\mathsf{std}}_{1},\mathcal{E}^{\mathsf{std}}_{2}\rangle\right)\hookrightarrow\operatorname{K_{0}}\left(\mathbf{D}(\Sigma_{2})\right) for i=3,4i=3,4. It is well known that any exceptional pair of 𝐃(1)\mathbf{D}(\mathbb{P}^{1}) is (up to shifts) of the form (𝒪1(a),𝒪1(a+1))\left(\mathcal{O}_{\mathbb{P}^{1}}(a),\mathcal{O}_{\mathbb{P}^{1}}(a+1)\right), hence is uniquely determined (up to shifts) by the class in the Grothendieck group. This immediately implies that ii𝗌𝗍𝖽\mathcal{E}_{i}\simeq\mathcal{E}^{\mathsf{std}}_{i} for i=3,4i=3,4, hence the conclusion. ∎

Step 3.

In the previous step, we killed the possible difference between ¯\underline{\mathcal{E}} and 𝗌𝗍𝖽\mathcal{E}^{\mathsf{std}} by spherical twists; more precisely, we found bBK0𝗍𝗋𝗂𝗏b\in B^{K_{0}-\mathsf{triv}} such that

¯=b(𝗌𝗍𝖽)\displaystyle\underline{\mathcal{E}}=b(\mathcal{E}^{\mathsf{std}}) (6.5)

(here we put bb on the right hand side intentionally). Recall that we wanted to kill the difference by a sequence of mutations and shifts, rather than spherical twists. In this last step, we confirm that bBb\in B in (6.5) can be replaced by a sequence of mutations. We begin with a lemma.

Lemma 6.3.

The following isomorphisms hold.

R𝒪Σ2(𝒪Σ2(C))T0𝒪Σ2L𝒪Σ2(f)𝒪Σ2(C+2f).\displaystyle R_{\mathcal{O}_{\Sigma_{2}}}(\mathcal{O}_{\Sigma_{2}}(-C))\simeq T_{0}\mathcal{O}_{\Sigma_{2}}\simeq L_{\mathcal{O}_{\Sigma_{2}}(f)}\mathcal{O}_{\Sigma_{2}}(C+2f). (6.6)
Proof.

By a direct computation, one can check that T0𝒪Σ2T_{0}\mathcal{O}_{\Sigma_{2}} is the cone of the (essentially) unique non-trivial morphism

𝒪C[2]𝒪Σ2.\displaystyle\mathcal{O}_{C}[-2]\to\mathcal{O}_{\Sigma_{2}}. (6.7)

The assertion immediately follows from this observation. ∎

Theorem 6.4.

For any bBb\in B, there exists σBr4\sigma\in\operatorname{Br}_{4} such that b(𝗌𝗍𝖽)=σ(𝗌𝗍𝖽)b(\mathcal{E}^{\mathsf{std}})=\sigma(\mathcal{E}^{\mathsf{std}}).

Proof.

By Theorem 2.12, bb is a product of copies of T1,T0T_{-1},T_{0} and their quasi-inverses. On the other hand, since autoequivalences and the action of the braid group commutes by Lemma 2.25, it is enough to show the assertion only for the two cases b=T1,T0b=T_{-1},T_{0}. We treat each case separately.

For T0T_{0}, we have

T0(𝗌𝗍𝖽)=(T0𝒪Σ2,𝒪Σ2(f),𝒪Σ2(C+2f)T0𝒪Σ2,𝒪Σ2(C+3f)).\displaystyle T_{0}(\mathcal{E}^{\mathsf{std}})=\left(T_{0}\mathcal{O}_{\Sigma_{2}},\mathcal{O}_{\Sigma_{2}}(f),\mathcal{O}_{\Sigma_{2}}(C+2f)\otimes T_{0}\mathcal{O}_{\Sigma_{2}},\mathcal{O}_{\Sigma_{2}}(C+3f)\right). (6.8)

By Lemma 6.3, one can easily verify that

(σ11σ2σ1)(𝗌𝗍𝖽)=T0(𝗌𝗍𝖽).\displaystyle\left(\sigma_{1}^{-1}\circ\sigma_{2}\circ\sigma_{1}\right)(\mathcal{E}^{\mathsf{std}})=T_{0}(\mathcal{E}^{\mathsf{std}}). (6.9)

For T1T_{-1}, we have

T1(𝗌𝗍𝖽)=(𝒪Σ2,𝒪Σ2(C+f),𝒪Σ2(C+2f),𝒪Σ2(2C+3f)).\displaystyle T_{-1}(\mathcal{E}^{\mathsf{std}})=\left(\mathcal{O}_{\Sigma_{2}},\mathcal{O}_{\Sigma_{2}}(C+f),\mathcal{O}_{\Sigma_{2}}(C+2f),\mathcal{O}_{\Sigma_{2}}(2C+3f)\right). (6.10)

By a direct computation, one can verify the following assertion.

(σ3σ2σ31)(𝗌𝗍𝖽)=((σ3σ2σ1)σ11σ31)(𝗌𝗍𝖽)=T1(𝗌𝗍𝖽).\displaystyle\left(\sigma_{3}\circ\sigma_{2}\circ\sigma_{3}^{-1}\right)(\mathcal{E}^{\mathsf{std}})=\left(\left(\sigma_{3}\circ\sigma_{2}\circ\sigma_{1}\right)\circ\sigma_{1}^{-1}\circ\sigma_{3}^{-1}\right)(\mathcal{E}^{\mathsf{std}})=T_{-1}(\mathcal{E}^{\mathsf{std}}). (6.11)

Below is an important consequence of Theorem 6.1.

Corollary 6.5.

𝖤𝖢4(Σ2)=𝖥𝖤𝖢(Σ2)\mathsf{EC}_{4}(\Sigma_{2})=\mathsf{FEC}(\Sigma_{2}).

Proof.

Since 𝗌𝗍𝖽\mathcal{E}^{\mathsf{std}} is full and the fullness is preserved under the action of the group G4G_{4}, this immediately follows from Theorem 6.1. ∎

References

  • [Bon89] A. I. Bondal. Representations of associative algebras and coherent sheaves. Izv. Akad. Nauk SSSR Ser. Mat., 53(1):25–44, 1989.
  • [Bon06] Juhani Bonsdorff. A Fourier transformation for Higgs bundles. J. Reine Angew. Math., 591:21–48, 2006.
  • [BOR20] Pieter Belmans, Shinnosuke Okawa, and Andrea T. Ricolfi. Moduli spaces of semiorthogonal decompositions in families. arXiv:2002.03303, February 2020.
  • [BP93] A. I. Bondal and A. E. Polishchuk. Homological properties of associative algebras: the method of helices. Izv. Ross. Akad. Nauk Ser. Mat., 57(2):3–50, 1993.
  • [CJ18] Pu Cao and Chen Jiang. Torsion exceptional sheaves on weak del Pezzo surfaces of type A. J. Algebra, 499:583–609, 2018.
  • [Dol12] Igor V. Dolgachev. Classical algebraic geometry. Cambridge University Press, Cambridge, 2012. A modern view.
  • [GK04] A. L. Gorodentsev and S. A. Kuleshov. Helix theory. Mosc. Math. J., 4(2):377–440, 535, 2004.
  • [GR87] A. L. Gorodentsev and A. N. Rudakov. Exceptional vector bundles on projective spaces. Duke Mathematical Journal, 54(1):115–130, 1987.
  • [Har66] Robin Hartshorne. Residues and duality. Lecture notes of a seminar on the work of A. Grothendieck, given at Harvard 1963/64. With an appendix by P. Deligne. Lecture Notes in Mathematics, No. 20. Springer-Verlag, Berlin-New York, 1966.
  • [Hir51] Friedrich Hirzebruch. Über eine Klasse von einfachzusammenhängenden komplexen Mannigfaltigkeiten. Math. Ann., 124:77–86, 1951.
  • [HL10] Daniel Huybrechts and Manfred Lehn. The geometry of moduli spaces of sheaves. Cambridge Mathematical Library. Cambridge University Press, Cambridge, second edition, 2010.
  • [HT10] Daniel Huybrechts and Richard P. Thomas. Deformation-obstruction theory for complexes via Atiyah and Kodaira-Spencer classes. Math. Ann., 346(3):545–569, 2010.
  • [Huy06] D. Huybrechts. Fourier-Mukai transforms in algebraic geometry. Oxford Mathematical Monographs. The Clarendon Press Oxford University Press, Oxford, 2006.
  • [IU05] Akira Ishii and Hokuto Uehara. Autoequivalences of derived categories on the minimal resolutions of AnA_{n}-singularities on surfaces. J. Differential Geom., 71(3):385–435, 2005.
  • [Kaw19] Kotaro Kawatani. Pure sheaves and Kleinian singularities. Manuscripta Math., 160(1-2):65–78, 2019.
  • [KO94] S. A. Kuleshov and D. O. Orlov. Exceptional sheaves on Del Pezzo surfaces. Izv. Ross. Akad. Nauk Ser. Mat., 58(3):53–87, 1994.
  • [Kod63] K. Kodaira. On stability of compact submanifolds of complex manifolds. Amer. J. Math., 85:79–94, 1963.
  • [Kul97] S. A. Kuleshov. Exceptional and rigid sheaves on surfaces with anticanonical class without base components. J. Math. Sci. (New York), 86(5):2951–3003, 1997. Algebraic geometry, 2.
  • [Kuz06] A. G. Kuznetsov. Hyperplane sections and derived categories. Izv. Ross. Akad. Nauk Ser. Mat., 70(3):23–128, 2006.
  • [Kuz11] Alexander Kuznetsov. Base change for semiorthogonal decompositions. Compos. Math., 147(3):852–876, 2011.
  • [Li17] Chunyi Li. The space of stability conditions on the projective plane. Selecta Math. (N.S.), 23(4):2927–2945, 2017.
  • [Lie06] Max Lieblich. Moduli of complexes on a proper morphism. J. Algebraic Geom., 15(1):175–206, 2006.
  • [LN07] Joseph Lipman and Amnon Neeman. Quasi-perfect scheme-maps and boundedness of the twisted inverse image functor. Illinois J. Math., 51(1):209–236, 2007.
  • [Orl97] D. O. Orlov. Equivalences of derived categories and K3K3 surfaces. J. Math. Sci. (New York), 84(5):1361–1381, 1997. Algebraic geometry, 7.
  • [OU15] Shinnosuke Okawa and Hokuto Uehara. Exceptional Sheaves on the Hirzebruch Surface 𝔽2\mathbb{F}_{2}. Int. Math. Res. Not. IMRN, (23):12781–12803, 2015.
  • [Pos95] Leonid Positselski. All strictly exceptional collections in Dcohb(Pm)D^{b}_{coh}(P^{m}) consist of vector bundles. alg-geom/9507014, July 1995.
  • [Rud88] A. N. Rudakov. Exceptional vector bundles on a quadric. Izv. Akad. Nauk SSSR Ser. Mat., 52(4):788–812, 896, 1988.
  • [ST01] Paul Seidel and Richard Thomas. Braid group actions on derived categories of coherent sheaves. Duke Math. J., 108(1):37–108, 2001.
  • [Sta16] The Stacks Project Authors. Stacks Project. http://stacks.math.columbia.edu, 2016.