Exceptional collections on
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 -reflections acting on the derived category.
This paper is devoted to solving the conjectures for the prototypical weak del Pezzo surface , the Hirzebruch surface of degree . 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 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 (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 be a smooth projective variety and be the bounded derived category of coherent sheaves on . Concerning the classification of the exceptional collections of , the prototypical results are given in [GR87, Section 5] for . The similar results for 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 be a del Pezzo surface over an algebraically closed field .
-
(1)
Any exceptional object is either a vector bundle or a line bundle on a -curve, up to shifts.
-
(2)
The isomorphism class of an exceptional object is determined by its class , up to shifts by .
-
(3)
Any exceptional collection of can be extended to a full exceptional collection.
-
(4)
Conjecture 1.2 below is true for ; i.e., the action is transitive.
The symbol denotes the Grothendieck group of the triangulated category . We briefly explain (4). Let denote the set of isomorphism classes of exceptional collections of length of , where . An important fact relevant to the classification is that there is a standard action of the group on by shifts and mutations, where is the braid group on 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 is a smooth projective variety such that admits a full exceptional collection of length . Then the action is transitive.
In dimensions greater than , the classification of exceptional collections is widely open. Even for , 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 whose anti-canonical bundle is nef and big. As it turns out, the generalization is not straightforward at all.
A weak del Pezzo surface which is not a del Pezzo surface admits at least one (and only finitely many) -curve(s); i.e., a smooth curve with and . Line bundles on , as objects of , are -spherical objects and hence yield non-trivial autoequivalences of called spherical twists due to Seidel and Thomas [ST01]. For a -spherical object , the corresponding spherical twist acts as a reflection on whereas it always satisfies . For example, the exceptional object has the same class as in but is not isomorphic to . In fact, by direct computation one can confirm the following.
(1.1) |
Moreover, by applying repeatedly, we obtain a collection of infinitely many exceptional objects of unbounded cohomological amplitudes which share the same class in . Hence Theorem 1.1 (1) (2) are not true for 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 ; namely, the Hirzebruch surface of degree , a paradigm of weak del Pezzo surfaces. In fact we formulated Conjecture 1.3 by generalizing the results we obtained for , Theorem 1.1, and [CJ18, Theorem 1.2] simultaneously.
Conjecture 1.3.
Let be a weak del Pezzo surface over an algebraically closed field .
-
(1)
For any exceptional object , there exists a sheaf on which is either an exceptional vector bundle or a line bundle on a -curve, a sequence of line bundles on -curves , and an integer such that
(1.2) -
(2)
For any pair of exceptional objects such that , there are a product of spherical twists and inverse spherical twists which acts trivially on and such that . Moreover, for each exceptional object , there is a unique exceptional vector bundle such that either or holds.
-
(3)
Any exceptional collection on can be extended to a full exceptional collection.
-
(4)
Conjecture 1.2 is true for .
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 .
More specifically, for
- •
- •
- •
- •
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.
- •
- •
-
•
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 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 of 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 of autoequivalences generated by spherical twists from [IU05]. Among others, we prove in Corollary 2.11 that , the subgroup of acting trivially on , is generated by squares of spherical twists by line bundles on .
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 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 there is a sequence of integers such that 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 . Among others we show that there is the index such that and for any the cohomology sheaf , if not , is a pure sheaf whose reduced support is . Then in Section 3.2, we prove that actually the schematic support of for is . 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 , where is an exceptional sheaf and is a torsion sheaf. We moreover show that if , then is a direct sum of copies of for some and that is a direct sum of copies of and . This integer plays a central role throughout the paper.
In Section 3.4 we investigate the relationship between the cohomology sheaves of and those of , the derived dual of . We in particular show in Corollary 3.27 that if is not isomorphic to a shift of a vector bundle and , then . In this paper we mainly discuss the case , and by this result we can settle the case where by passing to .
In Section 3.5 we introduce the notion of the length of the “torsion part” of an object at the generic point of . Formally speaking, for it is defined as . It follows that is isomorphic to a shift of a vector bundle if and only if , and hence it suffices to show that for some if . This is exactly what we achieve in Section 3.6. We show in Theorem 3.32 that works if and otherwise does.
Spherical objects on the minimal resolution of type 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 is never concentrated in the -curve . On the contrary the reduced support of a spherical object on is concentrated in , and it immediately implies that the schematic supports of the cohomology sheaves of the spherical object coincide with .
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 (Lemma 2.33).
Section 5 is devoted to the proof of Corollary 5.6. Take an exceptional collection . We first show in Theorem 5.4 that there is a product of spherical twists such that consists of vector bundles up to shifts. This is achieved in a one-by-one manner. The key is that if is an exceptional pair such that is a vector bundle and , then, surprisingly enough, remains to be a vector bundle (though it may not be isomorphic to ). Recall that depends only on and that strictly decreases the length of .
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 , we immediately obtain the proof of Corollary 5.6 for .
Section 6 is devoted to the proof of Theorem 6.1. We use the deformation of to . The difficulty is that there are infinitely many exceptional objects on which deform to the same exceptional object on (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 if and only if they have the same class in (up to shifts by ).
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 . Since deformation of exceptional collections commutes with mutations, the result for immediately implies that for any exceptional collection of length on there is a sequence of mutations which brings it to a collection which is numerically equivalent to the standard collection (i.e., having the same classes as in . See (2.78) for the definition of ).
It remains to show that an exceptional collection on which is numerically equivalent to can be sent to by mutations. In Step 2, as an intermediate step, we find such that . We construct such again in one-by-one manner.
In Step 3 we prove that 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 . Recall that generate . 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 . This is partly because the proof is rather involved already in this case. We nevertheless think that the case of 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 -curve; i.e., the fundamental cycle of the minimal resolution of the -singularity. So far such classification is achieved only for the minimal resolution of type 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 and type singularities. See [Kaw19] for results in this direction.
A weak del Pezzo surface over an algebraically closed field is isomorphic to either , or a blowup of 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 play an important role in showing the contractibility of (the main component of) the space of Bridgeland stability conditions in [Li17]. Our results should be similarly useful for studying . More specifically, they should be very closely related to the relationship between and ; see Remark 5.11 for details.
1.3. Notation and convention
We work over an algebraically closed field , unless otherwise stated. To ease notation, we will write and so on. Below is a list of frequently used symbols.
the Hirzebruch surface of degree (2.1) | |
the -curve of | |
the divisor class of a fiber of the morphism (2.4) | |
the derived dual of (2.6) | |
the (inverse) spherical twist by (2.15) | |
the group of autoequivalences generated by spherical twists (Definition 2.10) | |
various sets of exceptional collections (Definition 2.14) | |
(resp. ) | the braid group on strands (2.55) (resp. the extension of by (2.60)) |
the generalization map for exceptional collections from the central fiber to the generic fiber (2.65) | |
various sets of numerical exceptional collections (Definition 2.34) | |
the standard full exceptional collection of (Definition 2.36) | |
the standard full exceptional collection of (2.79) | |
the total (resp. -th) cohomology of with respect to the standard t-structure (3.3) | |
the unique index such that (Definition 3.7) | |
the complete local ring of the -singularity (Notation 3.10) | |
the maximal submodule of annihilated by the ideal (3.28) | |
the ideal sheaf of (3.30) | |
the dual -vector space of (3.36) | |
the canonical decomposition into an exceptional sheaf and a torsion sheaf (Lemma 3.21) | |
the torsion (resp. the torsion free) part of (Definition 3.22) | |
integers specified by the irreducible decomposition of (3.52) (see also Lemma 3.21) | |
“length of the torsion part of ” at the generic point of (Definition 3.28) | |
integers specified by the irreducible decomposition of an exceptional vector bundle restricted to (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 is the ruled surface
(2.1) |
The history of Hirzebruch surfaces goes back, at least, to the first paper by Hirzebruch [Hir51]. An explicit isotrivial degeneration of to for and is constructed in [Kod63, p.86 Example]. As a special case, there exists a smooth projective morphism (defined over )
(2.2) |
such that the fiber over is isomorphic to and the restriction of the family over the open subscheme is isomorphic to the trivial family .
In this paper we investigate the bounded derived category of coherent sheaves on , which is the most basic example of weak del Pezzo surfaces. We let denote the unique negative curve, and the (linear equivalence class of) the fiber of . Recall that
(2.3) |
where
(2.4) |
The anti-canonical bundle is given by
(2.5) |
2.2. Derived category, spherical twist, and the autoequivalence group of
Definition 2.1.
For a quasi-compact scheme , we let denote the perfect derived category of with the standard structure of a triangulated category. When is equipped with a morphism to , we think of as a triangulated -linear category. It comes with the natural symmetric monoidal structure given by the tensor product over , but we do not take it into account unless otherwise stated.
When is a smooth and projective variety over a field , we identify with the bounded derived category of coherent sheaves on .
The following equivalence of tensor triangulated categories
(2.6) |
will be called the derived dual.
One can easily verify that there exits a canonical natural isomorphism
(2.7) |
Definition 2.2.
For smooth projective varieties over and an object , the integral transform by the kernel will be denoted and defined as follows.
(2.8) |
Let be a smooth projective variety over a field . Recall that an object is spherical if and . The spherical twist by is the endofunctor
(2.9) |
defined by the kernel .
Consider the exchange automorphism
(2.10) |
Recall that
(2.11) |
is called the right adjoint kernel of and it enjoys the following adjoint property (see, say, [Huy06, Definition 5.7]).
(2.12) |
It follows that is an autoequivalence, so that . Note that there exists the obvious isomorphism
(2.13) |
so that
(2.14) |
A typical example of a spherical object on is for . The corresponding (inverse) spherical twist will be denoted as follows, for short.
(2.15) |
By definition, for each spherical object and any object , there are standard triangles as follows. To ease notation, we simply let denote the derived tensor product. The morphisms and are the evaluation and the coevaluation maps respectively, both of which are obtained from the standard adjoint pair of functors .
(2.16) | |||
(2.17) |
Lemma 2.3 ([IU05, Lemma 4.14]).
For each , there is an isomorphism of autoequivalences as follows.
(2.18) |
In particular, for any it holds that
(2.19) |
Proof.
Since is a smooth projective variety over , any autoequivalence of 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 is isomorphic to if and only if holds for any . 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)
For any spherical object , there exists a natural isomorphism of functors
(2.20) -
(2)
When , for any , there exists a natural isomorphism of functors
(2.21)
Proof.
There is a natural isomorphism
(2.22) | ||||
(2.23) | ||||
(2.24) | ||||
(2.25) | ||||
(2.26) |
where the isomorphism follows from [Har66, Chapter VII, Corollary 4.3 b)].
The second item follows from (2.20) and the isomorphism . ∎
Lemma 2.5 ([IU05, Lemma 4.15 (i) (2)]).
There is an isomorphism of autoequivalences
(2.27) |
which implies
(2.28) |
Definition 2.6.
For a triangulated category , we define the group of -trivial autoequivalences by the following exact sequence.
(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 , it holds that
(2.30) |
Proposition 2.8.
For any , there exists a sequence and such that
(2.31) |
Proof.
By (2.19), we may and will reduce the proof to that of the following claim.
Claim 2.9.
is contained in the subgroup of generated by and for all .
In the rest, we prove this claim. Consider first the case . We induct on . If it is , we have nothing to show. In general, we have
(2.32) |
By (2.27), the first term of the right hand side is isomorphic to . 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 . Then
(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 generated by spherical twists will be denoted by .
Corollary 2.11.
Fix any . Then elements of are exhausted by those of the following form, where .
(2.34) | |||
(2.35) |
Moreover, the normal subgroup
(2.36) |
is generated by .
Proof.
Lemma 2.3 implies that is a normal subgroup of . As we explain next, sets of generators of are well understood.
Theorem 2.12.
For any , the group is generated by the two spherical twists .
2.3. Deformation and mutation of exceptional collections
Let
(2.38) |
be a smooth projective morphism of Noetherian schemes with a closed point , and let
(2.39) |
be the central fiber. Note that the properness and smoothness of implies it is a perfect morphism, in the sense that the derived pushforward respects perfect complexes ([LN07, Proposition 2.1]).
Definition 2.13.
-
(1)
An object is -exceptional if is a line bundle on .
-
(2)
A collection of -exceptional objects is an -exceptional collection if for any with .
-
(3)
An -exceptional collection as above is said to be strong if moreover is isomorphic to a locally free sheaf (regarded as a complex concentrated in degree ) for any .
-
(4)
An -exceptional collection as above is said to be full if the minimal -linear (i.e., closed under by any ) triangulated subcategory which contains all of the objects in the collection is equivalent to .
Definition 2.14.
We say that two -exceptional collections and are isomorphic if for each . We will let (or , if is obvious from the context) denote the set of isomorphism classes of -exceptional collections of length . The set of isomorphism classes of -exceptional collections of length consisting entirely of locally free sheaves will be denoted by (or ), which comes with the obvious injection . Similarly, the set of equivalence classes of full -exceptional collections will be denoted by either or , and the set of isomorphism classes of full -exceptional collections consisting entirely of locally free sheaves will be denoted by (or ), which comes with the obvious injection .
Lemma 2.15.
-
(1)
Let be an -exceptional object. Then the functor
(2.40) is fully faithful and admits a right adjoint as follows.
(2.41) -
(2)
Let be an -exceptional collection. Then the smallest -linear triangulated subcategory of which contains admits a -linear semiorthogonal decomposition .
We say that a semiorthogonal decomposition is -linear if holds for any and (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 are perfect.
(2.42) |
Then the standard natural transformation of functors
(2.43) |
is an isomorphism if either or is flat.
Corollary 2.17.
Proof.
As is flat, the assertion immediately follows from the following computation. All functors are derived.
(2.45) |
∎
Lemma 2.18.
Suppose that for a complete local Noetherian ring . Then the natural restriction maps
(2.46) | |||
(2.47) |
obtained in Corollary 2.17 are bijections for any .
Proof.
Let us first show that any exceptional object deforms to an object . By definition of exceptional object, we know that . By the deformation theory of objects (see, say, [HT10, Corollary 3.4]), for each one finds the unique lift in of . Then it algebrizes uniquely to an actual object by [Lie06, Proposition 3.6.1].
We next show that is an -exceptional object. This is equivalent to the assertion , where is defined as the cone of the following standard morphism.
(2.48) |
Consider the following Cartesian diagram.
(2.49) |
By Lemma 2.16 it follows that
(2.50) |
so that . By Nakayama’s lemma, this implies that . The semiorthogonality of the collection is shown by similar arguments. ∎
Remark 2.19.
Definition 2.20.
Let be a morphism as in (2.38). An (-)exceptional pair is an (-)exceptional collection of length 2. For an -exceptional pair , the left mutation of through and the right mutation of through are defined by the following distinguished triangles.
(2.51) | |||
(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.
Next, let us recall a group action on from [BP93, Proposition 2.1]. Let be the braid group on strands, which admits the following famous presentation by generators and relations.
(2.55) | ||||
(2.56) |
Consider the action
(2.57) |
given by
(2.58) |
so that
(2.59) |
On the other hand, through the standard surjective homomorphism to the symmetric group of degree , the group acts naturally on the abelian group from the left. Let
(2.60) |
be the semi-direct product corresponding to the action.
One can verify that this, together with the action , where sends a collection to , extends to an action . In particular, one has the following induced action.
(2.61) |
When admits a full -exceptional collection of length , then one similarly obtains the action .
Remark 2.23.
Let be a smooth projective variety over the field . The autoequivalences of and the notion of exceptional collections are nicely compatible as we explain next.
Lemma 2.24.
If is an exceptional collection and , then so is . In particular, there is the natural action
(2.62) |
The following lemma is easy to verify and plays an important role in this paper.
2.4. Results obtained via deformation of to
Consider the (isotrivial) degeneration (2.2) over an algebraically closed field . Consider the discrete valuation ring and take the base change by of the family. We write
(2.63) |
by abuse of notation. The central fiber of is isomorphic to . Also, let
(2.64) |
be the geometric generic point of . The isotriviality of the family (2.2) outside the origin implies that the geometric generic fiber is isomorphic to over . Throughout this section, we freely use the symbols introduced in this paragraph.
Since the generic fiber 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)
Any exceptional object on is isomorphic to a shift of an exceptional vector bundle, so that the natural map is a bijection.
-
(2)
, and the action is transitive.
-
(3)
Any exceptional collection on can be extended to a full exceptional collection.
The non-triviality of the group implies that an exceptional object on is not uniquely determined by its class in , even modulo shifts by . 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 be exceptional vector bundles on such that . Then .
The degeneration (2.63) allows one to compare various invariants of to those of . Recall that .
Definition 2.28.
For an exceptional object , let denote the unique deformation of to and its restriction to the geometric generic fiber , respectively. For an exceptional collection on the central fiber, we will similarly write to mean its (unique) deformation to and its restriction to , respectively.
Let
(2.65) |
be the map which sends (an isomorphism class of) an exceptional collection of length 4 on to , which is obtained by restricting the deformation of to the -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 .
One similarly defines the map , to obtain the following diagram.
(2.66) |
We next compare of the surfaces. Note that we have the following diagram of schemes (the labels of the arrows in the diagram will be freely used).
(2.67) |
Applying the functor , we obtain the first two rows of Figure 2.68, which is a commutative diagram of commutative rings with units.
(2.68) |
The derived dual ∨ defined in (2.6) induces an automorphism of commutative rings
(2.69) |
The Euler pairing on (note that is a regular scheme) is the following bilinear pairing.
(2.70) |
We can similarly define
(2.71) | |||
(2.72) |
Lemma 2.29.
(2.73) | |||
(2.74) |
Let be a full exceptional collection of , and be the deformation of to and the restriction of to as defined in Definition 2.28. As pointed out in Remark 2.19, both are full exceptional collections.
Lemma 2.30.
are bases of , respectively. In particular, the horizontal maps in the first row of (2.68) are isomorphisms.
Proof.
Immediately follows from the fact that the collections are full exceptional collections of the triangulated categories over the base of length . ∎
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
(2.75) |
be the isomorphism of abelian groups obtained from the diagram Figure 2.68. Also, regard as -valued bilinear pairings by the diagram Figure 2.68. With all the preparations above, we can show the desired properties of the map .
Proposition 2.31.
The isomorphism of (2.75) respects the pairings on the source and the target abelian groups. Moreover, it fits in the following commutative diagram.
Proposition 2.32.
Let be exceptional objects. The following conditions are equivalent.
-
(1)
.
-
(2)
for some .
Proof.
For exceptional vector bundles on , we have the following reconstruction result. This is an immediate corollary of Lemma 2.27.
Lemma 2.33.
For any , the map is injective.
See Definition 2.34 for the definition of the set .
Definition 2.34.
Let be a smooth projective variety, and for simplicity let us assume that is isomorphic to the numerical Grothendieck group; i.e., the Euler pairing is non-degenerate. This in particular implies that is a free abelian group of finite rank.
An exceptional vector is an element such that .
A numerical exceptional collection on is a sequence of exceptional vectors such that for any . A numerical exceptional collection is said to be full if it is a basis of ; i.e, when .
The set of numerical exceptional collections of length (resp. full) on will be denoted by and , respectively.
For a numerical exceptional collection of length , 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.
(2.76) | |||
(2.77) |
By similar arguments for the action , one can verify that for a surface with there is an action , where the subgroup acts by the mutations (2.76) (2.77) and by the change of signs (hence the action descends to the quotient ).
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 .
Let us now introduce a particular (-)exceptional collection of invertible sheaves, which will be called the standard collection and serve as the base point of the sets and in this paper.
Definition 2.36.
The standard full exceptional collection of invertible sheaves on is defined as follows.
(2.78) |
By Lemma 2.18 and Remark 2.19, it uniquely deforms to a full -exceptional collection of invertible sheaves. We will write it , and its pullback to will be denoted by . Using the deformation invariance of the intersection numbers, one can easily confirm that
(2.79) |
under an isomorphism .
Corollary 2.37.
The map is surjective.
Proof.
Corollary 2.38.
For any , there exists such that .
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 , there exists an exceptional vector bundle and a sequence of integers such that
(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 be an integral noetherian scheme. For , we define
(3.2) |
and call it the schematic support of . It is universal among the closed subschemes which admits a coherent sheaf such that . The underlying closed subset of , or equivalently the reduced closed subscheme , is called the reduced support of . Also we let be the maximum torsion subsheaf of . For an object , we use the following notation.
(3.3) |
3.1. First properties of
As the first step toward the proof of Theorem 3.1, in this subsection we prove some basic properties of . Part of them concern , which can be summarized as follows.
There exists the unique integer with the following properties.
- •
- •
- •
for
Based on this, the similar statements for the schematic supports will be proved in the next subsection; namely, we will remove 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 . However, unfortunately, the proof of [IU05, Lemma 4.8] does not immediately apply to our situation. The fact prevents us from studying the problem locally around the curve .
Lemma 3.3.
The cohomology sheaves of an exceptional object enjoy the following properties.
-
(1)
is rigid ( in the spectral sequence (3.4)).
-
(2)
There exists a unique integer such that
-
•
.
-
•
for and are pure sheaves with , unless .
-
•
-
(3)
is rigid, and is an exceptional vector bundle.
See the following definition for the notion of rigidity.
Definition 3.4.
An object is said to be rigid if .
We will freely use the following standard fact on rigid objects.
Lemma 3.5.
Let be a rigid object. Then for any . In particular, for any such .
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 , there exists the following spectral sequence.
(3.4) |
Moreover, using the classes canonically determined by , the maps of (3.4) are given by
(3.5) |
Proof of Lemma 3.3.
Throughout the proof, we consider the spectral sequence (3.4) for the exceptional object . We prove the four items one by one.
(1) The exceptionality of is translated into the following conditions.
(3.6) |
Since is a smooth projective surface, unless . Hence (3.4) is -degenerate, and moreover is -degenerate at for any . Combined with (3.6), this implies that for any . In the next paragraph we also confirm , thereby concluding the rigidity of .
It follows from the explicit description (3.5) of maps that is in fact contained in , which implies that . Combined with the isomorphism from (3.6) and the epimorphism , this implies that and hence .
(2) We further obtain the following equalities from (3.6).
(3.7) | ||||
(3.8) |
To see (3.7) for , note that the arguments in the previous paragraph implies . Moreover, (1) implies that either . This follows from Lemma 3.5 and the orbit decomposition for the action .
Suppose for a contradiction that for every . Then coincides with and hence there is an isomorphism , from which we deduce by the Serre duality. This contradicts (3.6) and hence there must be at least one integer with .
Now consider the Serre duality
(3.9) |
Fix a non-trivial morphism such that the support of its cokernel is disjoint from the set of associated points of for all . This is possible, as the linear system 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 implies that for each the natural morphism
(3.10) |
is injective. Thus we obtain an injection of vector spaces
and hence an inequality
(3.11) |
for any .
Next we prove that is not surjective on the direct summands indexed by those with . To see this, (by slight abuse of notation) let be one of such indices and apply the functor to the following short exact sequence.
(3.12) |
The rigidity of , which we confirmed in (1), implies that we have the following short exact sequence.
(3.13) |
The 3rd term is not by the assumption and hence , which is identified with the direct summand of of interest, is not surjective.
Since we showed above that there is at least one such index , we have confirmed the non-surjectivity of and hence the following inequality.
(3.14) |
Summarizing the results so far, we obtain the following sequence of (in)equalities.
which implies that
(3.15) |
This means that the number of the indices with 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 for above, and for it is proven below.
(3) Put , and . Consider the spectral sequence
(3.16) |
arising from the short exact sequence , where and . For obvious reasons we see unless and .
On the other hand, since is rigid, it is stable under the action of by Lemma 3.5. Since the torsion part of a coherent sheaf is uniquely determined by the sheaf, it follows that is also stable under the same group action. Hence it follows that , so that
(3.17) |
Combined with the Serre duality, this implies the equality
(3.18) |
for .
It then follows that , where the first equality is the consequence of the fact that is torsion and is torsion free, and the second equality is the case of (3.18).
Thus we have confirmed that the spectral sequence (3.16) is -degenerate at . Hence , where the last vanishing is nothing but the rigidity of , which we confirmed in (1). Thus we have shown the rigidity of and .
Again in the spectral sequence (3.16), the vanishing implies . Hence is surjective. Also the vanishing implies . Since , the spectral sequence is -degenerate at and hence . This implies that is injective. Thus we obtain
(3.19) | |||||
(3.20) |
Therefore substituting (3.19) and (3.20) into
(3.21) |
and using (3.18), we obtain
Since is rigid, the same proof as in [OU15, Lemma 2.2] shows that is an exceptional vector bundle. ∎
Definition 3.7.
For an exceptional object , we will write for the unique integer such that .
Let be the rank function. If we let denote the embedding of the generic point, is concisely defined as the composition of the map and the isomorphism which sends the class of to . As a corollary of Lemma 3.3, we obtain the following
Corollary 3.8.
The equality
(3.22) |
holds for any exceptional object . In particular, .
3.2. Properties of the schematic support of
The aim of this subsection is to prove Proposition 3.18, which asserts that the schematic support of coincides with . This is the most technical part of this paper.
Remark 3.9.
To describe the schematic support of the cohomology sheaves of , we consider the anti-canonical morphism to the weighted projective plane of weight , which contracts the -curve to the singularity .
Notation 3.10.
Let denote the local ring of at the singular point. It is isomorphic to , the singularity.
Our first goal is to give an -module structure on for of the spectral sequence (3.4) with respect to which the differentials of the spectral sequence are -linear. The similar fact for spherical objects is used in [IU05], in which case the existence of such an -module structure is trivial. In fact, since the support of a spherical object is concentrated in , one can use the pushforward along .
Remark 3.11.
For any with , there is a natural homomorphism of -algebras which kills for some . Hence for and , there are natural -module structures of finite length on and given by precomposition and postcomposition, respectively. If is also supported in , we may use instead. However, we end up with the same -module structures defined via .
For , the reduced support of either or is contained in and therefore has a canonical -module structure by Remark 3.11. For and , .
For , we have
Since the reduced support of for is contained in , all the direct summands but the last one have canonical -module structures again by Remark 3.11.
Recall the following short exact sequence from the proof of Lemma 3.3 (3).
(3.23) |
Take to obtain the following exact sequence
where
(3.24) |
is the precomposition (think of as a morphism in ). The following vanishing follows from the exceptionality of .
Hence we conclude that the map is an isomorphism of -vector spaces. Since the reduced support of is contained in , the right hand side of (3.24) has a canonical -module structure.
Definition 3.12.
We transfer the -module structure on to via the isomorphism of (3.24), thereby giving an -module structure on .
Lemma 3.13.
For , the maps in the spectral sequence (3.4) are -linear.
Proof.
For , as already explained, the -module structures on the source and the target of are naturally defined and hence the -linearity of are rather obvious.
Let us show the -linearity of . Take an arbitrary element
(3.25) |
It suffices to show that the maps as follows, which appear in the description of given in Lemma 3.6, are -linear.
(3.26) | ||||
(3.27) |
Under the isomorphism of (3.24), the map (3.26) is identified with
and, similarly, the map (3.27) is identified with
These maps are -linear, for the reason that the -module structures given in Remark 3.11 are by means of precompositions. As the -module structure on is given in Definition 3.12 by transferring the -module structure on via , this is exactly what we had to prove. ∎
Suppose is a pure coherent sheaf on with . Then by Remark 3.11, has a standard -module structure of finite length and hence there is an integer such that . In general, for an -module and an ideal , we let denote the annihilator
(3.28) |
Geometrically speaking, this is the maximum submodule of which is “supported on the closed subscheme ”.
Lemma 3.14.
For a pure coherent sheaf on with , assume
where as above. Then the following strict inequality holds.
(3.29) |
Proof.
Without loss of generality, we may assume that is surjective; i.e., is -globally generated. In fact, if is not surjective, we can replace with the image . By standard arguments on the adjoint pair , there is a canonical isomorphism and is automatically -globally generated.
The following fact about the -singularity is known well: for each , the -th powers of the ideal sheaf
(3.30) |
of and are related to each other as follows.
(3.31) |
The vanishing means that is a sheaf on , from which we deduce that and hence its quotient are supported on . This means .
On the other hand, note that the image of the canonical morphism
(3.32) |
as a submodule of , is by (3.31). Thus we see , hence . Therefore we conclude
Since , we can naturally think of as an object of . Since we assumed that is pure, so is its subsheaf . Hence is a vector bundle on of positive rank.
Likewise can be thought of an object of , where is the sheaf version of the annihilator (3.28); i.e., the maximum subsheaf of whose schematic support is contained in . We claim that there is an isomorphism
(3.33) |
of vector bundles on . Indeed is induced from the product morphism , whose surjectivity implies that of . At the generic point of , the stalk is the maximal ideal of the discrete valuation ring and is a finite length -module. Hence one sees that is an isomorphism at (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 is not pure. Then there is a subsheaf such that is zero-dimensional. Then is non-zero and zero-dimensional, which contradicts the purity of . To see , note that there is an epimorphism
(3.34) |
and that .
Now consider the following -module.
To see the equality, note that the denominator of the left hand side, as a submodule of , coincides with the following subset.
Since , this coincides with the denominator of the right hand side.
Lemma 3.15.
Let and be a vector bundle and a line bundle on a smooth projective curve , respectively. Suppose and let be a non-zero linear subspace. Then we have a strict inequality
where denotes the product map.
Proof.
Without loss of generality, we may replace with the subsheaf generated by . Take a pair of linearly independent global sections . It follows that as subsheaves of . Since is locally free, this implies that as subsheaves of . Since and are the subsheaves of generated by and respectively, we conclude . Thus we see
(3.35) |
Taking , we obtain the assertion. ∎
Let be the category of Artinian () -modules. There is an (anti-)involution of categories
which is defined as follows.
(3.36) |
Corollary 3.16.
Under the same notation and assumption as in Lemma 3.14, put . Then the following strict inequality holds.
Proof.
Lemma 3.17.
For and integer , there are isomorphisms as follows.
(3.37) | ||||
(3.38) |
Proof.
To obtain the second isomorphism, replace with in the first isomorphism and then use . To see the first isomorphism, consider the following short exact sequence. The inclusion is the canonical one.
(3.39) |
By applying the anti-involution to this, we obtain the following short exact sequence.
(3.40) |
Then, as a submodule of , is computed as follows.
(3.41) |
Note that satisfies if and only if ; i.e., , where is the quotient map. Hence
(3.43) |
so that . ∎
Now we are ready to prove the following
Proposition 3.18.
For an exceptional object , it holds that
Proof.
We first discuss for . We may and will assume , since otherwise there is nothing to prove. Put
(3.44) |
In the spectral sequence (3.4) for the exceptional object , we have the isomorphism
(3.45) |
of -modules by Lemma 3.13. Note that , since for each either or holds.
On the other hand, for each there is an isomorphism of -modules given by the Serre duality:
This implies that . Combining this with the isomorphism of -modules (3.45), we obtain
Now assume for a contradiction that . Let
(3.46) |
so that . We can apply Lemma 3.14 to , to obtain the strict inequality
(3.47) |
On the other hand, let
(3.48) |
By Lemma 3.13 and the Serre duality, it follows that there is an isomorphism of -modules . We can apply Corollary 3.16 to , to obtain the strict inequality
(3.49) |
The strict inequalities (3.47) and (3.49) contradict the isomorphism (3.45). Hence we obtain , which means that is an -module for any . In fact, by rigidity, is a vector bundle on for any .
Finally, to investigate , we consider the derived dual . As we show in Lemma 3.25 below, there is an isomorphism as follows.
Since , by applying what we have just proved to the exceptional object , we see that the left hand side, hence the right hand side, is a vector bundle on . Hence so is . ∎
3.3. More on the structure of
In this subsection we give a structure theorem for in Lemma 3.21. It is then used to give a structure theorem for in Corollary 3.23.
Below is repeatedly used in this paper.
Lemma 3.20 ([Kul97, Remark 2.3.4]).
Let be an exceptional vector bundle of . Then there is an isomorphism
(3.50) |
for some and such that .
Note that the integers and in Lemma 3.20 are uniquely determined by .
Lemma 3.21.
Let be an exceptional object. Then the unique non-torsion cohomology sheaf decomposes as
where is an exceptional sheaf and is a vector bundle on . Moreover, if is not locally free, then there is an integer such that
-
•
The torsion part is a direct sum of copies of .
-
•
is a direct sum of copies of and .
Definition 3.22.
Proof of Lemma 3.21.
Consider the following standard short exact sequence.
(3.51) |
Lemma 3.3 (3) asserts that is an exceptional vector bundle. Hence we assume that , since otherwise there is nothing to prove.
is a rigid sheaf again by Lemma 3.3 (3). Combined with Proposition 3.18, this implies that there are such that
(3.52) |
By Lemma 3.20, is also rigid and hence there are such that
If , then the short exact sequence (3.51) splits and we are done. Therefore we assume , which implies
(3.53) |
In order to make the argument conceptual, fix a -vector space of dimension and replace with . Let
(3.54) |
be the quotient by the subsheaf , which fits in the following short exact sequence.
(3.55) |
From this we see
which implies that 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
(3.56) |
correspond to the extension (3.55). We will show that is injective. In the long exact sequence obtained by applying to (3.55),
-
•
the map is surjective since by the exceptionality of and
-
•
by the rigidity of .
Hence we obtain
Next we apply to (3.55) to obtain a surjective map
(3.57) |
For any , the map (3.57) sends to , where denotes the extension class (3.55). Take a basis of and decompose as under the isomorphism . Then one can confirm that (3.57) sends to and that sends a linear form to . Hence the surjectivity of the map (3.57) implies that generates , which in turn is equivalent to the injectivity of .
Now consider the universal extension of by .
(3.58) |
The inequality (3.53) implies , so that the distinguished triangle which (3.58) yields is isomorphic to the defining distinguished triangle (2.17) for the inverse spherical twist . In particular, there is an isomorphism and hence is an exceptional sheaf. Moreover, the injectivity of and the basic properties of universal extensions imply that there is an isomorphism
(3.59) |
In what follows let , so that
If , then is a vector bundle and we are done. So, in the rest of the proof, we assume ; i.e., we assume that .
Let us prove
(3.60) |
which together with (3.54) implies
(3.61) |
Since , this is the desired conclusion.
(3.60) follows from the local-to-global spectral sequence and the following vanishings.
(3.62) | ||||
(3.63) |
From Lemma 3.21 we immediately obtain
Corollary 3.23.
Suppose that is an exceptional object with . Then, with the notation of Lemma 3.21, it holds that
(3.65) |
for some .
Proof.
If , then and hence . Thus we see
Since is a direct summand of , this implies . Moreover, since , the local to global spectral sequence for Ext groups implies
(3.66) |
On the other hand, by Proposition 3.18 and the rigidity, there is a vector bundle on such that . Hence there are isomorphisms as follows.
(3.67) |
Combining these isomorphisms with (3.66), we obtain the following vanishings.
(3.68) |
From this we deduce that is of the form
for some and , concluding the proof. ∎
3.4. Derived dual of exceptional objects
Let be an exceptional object which is not isomorphic to a shift of a vector bundle, and let be the exceptional sheaf in Definition 3.22. In what follows we will mainly discuss the case where . If ( is a vector bundle), we will replace with its derived dual 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 on , the cohomology sheaves of the derived dual are related to those of as follows:
-
•
.
-
•
If , then .
-
•
For , fits into an exact sequence
-
•
For , the cohomology sheaf can also be written as
Proof.
Consider the following spectral sequence.
Since is a smooth projective surface, if or . Since is torsion for by Lemma 3.3 (2), for . Moreover, for any , is pure by Lemma 3.3 (2). Furthermore, is pure again by Lemma 3.3 (2) and is locally free by Lemma 3.3 (3). Hence by [HL10, Theorem 1.1.10, ], for any .
Summing up, we see that only if or . In particular, this spectral sequence is -degenerate. All assertions follow from these observations. ∎
Lemma 3.26.
Let be an exceptional object such that both and are vector bundles. Then .
Proof.
We consider the spectral sequence
As is a smooth surface, only if . In particular, it is -degenerate. Since for being torsion and being torsion free, it follows that and hence
(3.69) |
Take any
whose image under the surjection of the spectral sequence, which in fact is , corresponds to the natural inclusion under the isomorphisms (3.69). Then the derived dual
induces the surjection
in Lemma 3.25. As we assumed that is torsion free, by applying what we have just confirmed to , we similarly obtain a morphism
such that is the inclusion as a direct summand. Then it follows that the composite is an automorphism of and therefore splits as a direct sum of and . Since the exceptional object is indecomposable, this implies and hence , which is equivalent to . ∎
Thus we immediately obtain the following
Corollary 3.27.
Let be an exceptional object which is not isomorphic to a shift of a vector bundle. If is torsion free, then .
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 be the generic point of and the local ring of at . For a coherent sheaf , put
where is the stalk of at and is the length of its torsion part.
For an object , we define
Let us give a bit more concrete description for . Let be the pull-back of by the flat morphism . The length function is defined for objects in as well, and it immediately follows from the exactness of the (underived) pull-back functor that . On the other hand, since is a DVR, for there is an isomorphism and hence the equality . By the structure theorem for finitely generated modules over a DVR, is a direct sum of finite copies of and for various , where is a generator of the maximal ideal. Based on this, we obtain the following invariance.
Lemma 3.29.
For any object , .
Proof.
Note first that . Hence it is enough to show for all . By the explicit descriptions of we gave above, it is enough to show this for . In this case one can easily confirm , so we are done. ∎
If is an exceptional object with , then Proposition 3.18 implies
(3.70) |
where denotes the rank of a coherent sheaf on . Note that thus defined is the same as 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 is isomorphic to a shift of a vector bundle if and only if .
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
in , an inequality
holds. This is an equality if is a torsion sheaf.
Proof.
By taking the stalks at , it is enough to show the similar statements for . It follows from the standard arguments on local cohomology that
(3.71) |
is exact and that the last map is a surjection if is itself torsion (note that is isomorphic to the 0-th local cohomology functor at the closed point ). 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
from Lemma 3.21, where is an exceptional sheaf and is a torsion sheaf. There are two possibilities as follows.
-
(1)
is not torsion free. Then by Lemma 3.21, there are such that
(3.72) - (2)
Theorem 5.4 obviously follows from the following theorem, which asserts that it is always possible to decrease the length of by an appropriate spherical twist.
Theorem 3.32.
Proof.
In this proof, to make life easy, we assume . The general case is easily reduced to this just by replacing with .
Consider first the case (1). For each , by Corollary 3.23 there are such that . Putting and in (3.61), we can also write . Consider the following spectral sequence.
By direct computations we easily see that are as follows.
(3.73) |
Thus we see
(3.74) |
Noting that the differential map is non-zero only if , we easily see that are as follows.
For each , since is torsion, it follows from Lemma 3.31 that . Also, the inclusion implies . Thus we have confirmed the inequality
(3.75) |
Finally, the spectral sequence degenerates at and hence Lemma 3.31 implies
(3.76) |
4. Exceptional objects sharing the same class in
The goal of this section is to prove Corollary 4.4 on the set of exceptional objects sharing the same class in .
Let be an exceptional vector bundle of rank . Recall from (3.50) of Lemma 3.20 that there is an isomorphism
(4.1) |
for some and such that . We freely use this result, especially the symbols , and , throughout this section.
Lemma 4.1.
Let be an exceptional vector bundle. Then and are exceptional vector bundles.
Proof.
From the defining distinguished triangle of spherical twists (2.16), it immediately follows that is an exceptional sheaf. Note that it is isomorphic to . It then follows from the defining distinguished triangle for the inverse spherical twist (2.17) that is torsion free. Thus we see that and are both exceptional vector bundles. ∎
Lemma 4.2.
Let be an exceptional vector bundle such that holds in (3.50). Then and .
Proof.
Theorem 4.3.
Let be an exceptional object with . Then there exists and such that is an exceptional vector bundle.
Proof.
Corollary 4.4.
Let be an exceptional object. Then
- (1)
-
(2)
The action of the group on the following set is transitive.
(4.3)
5. Constructibility of exceptional collections
The aim of this section is Corollary 5.6, which asserts that any exceptional collection on is extendable to a full exceptional collection.
We first show that any exceptional collection on is sent to an exceptional collection consisting of (shifts of) vector bundles by a sequence of spherical twists.
Lemma 5.1.
Let be an exceptional pair such that is a vector bundle and is not isomorphic to a shift of a sheaf. Suppose also that defined in Definition 3.22 is not torsion free, so that is a direct sum of copies of for some . Then is isomorphic to either or ; in particular, it remains to be a vector bundle.
Proof.
Recall the decomposition from Lemma 3.21. Set , where by the assumption. By Corollary 3.23, there is an isomorphism
(5.1) |
for some . Also, again by Lemma 3.20, there is an isomorphism
for some and .
Consider the following spectral sequence.
(5.2) |
The assumption implies the vanishing of the limit for all . Also, since is a smooth projective surface, only if and hence (5.2) is -degenerate everywhere and -degenerate at . These imply that
-
•
for all , for being torsion and being torsion free. Hence (5.2) is -degenerate at for each , so that .
-
•
for all .
Thus we have confirmed that
-
•
only if or . Moreover,
-
•
is an isomorphism by the -degeneracy of (5.2) and the vanishing of the limit.
Note that and the Serre duality imply . Thus we see
(5.3) |
Take any . We know that is a direct sum of invertible sheaves on , so the vanishings imply if , and that is a direct sum of (possibly ) copies of if .
Let us first settle the case . As we mentioned in the previous paragraph, in this case only if or . Note in fact that , since it is assumed in the statement that is not isomorphic to a shift of a sheaf.
Lemma 5.2.
Let be an exceptional pair. Suppose that is a vector bundle and is a sheaf such that
Then is isomorphic to either or and it holds that .
Proof.
Consider the short exact sequence as follows.
(5.5) |
By [OU15, Theorem 1.4 (1)], we know that and is an exceptional vector bundle. Hence all we have to show is that is a vector bundle.
The vanishing implies . Especially one has
(5.6) | ||||
(5.7) |
If , then it follows that and hence . Therefore we may assume , or by taking the dual,
(5.8) |
The Serre dual of (5.6) and its restriction to yield the commutative square as follows.
(5.9) |
In the second row of the diagram (5.9), we omit by fixing an isomorphism and regard as a sheaf on . By restricting the locally split short exact sequence (5.5) to , we obtain the following short exact sequence.
(5.10) |
From (5.10) and the surjectivity of the second row in (5.9), we obtain the following isomorphism.
(5.11) |
As in the proof of Lemma 5.1, put
for and where . Then (5.7) implies
To determine the value of , 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 by [OU15, Theorem 1.4(1)]. We know moreover that by (5.10), where . Thus we obtain the following descriptions.
(5.12) |
Let us compute the dimension of the right hand side of (5.11). Since is an exceptional vector bundle, there are and such that
Note that by (5.8) and (5.9) we have
which implies .
Suppose for a contradiction that . Then we obtain and
(5.13) |
Using , 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
Then (5.12) implies either or . Thus we see that is isomorphic to or , respectively, by Lemma 4.1 and Lemma 4.2. Thus we conclude the proof. ∎
Corollary 5.3.
Let be an exceptional pair such that is a vector bundle and .
-
(1)
Suppose that is non-zero and is a direct sum of copies of . Then and is isomorphic to either or ; in particular, it is a vector bundle.
-
(2)
Suppose that , so that is non-zero by Lemma 3.26. Suppose that it is a direct sum of copies of . Then and is isomorphic to either or ; in particular, it is a vector bundle.
In particular, there is which depends only on , independent of in particular, such that and is isomorphic to either or .
Proof.
The assertions on the lengths are already proven in Theorem 3.32.
Let us first assume . The case where is not isomorphic to a shift of a sheaf is settled in Lemma 5.1. The case is isomorphic to a shift of a sheaf is settled in Lemma 5.2 (we may assume , without loss of generality).
If , we reduce the proof to the first case. In fact, it follows from Lemma 3.26 that . Note that
(5.14) |
is also an exceptional pair such that the first component is a vector bundle and . Suppose that is a direct sum of copies of . Then, by applying the conclusion of the previous paragraph to the pair (5.14), it follows that is isomorphic to either or . Then from the following computation we see that is isomorphic to either or .
(5.15) | |||
(5.16) |
∎
Theorem 5.4.
For any with and any exceptional collection , there exists a product of spherical twists of the form for some , denoted by such that .
Proof.
Without loss of generality, we may and will assume for all . We prove the assertion by an induction on .
The case when is nothing but Theorem 3.1. Consider the case when . By applying the induction hypothesis to the subcollection , we may and will assume that these are already vector bundles, say, . Suppose is not a vector bundle; i.e., . Otherwise there is nothing to show. In this case, since for each , if we take as in Corollary 5.3, then remains to be a vector bundle for all and it holds that . By repeating this process until reaches , 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 is not contained in the class, we can deduce the same assertion for from it:
Theorem 5.5.
Any exceptional collection on consisting of vector bundles can be extended to a full exceptional collection.
Proof.
Let
(5.17) |
be the blowup of in a point outside of the curve . Let be the exceptional curve.
Take , so that Write , and consider the right mutation . We claim that this is an exceptional vector bundle.
To see this, note first that the semiorthogonality implies , where . Thus we obtain the following short exact sequence.
(5.18) |
Suppose for a contradiction that the torsion part of the exceptional sheaf is nontrivial. Then it should map injectively to . It then implies that the canonical morphism from the locally free sheaf to the torsion free part of is both injective and surjective in codimension 1, which hence is an isomorphism. This contradicts the indecomposability of .
Thus we have obtained . By [Kul97, Theorem 3.1.8.2], it extends to a full exceptional collection on (note that is a weak del Pezzo surface obtained by blowing up first in a point and then in a point on the -curve, (hence) that is base point free and ). By applying the left mutation again, we obtain a full exceptional collection . Since , there are some objects such that . Since is an equivalence, we see that is a full exceptional collection of . ∎
We finally obtain the following constructibility theorem for .
Corollary 5.6.
Any exceptional collection on can be extended to a full exceptional collection.
Proof.
Remark 5.7.
Contrary to Theorem 4.3, an exceptional collection (consisting of objects of ) of length at least is not necessarily numerically equivalent to an exceptional collection of vector bundles. Namely, for each there is an exceptional collection of length and for for which there is no exceptional collection of vector bundles such that for . Example 5.8 below is such an example for . Examples for are obtained by extending examples of length by Corollary 5.6. For these exceptional collections, in particular, in Theorem 5.4 one can not take from . This is in contrast to Theorem 6.2.
Example 5.8.
Consider the following exceptional pair.
Then there is no exceptional pair of vector bundles such that for for the following reason.
Remark 5.9.
Given an exceptional collection such that for all , by Corollary 4.4 (1) there is a unique sequence of exceptional vector bundles such that for all . Example 5.8 implies that is not necessarily an exceptional collection.
It also implies that the map (the case appears in Figure 2.66) is not surjective for , though it is for by [OU15, Lemma 4.6] and Proposition 2.32. In fact, let be an exceptional collection (consisting of objects of ) which is not numerically equivalent to an exceptional collection of vector bundles. Then is not in the image of . In fact, an exceptional collection of vector bundles such that must satisfy for , which contradicts the choice of .
Despite Example 5.8, by Corollary 2.11, one can always bring an arbitrary exceptional collection (consisting of objects of ) to an exceptional collection of vector bundles by an element of up to a twist by .
Theorem 5.10.
Let be an exceptional collection with and for all . Then there exists such that .
Proof.
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 as in the following commutative diagram whose restriction to the subspaces of algebraic stability conditions is compatible with the generalization map
(5.20) |
where denote the set of isomorphism classes of full strong exceptional collections on (recall that a full strong exceptional collection yields a chamber of algebraic stability conditions in ). The fact that restricts to the sets of strong exceptional collections follows from Remark 2.19 and Corollary 2.17.
(5.21) |
Conjecturally, the Galois group ( the group of fiber-preserving automorphisms of ) of coincides with . As do not deform to , it is conceivable that coincides with the Galois group of . 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 , the action of the group on the following set is transitive.
(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 is transitive. Namely, Conjecture 1.2 holds true for .
The proof is divided into 3 steps. Let be the given exceptional collection of length .
Step 1.
Step 2.
Next, we show the following theorem.
Theorem 6.2.
For any satisfying
(6.2) |
there exists such that .
Proof of Theorem 6.2.
Write . We may and will assume that for all of the objects in the collection. By Corollary 4.4 (2), there exists such that . Hence by replacing with , we may and will assume that .
Next, by Corollary 5.3, there is such that and are both vector bundles. Recall that is like either (2.34) or (2.35). If is like (2.34), then it follows that both and are line bundles for some . Since and , it follows from Lemma 2.27 that and .
If is like (2.35) for , then both and are line bundles for some . Then it follows from the following computations and Lemma 2.27 that .
(6.3) | |||
(6.4) |
This immediately implies . Hence we may and will assume for .
At this point, in fact, we are done. To see this, note that both and are exceptional pairs of the triangulated subcategory , which is equivalent to , satisfying for . It is well known that any exceptional pair of is (up to shifts) of the form , hence is uniquely determined (up to shifts) by the class in the Grothendieck group. This immediately implies that for , hence the conclusion. ∎
Step 3.
In the previous step, we killed the possible difference between and by spherical twists; more precisely, we found such that
(6.5) |
(here we put 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 in (6.5) can be replaced by a sequence of mutations. We begin with a lemma.
Lemma 6.3.
The following isomorphisms hold.
(6.6) |
Proof.
By a direct computation, one can check that is the cone of the (essentially) unique non-trivial morphism
(6.7) |
The assertion immediately follows from this observation. ∎
Theorem 6.4.
For any , there exists such that .
Proof.
By Theorem 2.12, is a product of copies of 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 . We treat each case separately.
For , we have
(6.8) |
By Lemma 6.3, one can easily verify that
(6.9) |
For , we have
(6.10) |
By a direct computation, one can verify the following assertion.
(6.11) |
∎
Below is an important consequence of Theorem 6.1.
Corollary 6.5.
.
Proof.
Since is full and the fullness is preserved under the action of the group , 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 -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 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 . Int. Math. Res. Not. IMRN, (23):12781–12803, 2015.
- [Pos95] Leonid Positselski. All strictly exceptional collections in 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.