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

Categorical Torelli theorems for Gushel–Mukai threefolds

Augustinas Jacovskis, Xun Lin, Zhiyu Liu and Shizhuo Zhang
Abstract.

We show that a general ordinary Gushel–Mukai (GM) threefold XX can be reconstructed from its Kuznetsov component 𝒦u(X)\mathcal{K}u(X) together with an extra piece of data coming from tautological subbundle of the Grassmannian Gr(2,5)\mathrm{Gr}(2,5). We also prove that 𝒦u(X)\mathcal{K}u(X) determines the birational isomorphism class of XX, while 𝒦u(X)\mathcal{K}u(X^{\prime}) determines the isomorphism class of a special GM threefold XX^{\prime} if it is general. As an application, we prove a conjecture of Kuznetsov–Perry in dimension three under a mild assumption. Finally, we use 𝒦u(X)\mathcal{K}u(X) to restate a conjecture of Debarre–Iliev–Manivel regarding fibers of the period map for ordinary GM threefolds.

Key words and phrases:
Derived categories, Bridgeland moduli spaces, Kuznetsov components, Gushel–Mukai threefolds, Categorical Torelli theorem.
2020 Mathematics Subject Classification:
Primary 14F08; secondary 14J45, 14D20, 14D22

1. Introduction

In recent times, derived categories have played an important role in algebraic geometry; in many cases, much of the geometric information of a variety/scheme XX is encoded by its bounded derived category of coherent sheaves Db(X)\mathrm{D}^{b}(X). In this setting, one of the most fundamental questions that can be asked is whether Db(X)\mathrm{D}^{b}(X) recovers XX up to isomorphism, in other words, whether a derived Torelli theorem holds for XX. For varieties with ample or anti-ample canonical bundle (which include Fano varieties and varieties of general type), this question was answered affirmatively by Bondal–Orlov in [10].

1.1. Kuznetsov components and categorical Torelli theorems

Therefore, for the class of varieties above, it is natural to ask whether they are also determined up to isomorphism by less information than the whole derived category Db(X)\mathrm{D}^{b}(X). A natural candidate for this is a subcategory 𝒦u(X)\mathcal{K}u(X) of Db(X)\mathrm{D}^{b}(X) called the Kuznetsov component. This subcategory has been studied extensively by Kuznetsov and others (e.g. [32, 33, 29]) for many Fano varieties, including Gushel–Mukai (GM) varieties.

The question of whether 𝒦u(X)\mathcal{K}u(X) determines XX up to isomorphism has been studied for certain cases in the setting of Fano threefolds. In [7], the authors show that the Kuznetsov component completely determines cubic threefolds up to isomorphism, in other words, a categorical Torelli theorem holds for cubic threefolds YY. The same result was also verified in [52]. On the other hand, for many Fano varieties, the Kuznetsov component 𝒦u(X)\mathcal{K}u(X) does not determine the isomorphism class, but only the birational isomorphism class of XX. This is known as a birational categorical Torelli theorem. For instance, Kuznetsov components determine the birational isomorphism class of every index 11 prime Fano threefolds of even genus g8g\geq 8. For GM threefolds – the focus of our paper – by [30] it is known that there are birational GM threefolds with equivalent Kuznetsov components. So there are two natural questions to ask in this setting:

Question 1.1.
  1. (1)

    Does 𝒦u(X)\mathcal{K}u(X) determine the birational equivalence class of XX?

  2. (2)

    What extra data along with 𝒦u(X)\mathcal{K}u(X) do we need to identify a particular GM threefold XX from its birational equivalence class?

1.2. Main Results

1.2.1. (Refined) categorical Torelli for Gushel–Mukai threefolds

In the present paper, we deal with the case of index 11 prime Fano threefolds of degree 1010 and genus 66, also known as Gushel–Mukai threefolds (GM threefolds for short), which are split into two types: ordinary GM threefolds which arise as a quadric section of a linear section of the Grassmannian Gr(2,5)\mathrm{Gr}(2,5), and special GM threefolds which arise as double covers of a codimension three linear section of Gr(2,5)\mathrm{Gr}(2,5), branched over a degree ten K3 surface. By [29], we have a semiorthogonal decomposition

Db(X)=𝒦u(X),,𝒪X,\mathrm{D}^{b}(X)=\langle\mathcal{K}u(X),\mathcal{E},\operatorname{\mathcal{O}}_{X}\rangle,

where \mathcal{E} is the pull-back of the tautological subbundle on Gr(2,5)\operatorname{Gr}(2,5) along the natural map XGr(2,5)X\to\operatorname{Gr}(2,5).

Our first main theorem is concerned with ordinary GM threefolds and answers Question 1.1 (2):

Theorem 1.2 (Theorem 9.2).

Let XX be a general ordinary GM threefold and π:Db(X)𝒦u(X)\pi\colon\mathrm{D}^{b}(X)\rightarrow\mathcal{K}u(X) be the right adjoint to the inclusion 𝒦u(X)Db(X)\mathcal{K}u(X)\subset\mathrm{D}^{b}(X). Then the data of 𝒦u(X)\mathcal{K}u(X) along with the object π()\pi(\mathcal{E}) is enough to determine XX up to isomorphism.

On the other hand, for special GM threefolds which are general (“general special” for short), we show that a categorical Torelli theorem holds:

Theorem 1.3 (Theorem 9.9).

Let XX and XX^{\prime} be general special GM threefolds, and assume that there is an equivalence of categories 𝒦u(X)𝒦u(X)\mathcal{K}u(X)\simeq\mathcal{K}u(X^{\prime}). Then XX and XX^{\prime} are isomorphic.

1.2.2. Birational categorical Torelli for Gushel–Mukai threefolds

Next, returning to the setting of ordinary GM threefolds, we show that a birational categorical Torelli theorem holds for general ordinary GM threefolds, which answers Question 1.1 (1).

Theorem 1.4 (Theorem 9.3).

Let XX and XX^{\prime} be general ordinary GM threefolds, and suppose that there is an equivalence of categories 𝒦u(X)𝒦u(X)\mathcal{K}u(X)\simeq\mathcal{K}u(X^{\prime}). Then XX is birationally equivalent to XX^{\prime}.

In [30], the authors studied GM varieties of arbitrary dimension and proved the Duality Conjecture [29, Conjecture 3.7] for them, i.e. they showed that the period partner or period dual of a GM variety XX shares the same Kuznetsov component 𝒦u(X)\mathcal{K}u(X) as XX. Combining earlier results [14, Theorem 4.20] on the birational equivalence of these varieties, this gives strong evidence for the following conjecture:

Conjecture 1.5 ([30, Conjecture 1.7]).

If XX and XX^{\prime} are GM varieties of the same dimension such that there is an equivalence 𝒦u(X)𝒦u(X)\mathcal{K}u(X)\simeq\mathcal{K}u(X^{\prime}), then XX and XX^{\prime} are birationally equivalent.

Thus our result Theorem 1.4 actually proves Conjecture 1.5 under the assumption that XX and XX^{\prime} are both of dimension 33, ordinary and general.

Moreover, by a careful study of Bridgeland moduli spaces of stable objects in the Kuznetsov components 𝒜X\mathcal{A}_{X} for not only smooth ordinary GM threefolds but also special GM threefolds XX, we can prove that the Kuznetsov component of a general ordinary GM threefold can not be equivalent to the one of a general special GM threefold. Therefore, combined with Theorem 1.4 and 1.3, we have the following improved version of Theorem 1.4, which allows threefolds to be either ordinary or special:

Theorem 1.6 (Theorem 9.7 and Corollary 9.8).

If XX and XX^{\prime} are general ordinary or general special GM threefolds such that there is an equivalence 𝒦u(X)𝒦u(X)\mathcal{K}u(X)\simeq\mathcal{K}u(X^{\prime}), then XX and XX^{\prime} are birationally equivalent.

1.2.3. The Debarre–Iliev–Manivel Conjecture

In [13], the authors conjecture that the general fiber of the classical period map from the moduli space of ordinary GM threefolds to the moduli space of 1010 dimensional principally polarised abelian varieties is birational to the disjoint union of the minimal model 𝒞m(X)\mathcal{C}_{m}(X) of the Fano surface of conics and a moduli space of stable sheaves MGX(2,1,5)M_{G}^{X}(2,1,5), both quotiented by involutions, which we call the Debarre–Iliev–Manivel Conjecture (cf. Conjecture 10.1). Within the moduli space of smooth ordinary GM threefolds, we define the fiber of the “categorical period map” through [X][X] as the isomorphism classes of all ordinary GM threefolds XX^{\prime} whose Kuznetsov components satisfy 𝒦u(X)𝒦u(X)\mathcal{K}u(X^{\prime})\simeq\mathcal{K}u(X). Then the following categorical analogue of the Debarre–Iliev–Manivel conjecture follows from Theorem 1.4 and results on Bridgeland moduli spaces with respect to the two (1)(-1)-classes in the numerical Grothendieck group of 𝒜X\mathcal{A}_{X}.

Theorem 1.7 (Theorem 10.3).

A general fiber of the “categorical period map” through an ordinary GM threefold XX is the union of 𝒞m(X)/ι\mathcal{C}_{m}(X)/\iota and MGX(2,1,5)/ιM_{G}^{X}(2,1,5)/\iota^{\prime} where ι,ι\iota,\iota^{\prime} are geometrically meaningful involutions.

As an application, the Debarre–Iliev–Manivel Conjecture 10.1 can be restated in an equivalent form as follows:

Conjecture 1.8.

Let XX be a general ordinary GM threefold. The intermediate Jacobian J(X)J(X) determines the Kuznetsov component 𝒦u(X)\mathcal{K}u(X).

Remark 1.9.

In [13], the authors actually conjecture that a general fiber of the period map is birational to the disjoint union of two surfaces, parametrizing conic transforms and conic transforms of a line transform of XX, which is birational to the disjoint union of 𝒞m(X)\mathcal{C}_{m}(X) and MGX(2,1,5)M_{G}^{X}(2,1,5), both quotiented by involutions. In Corollary 9.5 we show that this birational equivalence is indeed an isomorphism.

1.2.4. Uniqueness of Serre-invariant stability conditions

One of the key steps when we identify Bridgeland moduli spaces via an equivalence of Kuznetsov components in the proofs of Theorems 1.2 and 1.4. A stability condition σ\sigma on the Kuznetsov component 𝒦u(X)\mathcal{K}u(X) of a prime Fano threefold XX is Serre-invariant if S𝒦u(X)σ=σgS_{\mathcal{K}u(X)}\cdot\sigma=\sigma\cdot g for some gGL~+(2,)g\in\widetilde{\mathrm{GL}}^{+}(2,\mathbb{R}) (see Section 4.4). Serre-invariance is one of the fundamental tools in studying relationship of classical Gieseker moduli spaces and Bridgeland moduli spaces for Kuznetsov components (cf. [1, 52, 54, 41, 16]). A natural question is whether any two Serre-invariant stability conditions are in the same GL~+(2,)\widetilde{\mathrm{GL}}^{+}(2,\mathbb{R})-orbit. In the present paper, we answer this question affirmatively.

Theorem 1.10 (Theorem A.10).

Let XX be a prime Fano threefold of index 11 of genus g6g\geq 6, or a del Pezzo threefold of degree d2d\geq 2. Then all Serre-invariant stability conditions on 𝒦u(X)\mathcal{K}u(X) are in the same GL~+(2,)\widetilde{\mathrm{GL}}^{+}(2,\mathbb{R})-orbit.

1.3. Methods

For convenience, we work with the alternative Kuznetsov component 𝒜X\mathcal{A}_{X}, defined by the semiorthogonal decomposition Db(X)=𝒜X,𝒪X,\mathrm{D}^{b}(X)=\langle\mathcal{A}_{X},\operatorname{\mathcal{O}}_{X},\mathcal{E}^{\vee}\rangle and there is an equivalence Ξ:𝒦u(X)𝒜X\Xi\colon\mathcal{K}u(X)\simeq\mathcal{A}_{X}. We prove the above theorems 1.2, 1.4, 1.6 and 1.7 by considering the moduli spaces of Bridgeland stable objects in the alternative Kuznetsov component 𝒜X\mathcal{A}_{X} with respect to (1)(-1)-classes in the numerical Grothendieck group of 𝒜X\mathcal{A}_{X}, i.e. a vector vv with χ(v,v)=1\chi(v,v)=-1 where χ\chi is the Euler form. Up to sign, there are two (1)(-1)-classes in the numerical Grothendieck group of 𝒜X\mathcal{A}_{X}, call them x-x and y2xy-2x.

First, we show that the moduli space with the class x-x is isomorphic to the minimal model 𝒞m(X)\mathcal{C}_{m}(X) of the Fano surface of conics (Theorem 7.12). Indeed, we first show that the unique exceptional curve contracted in 𝒞(X)\mathcal{C}(X) is the rational curve of conics whose ideal sheaf ICI_{C} is not in 𝒜X\mathcal{A}_{X} and that the image is the smooth point represented by π()\pi(\mathcal{E}) (Proposition 7.1), so 𝒞m(X)\mathcal{C}_{m}(X) forms an irreducible component of the moduli space σ(𝒜X,x)\mathcal{M}_{\sigma}(\mathcal{A}_{X},-x) of stable objects in 𝒜X\mathcal{A}_{X} with respect to x-x.

Then we show this component actually occupies the whole moduli space (Proposition 7.11), which is the most difficult and technical part of the article and we only briefly sketch the argument here. We start with a stable object F𝒜XF\in\mathcal{A}_{X} of the class x-x. It suffices to show that FF is isomorphic to the projection of ideal sheaf ICI_{C} of a conic CXC\subset X. First, we assume that FF is semistable in the double tilted heart Cohα,β0(X)\operatorname{\mathrm{Coh}}^{0}_{\alpha,\beta}(X) (cf. Section 4.4). Then by a wall-crossing argument, we prove that F[1]F[-1] is a slope-semistable sheaf of rank one. Since its class is [F]=[IC][F]=-[I_{C}], we get FIC[1]F\cong I_{C}[1]. Next, we assume that FF is not semistable in the double tilted heart Cohα,β0(X)\operatorname{\mathrm{Coh}}^{0}_{\alpha,\beta}(X). Our main tools are inequalities in [36] and Theorem 4.7, which allow us to bound the rank and first two Chern characters ch1,ch2\mathrm{ch}_{1},\mathrm{ch}_{2} of the destabilizing objects and their cohomology objects. Since F𝒜XF\in\mathcal{A}_{X}, by using the Euler characteristics χ(𝒪X,)\chi(\mathcal{O}_{X},-) and χ(,)\chi(\mathcal{E}^{\vee},-) we can obtain a bound on ch3\mathrm{ch}_{3}. Then we deduce that the Harder–Narasimhan factors of FF are the expected ones (Proposition 7.10). As a result, σ(𝒜X,x)𝒞m(X)\mathcal{M}_{\sigma}(\mathcal{A}_{X},-x)\cong\mathcal{C}_{m}(X). Similarly, we identify the moduli space MGX(2,1,5)M_{G}^{X}(2,1,5) of Gieseker semistable sheaves of rank 22, c1=1,c2=5c_{1}=1,c_{2}=5 and c3=0c_{3}=0 on XX with the Bridgeland moduli space σ(𝒜X,y2x)\mathcal{M}_{\sigma}(\mathcal{A}_{X},y-2x) in Theorem 8.9.

As we have seen, 𝒞(X)\mathcal{C}(X) is exactly the blow-up of 𝒞m(X)σ(𝒜X,x)\mathcal{C}_{m}(X)\cong\mathcal{M}_{\sigma}(\mathcal{A}_{X},-x) at the point Ξ(π())\Xi(\pi(\mathcal{E})), hence the data (𝒦u(X),π())(\mathcal{K}u(X),\pi(\mathcal{E})) determines 𝒞(X)\mathcal{C}(X). A classical result of Logachev [39] states that XX can be determined up to isomorphism from 𝒞(X)\mathcal{C}(X). Thus Theorem 1.2 is proved.

We prove Theorem 1.3 via another method. By considering the equivariant Kuznetsov components 𝒦u(X)μ2\mathcal{K}u(X)^{\mu_{2}}, first discussed in [28], and exploiting the fact that XX is the double cover of a degree 55 index 22 prime Fano threefold YY, branched over a quadric hypersurface Y\mathcal{B}\subset Y. In this case, the equivariant Kuznetsov component is equivalent to Db()\mathrm{D}^{b}(\mathcal{B}) where \mathcal{B} is a K3 surface. Therefore, a number of results concerning the Fourier–Mukai partners of K3 surfaces can be used to deduce that 𝒦u(X)μ2𝒦u(X)μ2\mathcal{K}u(X)^{\mu_{2}}\simeq\mathcal{K}u(X^{\prime})^{\mu_{2}} implies \mathcal{B}\cong\mathcal{B}^{\prime}. Then the fact that the del Pezzo threefold YY of degree 55 is rigid can be used to deduce that indeed, XXX\cong X^{\prime}.

To prove Theorem 1.4, we invoke a few more results from [13]. More precisely, an equivalence of categories Φ:𝒜X𝒜X\Phi\colon\mathcal{A}_{X}\simeq\mathcal{A}_{X^{\prime}} identifies the moduli space σ(𝒜X,x)\mathcal{M}_{\sigma}(\mathcal{A}_{X},-x) with either σ(𝒜X,x)\mathcal{M}_{\sigma}(\mathcal{A}_{X},-x) or σ(𝒜X,y2x)\mathcal{M}_{\sigma}(\mathcal{A}_{X^{\prime}},y-2x). The former case gives an isomorphism of minimal surfaces 𝒞m(X)𝒞m(X)\mathcal{C}_{m}(X)\cong\mathcal{C}_{m}(X^{\prime}). Blowing 𝒞m(X)\mathcal{C}_{m}(X) up at the smooth point associated to π()\pi(\mathcal{E}) gives 𝒞(X)\mathcal{C}(X), and blowing up 𝒞m(X)\mathcal{C}_{m}(X^{\prime}) at the image of π()\pi(\mathcal{E}) under Φ\Phi gives 𝒞(XC)\mathcal{C}(X^{\prime}_{C}), where XCX^{\prime}_{C} is certain birational transformation of XX^{\prime}, associated with a conic CXC\subset X^{\prime}. Then by Logachev’s Reconstruction Theorem for 𝒞(X)\mathcal{C}(X), XX is isomorphic to XCX^{\prime}_{C} which is birational to XX^{\prime}. For the latter case, we start with the isomorphism 𝒞m(X)MGX(2,1,5)\mathcal{C}_{m}(X)\cong M_{G}^{X^{\prime}}(2,1,5). In fact, MGX(2,1,5)M_{G}^{X^{\prime}}(2,1,5) is birational to 𝒞(XL)\mathcal{C}(X^{\prime}_{L}), where XLX^{\prime}_{L} is another birational transformation of XX^{\prime}, associated with a line LXL\subset X^{\prime}. Since 𝒞(XL)\mathcal{C}(X^{\prime}_{L}) is a surface of general type, we get 𝒞m(X)𝒞m(XL)\mathcal{C}_{m}(X)\cong\mathcal{C}_{m}(X_{L}^{\prime}). Then by the same argument as in the previous case, XX is isomorphic to some birational transformation of XX^{\prime}.

Finally, the proof of Theorem 1.6 is similar to that of Theorem 1.4. Firstly, we identify the Bridgeland moduli spaces σ(𝒜X,x)\mathcal{M}_{\sigma}(\mathcal{A}_{X^{\prime}},-x) and σ(𝒜X,y2x)\mathcal{M}_{\sigma}(\mathcal{A}_{X^{\prime}},y-2x) on a special GM threefold XX^{\prime} with 𝒞m(X)\mathcal{C}_{m}(X^{\prime}) and MGX(2,1,5)M^{X^{\prime}}_{G}(2,1,5) respectively (Theorem 7.12 and Theorem 8.9), where 𝒞m(X)\mathcal{C}_{m}(X^{\prime}) is the contraction of the Fano surface 𝒞(X)\mathcal{C}(X^{\prime}) of conics on XX^{\prime} along one of the components to a singular point. Then if XX is ordinary, the equivalence Φ:𝒜X𝒜X\Phi\colon\mathcal{A}_{X}\simeq\mathcal{A}_{X^{\prime}} would identify those moduli spaces on a general ordinary GM threefold XX with those on a special GM threefold XX^{\prime}; we show that this is impossible by analyzing their singularities. Then Theorem 1.6 reduces to Theorem 1.4 and Theorem 1.3.

1.4. Related Work

1.4.1. Categorical Torelli theorems

There is a very nice survey article [51] on recent results and remaining open questions on this topic. In [7] and [52], the authors prove categorical Torelli theorems for cubic threefolds. In [1] and [12], the authors prove categorical Torelli theorems for general quartic double solids. In [38] and [40], the authors prove a refined categorical Torelli theorem for Enriques surfaces. In [26], the authors generalize Theorem 9.2 to all prime Fano threefolds of genus g6g\geq 6. In [18], the authors prove a birational categorical Torelli theorem for general non-Hodge-special Gushel–Mukai fourfolds.

1.4.2. Identifying classical moduli spaces as Bridgeland moduli spaces for Kuznetsov components

In the present article, we realize the Fano surface of conics and a certain Gieseker moduli space of semistable sheaves as Bridgeland moduli spaces of stable objects in Kuznetsov components of GM threefolds. In [52], the authors realize the Fano surface of lines Σ(Yd)\Sigma(Y_{d}) (for d2d\geq 2) as a Bridgeland moduli space of stable objects in the Kuznetsov component 𝒦u(Yd)\mathcal{K}u(Y_{d}). In [41], the authors realize the moduli space of rank two instanton sheaves on a del Pezzo threefold YdY_{d} (for d3d\geq 3) and the compactification of the moduli space of ACM sheaves on X4d+2X_{4d+2} (for d3d\geq 3) as Bridgeland moduli spaces of stable objects in 𝒦u(Yd)\mathcal{K}u(Y_{d}) and 𝒦u(X4d+2)\mathcal{K}u(X_{4d+2}), respectively. In [16], the authors realize the moduli space of Ulrich bundles of arbitrary rank on a cubic threefold Y3Y_{3} as an open locus of a Bridgeland moduli space of stable objects in 𝒦u(Y3)\mathcal{K}u(Y_{3}).

1.4.3. Serre-invariant stability conditions

In [48] and [52], the authors prove that stability conditions on Kuznetsov components of every del Pezzo threefold YdY_{d} of degree d1d\geq 1 and every index 11 prime Fano threefold of genus g6g\geq 6 are Serre-invariant. In [16], the authors prove the uniqueness of Serre-invariant stability conditions for a general triangulated category satisfying a list of very natural assumptions, which include Kuznetsov components of a series of prime Fano threefolds.

1.5. Notation and conventions

  • We work over the field k=k=\mathbb{C}. All triangulated categories and abelian categories are assumed to be kk-linear.

  • We use hom\hom and exti\operatorname{ext}^{i} to represent the dimension of the vector spaces Hom\operatorname{Hom} and Exti\operatorname{Ext}^{i}.

  • The numerical KK group of a triangulated category 𝒟\mathcal{D} is denoted by 𝒩(𝒟)\mathcal{N}(\mathcal{D}), which is the Grothendieck group K0(𝒟)K_{0}(\mathcal{D}) modulo the kernel of the Euler form χ(E,F)=i(1)iexti(E,F)\chi(E,F)=\sum_{i}(-1)^{i}\operatorname{ext}^{i}(E,F)

  • We denote the bounded derived category of a smooth projective variety XX by Db(X)\mathrm{D}^{b}(X). The derived dual functor is denoted by 𝔻:=RomX(,𝒪X)\mathbb{D}:=\mathrm{R}\mathcal{H}om_{X}(-,\operatorname{\mathcal{O}}_{X}).

  • We denote the phase and slope with respect to a weak stability condition σ\sigma by ϕσ\phi_{\sigma} and μσ\mu_{\sigma}, respectively. The maximal and minimal slopes (phases) of the Harder–Narasimhan factors of a given object FF will be denoted by μσ+(F)\mu_{\sigma}^{+}(F) (ϕσ+(F)\phi^{+}_{\sigma}(F)) and μσ(F)\mu_{\sigma}^{-}(F) (ϕσ(F)\phi^{-}_{\sigma}(F)), respectively.

  • 𝒜i\mathcal{H}^{i}_{\mathcal{A}} means the ii-th cohomology with respect to the heart 𝒜\mathcal{A}. When the 𝒜\mathcal{A} subscript is dropped, we take the heart to be Coh(X)\operatorname{\mathrm{Coh}}(X).

  • The symbol \simeq denotes an equivalence of categories and a birational equivalence of varieties. The symbol \cong denotes an isomorphism of varieties.

  • Let XX be a GM threefold. Then a conic means a closed subscheme CXC\subset X with Hilbert polynomial pC(t)=1+2tp_{C}(t)=1+2t, and a line means a closed subscheme LXL\subset X with Hilbert polynomial pL(t)=1+tp_{L}(t)=1+t.

1.6. Organization of the paper

In Section 2, we collect basic facts about semiorthogonal decompositions. In Section 3, we introduce Gushel–Mukai threefolds and their Kuznetsov components. In Section 4, we introduce the definition of weak stability conditions on Db(X)\mathrm{D}^{b}(X), and the induced stability conditions on the alternative Kuznetsov components 𝒜X\mathcal{A}_{X} of GM threefolds. In Section 5, we introduce a distinguished object π()𝒦u(X)\pi(\mathcal{E})\in\mathcal{K}u(X) and its alternative Kuznetsov component analogue Ξ(π())𝒜X\Xi(\pi(\mathcal{E}))\in\mathcal{A}_{X} and prove its stability. In Section 6 we discuss the geometry of the Fano surface of conics of a GM threefold. In Section 7, we construct the Bridgeland moduli space of σ\sigma-stable objects with class x-x in 𝒜X\mathcal{A}_{X}. In Section 8, we construct the Bridgeland moduli space of σ\sigma-stable objects with respect to the other (1)(-1)-class y2xy-2x in 𝒜X\mathcal{A}_{X}. In Section 9, we prove several birational/refined categorical Torelli theorems (Theorems 1.2, 1.3 and 1.4) and Conjecture 1.5 in dimension three with mild assumptions. In Section 10, we describe the general fiber of the “categorical period map” for ordinary GM threefolds 1.7, and restate the Debarre–Iliev–Manivel conjecture in terms of Conjecture 10.6. Finally, we study Serre-invariant stability conditions on Kuznetsov components and show that they are contained in one GL~+(2,)\widetilde{\mathrm{GL}}^{+}(2,\mathbb{R}) orbit in Appendix A.

Acknowledgements

Firstly, it is our pleasure to thank Arend Bayer for very useful discussions on the topics of this project. We would like to thank Sasha Kuznetsov for answering many of our questions on Gushel–Mukai threefolds. We thank Atanas Iliev, Laurent Manivel, Daniele Faenzi, Dmitry Logachev, Will Donovan, Bernhard Keller, Alexey Elagin, Xiaolei Zhao, Chunyi Li, Laura Pertusi, Song Yang, Alex Perry, Pieter Belmans, Qingyuan Jiang, Enrico Fatighenti, Naoki Koseki, Bingyu Xia, Yong Hu and Luigi Martinelli for helpful conversations on several related topics. We would like to thank Daniele Faenzi for sending us the preprint [17] and Soheyla Feyzbakhsh for sending us the preprint [16]. We thank Pieter Belmans for useful comments on the first draft of our article. The last author would like to thank Tingyu Sun for constant support and encouragement. We would like to thank the anonymous referee for their careful reading of our paper, and for their very useful and insightful comments.

The first and last authors are supported by ERC Consolidator Grant WallCrossAG, no. 819864. The first author was also supported by the Luxembourg National Research Fund (FNR–17113194).

2. Semiorthogonal decompositions

In this section, we collect some useful facts about semiorthogonal decompositions. Background on triangulated categories and derived categories of coherent sheaves can be found in [21], for example. From now on, let Db(X)\mathrm{D}^{b}(X) denote the bounded derived category of coherent sheaves on a smooth projective variety XX, and for E,FDb(X)E,F\in\mathrm{D}^{b}(X), define

RHom(E,F)=iHom(E,F[i])[i].\operatorname{RHom}^{\bullet}(E,F)=\bigoplus_{i\in\mathbb{Z}}\operatorname{Hom}(E,F[i])[-i].

2.1. Exceptional collections and semiorthogonal decompositions

Definition 2.1.

Let 𝒟\mathcal{D} be a triangulated category and E𝒟E\in\mathcal{D}. We say that EE is an exceptional object if RHom(E,E)=k\operatorname{RHom}^{\bullet}(E,E)=k. Now let (E1,,Em)(E_{1},\dots,E_{m}) be a collection of exceptional objects in 𝒟\mathcal{D}. We say it is an exceptional collection if RHom(Ei,Ej)=0\operatorname{RHom}^{\bullet}(E_{i},E_{j})=0 for i>ji>j.

Definition 2.2.

Let 𝒟\mathcal{D} be a triangulated category and 𝒞\mathcal{C} be a triangulated subcategory of 𝒟\mathcal{D}. We define the right orthogonal complement of 𝒞\mathcal{C} in 𝒟\mathcal{D} as the full triangulated subcategory

𝒞={X𝒟Hom(Y,X)=0 for all Y𝒞}.\mathcal{C}^{\bot}=\{X\in\mathcal{D}\mid\operatorname{Hom}(Y,X)=0\text{ for all }Y\in\mathcal{C}\}.

The left orthogonal complement is defined similarly, as

𝒞={X𝒟Hom(X,Y)=0 for all Y𝒞}.{}^{\bot}\mathcal{C}=\{X\in\mathcal{D}\mid\operatorname{Hom}(X,Y)=0\text{ for all }Y\in\mathcal{C}\}.
Definition 2.3.

Let 𝒟\mathcal{D} be a triangulated category. We say a triangulated subcategory 𝒞𝒟\mathcal{C}\subset\mathcal{D} is admissible if the inclusion functor i:𝒞𝒟i\colon\mathcal{C}\hookrightarrow\mathcal{D} has left adjoint ii^{*} and right adjoint i!i^{!}.

Definition 2.4.

Let 𝒟\mathcal{D} be a triangulated category, and (𝒞1,,𝒞m)(\mathcal{C}_{1},\dots,\mathcal{C}_{m}) be a collection of full admissible subcategories of 𝒟\mathcal{D}. We say that 𝒟=𝒞1,,𝒞m\mathcal{D}=\langle\mathcal{C}_{1},\dots,\mathcal{C}_{m}\rangle is a semiorthogonal decomposition of 𝒟\mathcal{D} if 𝒞j𝒞i\mathcal{C}_{j}\subset\mathcal{C}_{i}^{\bot} for all i>ji>j, and the subcategories (𝒞1,,𝒞m)(\mathcal{C}_{1},\dots,\mathcal{C}_{m}) generate 𝒟\mathcal{D}, i.e. the category resulting from taking all shifts and cones of objects in the categories (𝒞1,,𝒞m)(\mathcal{C}_{1},\dots,\mathcal{C}_{m}) is equivalent to 𝒟\mathcal{D}.

Let S𝒟S_{\mathcal{D}} be the Serre functor of 𝒟\mathcal{D}, then we have the following standard result, see e.g. [5, Section 3]:

Proposition 2.5 ([5, Section 3]).

If 𝒟=𝒟1,𝒟2\mathcal{D}=\langle\mathcal{D}_{1},\mathcal{D}_{2}\rangle is a semiorthogonal decomposition, then 𝒟=S𝒟(𝒟2),𝒟1=𝒟2,S𝒟1(𝒟1)\mathcal{D}=\langle S_{\mathcal{D}}(\mathcal{D}_{2}),\mathcal{D}_{1}\rangle=\langle\mathcal{D}_{2},S^{-1}_{\mathcal{D}}(\mathcal{D}_{1})\rangle are also semiorthogonal decompositions.

2.2. Mutations

Let 𝒞𝒟\mathcal{C}\subset\mathcal{D} be an admissible triangulated subcategory. Then the left mutation functor 𝐋𝒞\bm{\mathrm{L}}_{\mathcal{C}} through 𝒞\mathcal{C} is defined as the functor lying in the canonical functorial exact triangle

ii!id𝐋𝒞ii^{!}\rightarrow\operatorname{id}\rightarrow\bm{\mathrm{L}}_{\mathcal{C}}

and the right mutation functor 𝐑𝒞\bm{\mathrm{R}}_{\mathcal{C}} through 𝒞\mathcal{C} is defined similarly, by the triangle

𝐑𝒞idii.\bm{\mathrm{R}}_{\mathcal{C}}\rightarrow\operatorname{id}\rightarrow ii^{*}.

When EDb(X)E\in\mathrm{D}^{b}(X) is an exceptional object, and FDb(X)F\in\mathrm{D}^{b}(X) is any object, the left mutation 𝐋EF\bm{\mathrm{L}}_{E}F fits into the triangle

ERHom(E,F)F𝐋EF,E\otimes\operatorname{RHom}^{\bullet}(E,F)\rightarrow F\rightarrow\bm{\mathrm{L}}_{E}F,

and the right mutation 𝐑EF\bm{\mathrm{R}}_{E}F fits into the triangle

𝐑EFFERHom(F,E).\bm{\mathrm{R}}_{E}F\rightarrow F\rightarrow E\otimes\operatorname{RHom}^{\bullet}(F,E)^{\vee}.
Proposition 2.6 ([35, Lemma 2.6]).

Let 𝒟=𝒜,\mathcal{D}=\langle\mathcal{A},\mathcal{B}\rangle be a semiorthogonal decomposition. Then

S=𝐑𝒜S𝒟 and S𝒜1=𝐋S𝒟1.S_{\mathcal{B}}=\bm{\mathrm{R}}_{\mathcal{A}}\circ S_{\mathcal{D}}\,\,\,\text{ and }\,\,\,S_{\mathcal{A}}^{-1}=\bm{\mathrm{L}}_{\mathcal{B}}\circ S_{\mathcal{D}}^{-1}.
Lemma 2.7 ([34, Lemma 2.7]).

Let 𝒟=𝒞1,𝒞2,,𝒞n\mathcal{D}=\langle\mathcal{C}_{1},\mathcal{C}_{2},...,\mathcal{C}_{n}\rangle be a semiorthogonal decomposition with all components being admissible. Then for each 1kn11\leq k\leq n-1, there is a semiorthogonal decomposition

𝒟=𝒞1,,𝒞k1,𝐋𝒞k𝒞k+1,𝒞k,𝒞k+2,𝒞n\mathcal{D}=\langle\mathcal{C}_{1},...,\mathcal{C}_{k-1},\bm{\mathrm{L}}_{\mathcal{C}_{k}}\mathcal{C}_{k+1},\mathcal{C}_{k},\mathcal{C}_{k+2}...,\mathcal{C}_{n}\rangle

and for each 2kn2\leq k\leq n there is a semiorthogonal decomposition

𝒟=𝒞1,,𝒞k2,𝒞k,𝐋𝒞k𝒞k1,𝒞k+1,𝒞n.\mathcal{D}=\langle\mathcal{C}_{1},...,\mathcal{C}_{k-2},\mathcal{C}_{k},\bm{\mathrm{L}}_{\mathcal{C}_{k}}\mathcal{C}_{k-1},\mathcal{C}_{k+1}...,\mathcal{C}_{n}\rangle.

3. Gushel–Mukai threefolds and their derived categories

Let XX be a prime Fano threefold of index 11 and degree H3=10H^{3}=10, where HH is the ample generator of CaCl(X)\operatorname{CaCl}(X). Then XX is either a quadric section of a linear section of codimension 22 of the Grassmannian Gr(2,5)\mathrm{Gr}(2,5), in which case it is called an ordinary Gushel–Mukai (GM) threefold, or XX is a double cover of a degree 55 and index 22 Fano threefold YY ramified in a quadric hypersurface, in which case it is called a special GM threefold. In the latter case, it has a natural involution τ:XX\tau\colon X\rightarrow X induced by the double cover π:XY\pi\colon X\rightarrow Y. By [43, 6], there exists a stable vector bundle \mathcal{E} of rank 22 with c1()=H,c2()=4Lc_{1}(\mathcal{E})=-H,c_{2}(\mathcal{E})=4L and c3()=0c_{3}(\mathcal{E})=0, where LL is the class of a line on XX. In addition, \mathcal{E} is exceptional and H(X,)=0H^{\bullet}(X,\mathcal{E})=0. In fact, \mathcal{E} is the pullback of the tautological bundle on the Grassmannian Gr(2,5)\operatorname{Gr}(2,5). By [13, Proposition 4.1], \mathcal{E} is the unique stable sheaf with c1()=H,c2()=4Lc_{1}(\mathcal{E})=-H,c_{2}(\mathcal{E})=4L and c3()=0c_{3}(\mathcal{E})=0.

Furthermore, there is a standard short exact sequence

0𝒪X5𝒬00\rightarrow\mathcal{E}\rightarrow\operatorname{\mathcal{O}}_{X}^{\oplus 5}\rightarrow\mathcal{Q}\rightarrow 0 (1)

where 𝒬\mathcal{Q} is the pull-back of the tautological quotient bundle on Gr(2,5)\operatorname{Gr}(2,5) along the natural map XGr(2,5)X\to\operatorname{Gr}(2,5). Since rk()=2\operatorname{rk}(\mathcal{E})=2, we have (H)\mathcal{E}(H)\cong\mathcal{E}^{\vee}.

Definition 3.1.

Let XX be a GM threefold.

  • The Kuznetsov component of XX is defined as 𝒦u(X):=,𝒪X\mathcal{K}u(X):=\langle\mathcal{E},\operatorname{\mathcal{O}}_{X}\rangle^{\bot}. In particular, it fits into the semiorthogonal decomposition Db(X)=𝒦u(X),,𝒪X\mathrm{D}^{b}(X)=\langle\mathcal{K}u(X),\mathcal{E},\operatorname{\mathcal{O}}_{X}\rangle;

  • The alternative Kuznetsov component of XX is defined as 𝒜X:=𝒪X,\mathcal{A}_{X}:=\langle\operatorname{\mathcal{O}}_{X},\mathcal{E}^{\vee}\rangle^{\bot}. In particular, it fits into the semiorthogonal decomposition Db(X)=𝒜X,𝒪X,\mathrm{D}^{b}(X)=\langle\mathcal{A}_{X},\operatorname{\mathcal{O}}_{X},\mathcal{E}^{\vee}\rangle.

Remark 3.2.

By [29, Proposition 2.6], there is a natural involutive autoequivalence functor τ𝒜:=S𝒜X[2]\tau_{\mathcal{A}}:=S_{\mathcal{A}_{X}}[-2] of 𝒜X\mathcal{A}_{X}. When XX is special, it is induced by the natural involution τ\tau on XX as τ𝒜=τ|𝒜X\tau_{\mathcal{A}}=\tau^{*}|_{\mathcal{A}_{X}}.

Definition 3.3.

The left adjoint to the inclusion 𝒜XDb(X)\mathcal{A}_{X}\hookrightarrow\mathrm{D}^{b}(X) is given by pr:=𝐋𝒪X𝐋:Db(X)𝒜X\mathrm{pr}:=\bm{\mathrm{L}}_{\operatorname{\mathcal{O}}_{X}}\bm{\mathrm{L}}_{\mathcal{E}^{\vee}}\colon\mathrm{D}^{b}(X)\rightarrow\mathcal{A}_{X}. We call this the projection functor.

The analogous natural projection functor can be defined for 𝒦u(X)\mathcal{K}u(X), and we denote it by pr:=𝐋𝐋𝒪X\mathrm{pr}^{\prime}:=\bm{\mathrm{L}}_{\mathcal{E}}\bm{\mathrm{L}}_{\operatorname{\mathcal{O}}_{X}}.

3.1. Kuznetsov components

Let K0(𝒟)K_{0}(\mathcal{D}) denote the Grothendieck group of a triangulated category 𝒟\mathcal{D}. We have the bilinear Euler form

χ(E,F)=i(1)iexti(E,F)\chi(E,F)=\sum_{i\in\mathbb{Z}}(-1)^{i}\operatorname{ext}^{i}(E,F)

for [E],[F]K0(𝒟)[E],[F]\in K_{0}(\mathcal{D}). By the Hirzebruch–Riemann–Roch formula, it takes the following form on GM threefolds. We have [33, p. 5] χ(u,v)=χ0(uv)\chi(u,v)=\chi_{0}(u^{*}\cap v) where uuu\mapsto u^{*} is an involution of i=03Hi(X,)\oplus_{i=0}^{3}H^{i}(X,\mathbb{Q}) given by multiplication with (1)i(-1)^{i} on H2i(X,)H^{2i}(X,\mathbb{Q}), and χ0\chi_{0} is given by

χ0(x+yH+zL+wP)=x+176y+12z+w,\chi_{0}(x+yH+zL+wP)=x+\frac{17}{6}y+\frac{1}{2}z+w,

where LL is the class of lines and PP is the class of points. The numerical Grothendieck group of 𝒟\mathcal{D} is 𝒩(𝒟)=K0(𝒟)/kerχ\mathcal{N}(\mathcal{D})=K_{0}(\mathcal{D})/\ker\chi.

Lemma 3.4 ([33, p. 5]).

The numerical Grothendieck group 𝒩(𝒦u(X))\mathcal{N}(\mathcal{K}u(X)) of the Kuznetsov component is a rank 22 integral lattice generated by the basis elements v=13L+12Pv=1-3L+\frac{1}{2}P and w=H6L+16Pw=H-6L+\frac{1}{6}P. Using this basis, χ\chi is given by the matrix

(2335).\begin{pmatrix}-2&-3\\ -3&-5\end{pmatrix}.

3.2. Alternative Kuznetsov components

As in [33, Proposition 3.9], the following lemma follows from a straightforward computation.

Lemma 3.5.

The numerical Grothendieck group of 𝒜X\mathcal{A}_{X} is a rank 2 integral lattice with basis vectors x=12Lx=1-2L and y=H4L56Py=H-4L-\frac{5}{6}P, and the Euler form with respect to the basis is

(1225).\begin{pmatrix}-1&-2\\ -2&-5\end{pmatrix}.
Remark 3.6.

It is straightforward to check that the (1)(-1)-classes of 𝒩(𝒜X)\mathcal{N}(\mathcal{A}_{X}) are x=12Lx=1-2L and 2xy=2H+56P2x-y=2-H+\frac{5}{6}P, up to sign.

Indeed, the Kuznetsov components from Subsection 3.1 and the alternative Kuznetsov components from this section are equivalent:

Lemma 3.7.

The original and alternative Kuznetsov components are equivalent. More precisely, there is an equivalence of categories Ξ:𝒦u(X)𝒜X\Xi\colon\mathcal{K}u(X)\xrightarrow{\sim}\mathcal{A}_{X} given by E𝐋𝒪X(E𝒪X(H))E\mapsto\bm{\mathrm{L}}_{\operatorname{\mathcal{O}}_{X}}(E\otimes\mathcal{O}_{X}(H)), with inverse given by F(𝐑𝒪XF)𝒪X(H)F\mapsto(\bm{\mathrm{R}}_{\operatorname{\mathcal{O}}_{X}}F)\otimes\mathcal{O}_{X}(-H).

Proof.

Using Lemma 2.7 and noting that 𝒪X(H)\mathcal{E}\otimes\operatorname{\mathcal{O}}_{X}(H)\cong\mathcal{E}^{\vee}, we manipulate the semiorthogonal decomposition as follows:

Db(X)\displaystyle\mathrm{D}^{b}(X) =𝒦u(X),,𝒪X\displaystyle=\langle\mathcal{K}u(X),\mathcal{E},\mathcal{O}_{X}\rangle
𝒦u(X)𝒪X(H),,𝒪X(H)\displaystyle\simeq\langle\mathcal{K}u(X)\otimes\mathcal{O}_{X}(H),\mathcal{E}^{\vee},\mathcal{O}_{X}(H)\rangle
𝒪X,𝒦u(X)𝒪X(H),\displaystyle\simeq\langle\mathcal{O}_{X},\mathcal{K}u(X)\otimes\mathcal{O}_{X}(H),\mathcal{E}^{\vee}\rangle
𝐋𝒪X(𝒦u(X)𝒪X(H)),𝒪X,.\displaystyle\simeq\langle\bm{\mathrm{L}}_{\operatorname{\mathcal{O}}_{X}}(\mathcal{K}u(X)\otimes\mathcal{O}_{X}(H)),\operatorname{\mathcal{O}}_{X},\mathcal{E}^{\vee}\rangle.

Now comparing with the definition of 𝒜X\mathcal{A}_{X}, we get 𝒜X𝐋𝒪X(𝒦u(X)𝒪X(H))\mathcal{A}_{X}\simeq\bm{\mathrm{L}}_{\operatorname{\mathcal{O}}_{X}}(\mathcal{K}u(X)\otimes\mathcal{O}_{X}(H)) and the desired result follows. The reverse direction is similar. ∎

4. Bridgeland stability conditions

In this section, we recall (weak) Bridgeland stability conditions on Db(X)\mathrm{D}^{b}(X), and the notions of tilt stability, double-tilt stability, and stability conditions induced on Kuznetsov components from weak stability conditions on Db(X)\mathrm{D}^{b}(X). We follow [5, § 2].

4.1. Weak stability conditions

Let 𝒟\mathcal{D} be a triangulated category, and K0(𝒟)K_{0}(\mathcal{D}) its Grothendieck group. Fix a surjective morphism v:K0(𝒟)Λv\colon K_{0}(\mathcal{D})\rightarrow\Lambda to a finite rank lattice.

Definition 4.1.

A stability condition (resp. weak stability condition) on 𝒟\mathcal{D} is a pair σ=(𝒜,Z)\sigma=(\mathcal{A},Z) where 𝒜\mathcal{A} is the heart of a bounded t-structure on 𝒟\mathcal{D}, and Z:ΛZ\colon\Lambda\rightarrow\mathbb{C} is a group homomorphism such that the following conditions hold:

  1. (1)

    The composition Zv:K0(𝒜)K0(𝒟)Z\circ v\colon K_{0}(\mathcal{A})\cong K_{0}(\mathcal{D})\rightarrow\mathbb{C} satisfies: for any E0𝒟E\neq 0\in\mathcal{D} we have ImZ(E)0\operatorname{Im}Z(E)\geq 0 and if ImZ(E)=0\operatorname{Im}Z(E)=0 then ReZ(E)<0\operatorname{Re}Z(E)<0 (resp. ReZ(E)0\operatorname{Re}Z(E)\leq 0). From now on, we write Z(E)Z(E) rather than Z(v(E))Z(v(E)).

We define a slope function μσ\mu_{\sigma} for σ\sigma using ZZ. For any E𝒜E\in\mathcal{A}, set

μσ(E):={ReZ(E)ImZ(E),ImZ(E)>0+,else.\mu_{\sigma}(E):=\begin{cases}-\frac{\operatorname{Re}Z(E)}{\operatorname{Im}Z(E)},&\operatorname{Im}Z(E)>0\\ +\infty,&\text{else}.\end{cases}

We say an object 0E𝒜0\neq E\in\mathcal{A} is σ\sigma-(semi)stable if μσ(F)<μσ(E/F)\mu_{\sigma}(F)<\mu_{\sigma}(E/F) (respectively μσ(F)μσ(E/F)\mu_{\sigma}(F)\leq\mu_{\sigma}(E/F)) for all proper subobjects FEF\subset E.

  1. (2)

    Any object E𝒜E\in\mathcal{A} has a Harder–Narasimhan filtration in terms of σ\sigma-semistability defined above.

  2. (3)

    There exists a quadratic form QQ on Λ\Lambda\otimes\mathbb{R} such that Q|kerZQ|_{\ker Z} is negative definite, and Q(E)0Q(E)\geq 0 for all σ\sigma-semistable objects E𝒜E\in\mathcal{A}. This is known as the support property.

Definition 4.2.

Let σ=(𝒜,Z)\sigma=(\mathcal{A},Z) be a stability condition on 𝒟\mathcal{D}. The phase of a σ\sigma-semistable object E𝒜E\in\mathcal{A} is

ϕ(E):=1πarg(Z(E))(0,1].\phi(E):=\frac{1}{\pi}\mathrm{arg}(Z(E))\in(0,1].

Specially, if Z(E)=0Z(E)=0 then ϕ(E)=1\phi(E)=1. If F=E[n]F=E[n], then we define

ϕ(F):=ϕ(E)+n\phi(F):=\phi(E)+n

A slicing 𝒫\mathcal{P} of 𝒟\mathcal{D} consists of full additive subcategories 𝒫(ϕ)𝒟\mathcal{P}(\phi)\subset\mathcal{D} for each ϕ\phi\in\mathbb{R} satisfying

  1. (4)

    for ϕ(0,1]\phi\in(0,1], the subcategory 𝒫(ϕ)\mathcal{P}(\phi) is given by the zero object and all σ\sigma-semistable objects whose phase is ϕ\phi;

  2. (5)

    for ϕ+n\phi+n with ϕ(0,1]\phi\in(0,1] and nn\in\mathbb{Z}, we set 𝒫(ϕ+n):=𝒫(ϕ)[n]\mathcal{P}(\phi+n):=\mathcal{P}(\phi)[n].

We will use both notations σ=(𝒜,Z)\sigma=(\mathcal{A},Z) and σ=(𝒫,Z)\sigma=(\mathcal{P},Z) for a stability condition σ\sigma with heart 𝒜=𝒫((0,1])\mathcal{A}=\mathcal{P}((0,1]) where 𝒫\mathcal{P} is the slicing of σ\sigma.

We say σ\sigma is a numerical stability condition on 𝒟\mathcal{D} if the surjective morphism v:K0(𝒟)Λv\colon K_{0}(\mathcal{D})\to\Lambda factors through the natural surjection K0(𝒟)𝒩(𝒟)K_{0}(\mathcal{D})\twoheadrightarrow\mathcal{N}(\mathcal{D}) (assuming 𝒩(𝒟)\mathcal{N}(\mathcal{D}) is well-defined).

Next, we recall two natural group actions on the set of stability conditions Stab(𝒟)\operatorname{Stab}(\mathcal{D}).

  1. (1)

    An element g~=(g,G)\tilde{g}=(g,G) in the universal covering GL~+(2,)\widetilde{\mathrm{GL}}^{+}(2,\mathbb{R}) of the group GL+(2,)\mathrm{GL}^{+}(2,\mathbb{R}) consists of an increasing function g:g\colon\mathbb{R}\rightarrow\mathbb{R} such that g(ϕ+1)=g(ϕ)+1g(\phi+1)=g(\phi)+1 and a matrix GGL+(2,)G\in\mathrm{GL}^{+}(2,\mathbb{R}) with det(G)>0\det(G)>0. It acts on the right on the stability manifold by σg~:=(G1Z,𝒫(g(ϕ)))\sigma\cdot\tilde{g}:=(G^{-1}\circ Z,\mathcal{P}(g(\phi))) for any σ=(𝒫,Z)Stab(𝒟)\sigma=(\mathcal{P},Z)\in\operatorname{Stab}(\mathcal{D}) (see [11, Lemma 8.2]).

  2. (2)

    Let AutΛ(𝒟)\mathrm{Aut}_{\Lambda}(\mathcal{D}) be the group of exact autoequivalences of 𝒟\mathcal{D}, whose action Φ\Phi_{*} on K0(𝒟)K_{0}(\mathcal{D}) is compatible with v:K0(𝒟)Λv\colon K_{0}(\mathcal{D})\to\Lambda. For ΦAutΛ(𝒟)\Phi\in\mathrm{Aut}_{\Lambda}(\mathcal{D}) and σ=(𝒫,Z)Stab(𝒟)\sigma=(\mathcal{P},Z)\in\mathrm{Stab}(\mathcal{D}), we define a left action of the group of linear exact autoequivalences AutΛ(𝒟)\mathrm{Aut}_{\Lambda}(\mathcal{D}) by Φσ=(Φ(𝒫),ZΦ1)\Phi\cdot\sigma=(\Phi(\mathcal{P}),Z\circ\Phi_{*}^{-1}), where Φ\Phi_{*} is the automorphism of K0(𝒟)K_{0}(\mathcal{D}) induced by Φ\Phi.

4.2. Tilt-stability

Let (X,H)(X,H) be a polarised smooth projective variety of dimension nn and σH=(Coh(X),ZH)\sigma_{H}=(\operatorname{\mathrm{Coh}}(X),Z_{H}) be the standard weak stability condition on Coh(X)\operatorname{\mathrm{Coh}}(X) defined as

ZH(E):=Hn1ch1(E)+𝔦Hnrk(E).Z_{H}(E):=-H^{n-1}\mathrm{ch}_{1}(E)+\mathfrak{i}H^{n}\operatorname{rk}(E).

Its σH\sigma_{H}-stability coincides with classical μH\mu_{H}-stability (slope stability). Now for a fixed real number β\beta, consider the following subcategories111The angle brackets here mean extension closure. of Coh(X)\operatorname{\mathrm{Coh}}(X):

𝒯β\displaystyle\mathcal{T}^{\beta} =ECoh(X)E is σH-semistable with μσH(E)>β\displaystyle=\langle E\in\operatorname{\mathrm{Coh}}(X)\mid E\text{ is }\sigma_{H}\text{-semistable with }\mu_{\sigma_{H}}(E)>\beta\rangle
β\displaystyle\operatorname{\mathcal{F}}^{\beta} =ECoh(X)E is σH-semistable with μσH(E)β.\displaystyle=\langle E\in\operatorname{\mathrm{Coh}}(X)\mid E\text{ is }\sigma_{H}\text{-semistable with }\mu_{\sigma_{H}}(E)\leq\beta\rangle.

Then it is a result of [20] that the tilted heart Cohβ(X):=𝒯β,β[1]\operatorname{\mathrm{Coh}}^{\beta}(X):=\langle\mathcal{T}^{\beta},\operatorname{\mathcal{F}}^{\beta}[1]\rangle is the heart of a bounded t-structure on Coh(X)\operatorname{\mathrm{Coh}}(X).

Proposition 4.3 ([9, 8]).

Let α>0\alpha>0 and β\beta\in\mathbb{R}. Then the pair σα,β=(Cohβ(X),Zα,β)\sigma_{\alpha,\beta}=(\operatorname{\mathrm{Coh}}^{\beta}(X),Z_{\alpha,\beta}) defines a weak stability condition on Db(X)\mathrm{D}^{b}(X), where

Zα,β(E)=12α2Hnch0β(E)Hn2ch2β(E)+𝔦Hn1ch1β(E).Z_{\alpha,\beta}(E)=\frac{1}{2}\alpha^{2}H^{n}\mathrm{ch}_{0}^{\beta}(E)-H^{n-2}\mathrm{ch}_{2}^{\beta}(E)+\mathfrak{i}H^{n-1}\mathrm{ch}_{1}^{\beta}(E).

The quadratic form QQ is given by the discriminant

ΔH(E)=(Hn1ch1(E))22Hnch0(E)Hn2ch2(E).\Delta_{H}(E)=(H^{n-1}\mathrm{ch}_{1}(E))^{2}-2H^{n}\mathrm{ch}_{0}(E)H^{n-2}\mathrm{ch}_{2}(E).

We denote the slope function by μα,β:=μσα,β\mu_{\alpha,\beta}:=\mu_{\sigma_{\alpha,\beta}}.

The weak stability conditions σα,β\sigma_{\alpha,\beta} constructed above are also known as tilt-stability and the heart Cohβ(X)\operatorname{\mathrm{Coh}}^{\beta}(X) are called the tilted heart.

Now pick a weak stability condition σα,β\sigma_{\alpha,\beta}. We define

𝒯α,β0\displaystyle\mathcal{T}^{0}_{\alpha,\beta} =ECohβ(X)E is σα,β-semistable with μα,β(E)>0\displaystyle=\langle E\in\operatorname{\mathrm{Coh}}^{\beta}(X)\mid E\text{ is }\sigma_{\alpha,\beta}\text{-semistable with }\mu_{\alpha,\beta}(E)>0\rangle
α,β0\displaystyle\operatorname{\mathcal{F}}^{0}_{\alpha,\beta} =ECohβ(X)E is σα,β-semistable with μα,β(E)0.\displaystyle=\langle E\in\operatorname{\mathrm{Coh}}^{\beta}(X)\mid E\text{ is }\sigma_{\alpha,\beta}\text{-semistable with }\mu_{\alpha,\beta}(E)\leq 0\rangle.

Moreover, we “rotate” the stability function Zα,βZ_{\alpha,\beta} by setting

Zα,β0:=1𝔦Zα,β.Z_{\alpha,\beta}^{0}:=\frac{1}{\mathfrak{i}}Z_{\alpha,\beta}.

Then we have the following result:

Proposition 4.4 ([5, Proposition 2.15]).

The pair σα,β0=(Cohα,β0(X)=𝒯α,β0,α,β0[1],Zα,β0)\sigma^{0}_{\alpha,\beta}=(\operatorname{\mathrm{Coh}}^{0}_{\alpha,\beta}(X)=\langle\mathcal{T}^{0}_{\alpha,\beta},\operatorname{\mathcal{F}}^{0}_{\alpha,\beta}[1]\rangle,Z_{\alpha,\beta}^{0}) defines a weak stability condition on Db(X)\mathrm{D}^{b}(X). We denote the slope function by μα,β0:=μσα,β0\mu^{0}_{\alpha,\beta}:=\mu_{\sigma^{0}_{\alpha,\beta}}.

We now state a useful lemma that relates 2-Gieseker-stability (see [3, Definition 4.3]) and tilt-stability.

Lemma 4.5 ([8, Lemma 2.7], [3, Proposition 4.8, 4.9]).

Let EDb(X)E\in\mathrm{D}^{b}(X).

  1. (1)

    Let β<μ(E)\beta<\mu(E). Then ECohβ(X)E\in\operatorname{\mathrm{Coh}}^{\beta}(X) is σα,β\sigma_{\alpha,\beta}-(semi)stable for α0\alpha\gg 0 if and only if ECoh(X)E\in\operatorname{\mathrm{Coh}}(X) and EE is 2-Gieseker-(semi)stable.

  2. (2)

    If ECohβ(X)E\in\operatorname{\mathrm{Coh}}^{\beta}(X) is σα,β\sigma_{\alpha,\beta}-semistable for βμ(E)\beta\geq\mu(E) and α0\alpha\gg 0, then 1(E)\mathcal{H}^{-1}(E) is a torsion free μ\mu-semistable sheaf and 0(E)\mathcal{H}^{0}(E) is supported in dimension not greater than one. If β>μ(E)\beta>\mu(E) and α>0\alpha>0, then 1(E)\mathcal{H}^{-1}(E) is also reflexive.

4.3. Stronger BG inequalities

In this subsection, we state stronger Bogomolov–Gieseker (BG) style inequalities, which hold for tilt-semistable objects. These will be useful later on for ruling out potential walls for tilt-stability of objects in Db(X)\mathrm{D}^{b}(X). The first is a stronger version of Proposition 4.3, which was proved by Chunyi Li in [36, Proposition 3.2] for Fano threefolds of Picard number one.

Lemma 4.6 (Stronger BG I).

Let XX be an index 11 prime Fano threefold with degree dd, and EDb(X)E\in\mathrm{D}^{b}(X) a σα,β\sigma_{\alpha,\beta}-stable object where α>0\alpha>0. Let k:=μ(E)k:=\lfloor\mu(E)\rfloor. Then we have:

Hch2(E)H3ch0(E)max{kμH(E)k22,12μH(E)234d,(k+1)μH(E)(k+1)22}.\frac{H\cdot\mathrm{ch}_{2}(E)}{H^{3}\cdot\mathrm{ch}_{0}(E)}\leq\max\left\{k\mu_{H}(E)-\frac{k^{2}}{2},\frac{1}{2}\mu_{H}(E)^{2}-\frac{3}{4d},(k+1)\mu_{H}(E)-\frac{(k+1)^{2}}{2}\right\}.

Moreover, if the equality holds, then EE has rank one or two.

The second is due to Naoki Koseki and Chunyi Li. It is based on [27, Lemma 4.2, Theorem 4.3]. Chunyi Li also sent us a similar inequality from his upcoming paper [37].

Theorem 4.7 (Stronger BG II).

Let XX be an index 11 Fano threefold of degree dd, and ECoh0(X)E\in\operatorname{\mathrm{Coh}}^{0}(X) be a σα,0\sigma_{\alpha,0}-semistable object for some α>0\alpha>0 with |μH(E)|[0,1]|\mu_{H}(E)|\in[0,1] and rk(E)2\operatorname{rk}(E)\geq 2. Then

Hch2(E)H3ch0(E)max{12μH(E)234d,μH(E)212|μH(E)|}.\frac{H\cdot\mathrm{ch}_{2}(E)}{H^{3}\cdot\mathrm{ch}_{0}(E)}\leq\max\left\{\frac{1}{2}\mu_{H}(E)^{2}-\frac{3}{4d},\mu_{H}(E)^{2}-\frac{1}{2}|\mu_{H}(E)|\right\}.

Before we prove Theorem 4.7, we first state an easy lemma.

Lemma 4.8.

Let SS be a K3 surface of degree dd and HSH_{S} the ample polarisation. Let EE be a μHS\mu_{H_{S}}-semistable sheaf in Db(S)\mathrm{D}^{b}(S) with rk(E)2\operatorname{rk}(E)\geq 2. Then

ch2(E)HS2rk(E)12μHS(E)234d.\frac{\mathrm{ch}_{2}(E)}{H_{S}^{2}\cdot\operatorname{rk}(E)}\leq\frac{1}{2}\mu_{H_{S}}(E)^{2}-\frac{3}{4d}.
Proof.

Let v(E)v(E) be the Mukai vector of EE. We have

v(E)2\displaystyle v(E)^{2} =HS2ch1(E)22rk(E)22rk(E)ch2(E)\displaystyle=H_{S}^{2}\cdot\mathrm{ch}_{1}(E)^{2}-2\operatorname{rk}(E)^{2}-2\operatorname{rk}(E)\cdot\mathrm{ch}_{2}(E)
212rk(E)2.\displaystyle\geq-2\geq-\frac{1}{2}\operatorname{rk}(E)^{2}.

Dividing through by rk(E)2\operatorname{rk}(E)^{2} and rearranging, we get

ch2(E)rk(E)\displaystyle\frac{\mathrm{ch}_{2}(E)}{\operatorname{rk}(E)} 12μHS(E)2HS234\displaystyle\leq\frac{1}{2}\mu_{H_{S}}(E)^{2}H_{S}^{2}-\frac{3}{4}

as required. ∎

Proof of Theorem 4.7.

Let f:[0,1]f\colon[0,1]\rightarrow\mathbb{R} be defined as

f(t):=max{12t234d,t212t}f(t):=\max\left\{\frac{1}{2}t^{2}-\frac{3}{4d},t^{2}-\frac{1}{2}t\right\}

Note that ff is star-shaped [27, Definition 3.2] and satisfies f(0)=0f(0)=0 and f(1)=1/2f(1)=1/2 as well as

t212tf(t)12t2t^{2}-\frac{1}{2}t\leq f(t)\leq\frac{1}{2}t^{2}

for all t[0,1]t\in[0,1]. We now follow the strategy of proof in [27, Theorem 4.3]. Assume for a contradiction that there is an EDb(X)E\in\mathrm{D}^{b}(X) such that the inequality in the statement of Theorem 4.7 is not true. Then conditions (a) and (b) in [27, Lemma 3.3] are satisfied for ff. Then by loc. cit., the restriction E|SdE|_{S_{d}} where SdS_{d} is a general hyperplane section of XdX_{d} is μHSd\mu_{H_{S_{d}}}-semistable. Also note that μHSd(E|Sd)=μH(E)\mu_{H_{S_{d}}}(E|_{S_{d}})=\mu_{H}(E) and SdS_{d} is a smooth K3 surface. But then by assumption

ch2(E|Sd)HSd2rk(E|Sd)>12μHSd(E)234d\frac{\mathrm{ch}_{2}(E|_{S_{d}})}{H_{S_{d}}^{2}\cdot\operatorname{rk}(E|_{S_{d}})}>\frac{1}{2}\mu_{H_{S_{d}}}(E)^{2}-\frac{3}{4d}

which contradicts Proposition 4.8, so the assumption is false and the result follows. ∎

4.4. Stability conditions on the Kuznetsov component of a GM threefold

Proposition 5.1 in [5] gives a criterion for checking when weak stability conditions on a triangulated category can be used to induce stability conditions on a subcategory. Each of the criteria of this proposition can be checked for 𝒜XDb(X)\mathcal{A}_{X}\subset\mathrm{D}^{b}(X) to give stability conditions on 𝒜X\mathcal{A}_{X}.

More precisely, let 𝒜(α,β)=Cohα,β0(X)𝒜X\mathcal{A}(\alpha,\beta)=\operatorname{\mathrm{Coh}}^{0}_{\alpha,\beta}(X)\cap\mathcal{A}_{X} and Z(α,β)=Zα,β0|𝒜XZ(\alpha,\beta)=Z_{\alpha,\beta}^{0}|_{\mathcal{A}_{X}}. Furthermore, if we take suitable (α,β)(\alpha,\beta), by [5, Theorem 6.9] and [48, Proposition 3.2] we have:

Theorem 4.9.

Let XX be a GM threefold. Then σ(α,β)\sigma(\alpha,\beta) is a stability condition on 𝒜X\mathcal{A}_{X} for all (α,β)V(\alpha,\beta)\in V, where

V:={(α,β):110<β<0,0<α<β}.V:=\{(\alpha,\beta)\colon-\frac{1}{10}<\beta<0,0<\alpha<-\beta\}.

Now we introduce a special class of stability condition, which will play a central role in our paper.

Definition 4.10.

Let σ\sigma be a stability condition on a triangulated category 𝒟\mathcal{D}. It is called Serre-invariant if S𝒟σ=σgS_{\mathcal{D}}\cdot\sigma=\sigma\cdot g for some gGL~+(2,)g\in\widetilde{\mathrm{GL}}^{+}(2,\mathbb{R}), where S𝒟S_{\mathcal{D}} is the Serre functor of 𝒟\mathcal{D}.

We recall a recent result proved in [48].

Theorem 4.11.

Let XX be a GM threefold and σ\sigma (resp. σ\sigma^{\prime}) be a stability condition on 𝒦u(X)\mathcal{K}u(X) (resp. 𝒜X\mathcal{A}_{X}) defined by [5]. Then σ\sigma (resp. σ\sigma^{\prime}) is Serre-invariant.

Proposition 4.12.

Let XX be a GM threefold and EE a non-zero object in 𝒜X\mathcal{A}_{X} such that ext1(E,E)3\mathrm{ext}^{1}(E,E)\leq 3 and χ(E,E)-\chi(E,E) is not a perfect square. Then EE is σ\sigma-stable for every Serre-invariant stability condition σ\sigma on 𝒜X\mathcal{A}_{X}.

Proof.

The proof is the same as in [54, Lemma 9.12]. We omit the details. ∎

5. Projection of \mathcal{E} into 𝒦u(X)\mathcal{K}u(X)

In this section, we consider the object that results from projecting the vector bundle \mathcal{E} into 𝒦u(X)\mathcal{K}u(X), and its stability in 𝒦u(X)\mathcal{K}u(X). We start with a lemma.

Lemma 5.1.

Let XX be a GM threefold.

  1. (1)

    RHom(𝒬(H),)=RHom(,𝒬)=k2\operatorname{RHom}^{\bullet}(\mathcal{Q}(-H),\mathcal{E})=\operatorname{RHom}^{\bullet}(\mathcal{E},\mathcal{Q}^{\vee})=k^{2} when XX is ordinary.

  2. (2)

    RHom(𝒬(H),)=RHom(,𝒬)=k3k[1]\operatorname{RHom}^{\bullet}(\mathcal{Q}(-H),\mathcal{E})=\operatorname{RHom}^{\bullet}(\mathcal{E},\mathcal{Q}^{\vee})=k^{3}\oplus k[-1] when XX is special.

  3. (3)

    RHom(,𝒬(H))=k[2]\operatorname{RHom}^{\bullet}(\mathcal{E},\mathcal{Q}(-H))=k[-2].

  4. (4)

    RHom(𝒬,)=RHom(,𝒬)=k[2]\mathrm{RHom}^{\bullet}(\mathcal{Q}^{\vee},\mathcal{E})=\mathrm{RHom}^{\bullet}(\mathcal{E}^{\vee},\mathcal{Q})=k[-2].

Proof.

When XX is ordinary, (1) and (2) follow from the Koszul resolution of XGr(2,5)X\subset\mathrm{Gr}(2,5) and the Borel–Weil–Bott Theorem. When XX is special with the double cover π:XY\pi\colon X\to Y, note that π𝒪X=𝒪Y𝒪Y(1)\pi_{*}\operatorname{\mathcal{O}}_{X}=\operatorname{\mathcal{O}}_{Y}\oplus\operatorname{\mathcal{O}}_{Y}(-1). Then the (1) and (2) follow from the projection formula and [53, Lemma 2.14, Proposition 2.15]. And applying Hom(,)\operatorname{Hom}(-,\mathcal{E}) to (1) and using Serre duality, we get RHom(,𝒬(H))=RHom(𝒬,)[3]=k[2]\operatorname{RHom}^{\bullet}(\mathcal{E},\mathcal{Q}(-H))=\mathrm{RHom}^{\bullet}(\mathcal{Q},\mathcal{E})^{\vee}[-3]=k[-2], which proves (3). Finally, (4) follows from applying Hom(,)\operatorname{Hom}(-,\mathcal{E}) to (1) and using Serre duality and RHom(,)=k\mathrm{RHom}^{\bullet}(\mathcal{E},\mathcal{E})=k. ∎

5.1. The projection of \mathcal{E} into 𝒦u(X)\mathcal{K}u(X)

Let π:=𝐑𝒪X(H)𝐑(H):Db(X)𝒦u(X)\pi:=\bm{\mathrm{R}}_{\operatorname{\mathcal{O}}_{X}(-H)}\bm{\mathrm{R}}_{\mathcal{E}(-H)}\colon\mathrm{D}^{b}(X)\rightarrow\mathcal{K}u(X) be the right adjoint to the inclusion 𝒦u(X)Db(X)\mathcal{K}u(X)\hookrightarrow\mathrm{D}^{b}(X). Here 𝒦u(X)=,𝒪X\mathcal{K}u(X)=\langle\mathcal{E},\operatorname{\mathcal{O}}_{X}\rangle^{\perp} is the original Kuznetsov component.

Lemma 5.2.

The projection object π()\pi(\operatorname{\mathcal{E}}) is the unique object that fits into a non-trivial exact triangle

𝒬(H)[1]π().\mathcal{Q}(-H)[1]\to\pi(\mathcal{E})\to\mathcal{E}. (2)
Proof.

By Serre duality, we have RHom(,(H))=RHom(,)[3]=k[3]\operatorname{RHom}^{\bullet}(\mathcal{E},\mathcal{E}(-H))=\operatorname{RHom}^{\bullet}(\mathcal{E},\mathcal{E})^{\vee}[-3]=k[-3]. Then we have an exact triangle 𝐑(H)(H)[3]\bm{\mathrm{R}}_{\mathcal{E}(-H)}\mathcal{E}\to\mathcal{E}\to\mathcal{E}(-H)[3]. And by (1), we see 𝐑𝒪X(H)(H)=𝒬(H)[1]\bm{\mathrm{R}}_{\operatorname{\mathcal{O}}_{X}(-H)}\mathcal{E}(-H)=\mathcal{Q}(-H)[-1]. Thus, from 𝐑𝒪X(H)=\bm{\mathrm{R}}_{\operatorname{\mathcal{O}}_{X}(-H)}\mathcal{E}=\mathcal{E} we obtain the triangle (2). It is non-trivial since π()𝒦u(X)\pi(\mathcal{E})\in\mathcal{K}u(X), so \mathcal{E} cannot be a direct summand of π()\pi(\mathcal{E}). Finally, the uniqueness follows from Lemma 5.1 (4). ∎

Lemma 5.3.

Let XX be a GM threefold. Then we have

  • RHom(π(),π())=kk2[1]\operatorname{RHom}^{\bullet}(\pi(\mathcal{E}),\pi(\mathcal{E}))=k\oplus k^{2}[-1] when XX is ordinary.

  • RHom(π(),π())=kk3[1]k[2]\operatorname{RHom}^{\bullet}(\pi(\mathcal{E}),\pi(\mathcal{E}))=k\oplus k^{3}[-1]\oplus k[-2] when XX is special.

Hence π()\pi(\mathcal{E}) is stable with respect to every Serre-invariant stability condition on 𝒦u(X)\mathcal{K}u(X).

Proof.

The first statement follows from applying Hom(,)\operatorname{Hom}(-,\mathcal{E}) to triangle (2) and Lemma 5.1, and also the fact that RHom(π(),π())=RHom(π(),)\operatorname{RHom}^{\bullet}(\pi(\mathcal{E}),\pi(\mathcal{E}))=\operatorname{RHom}^{\bullet}(\pi(\mathcal{E}),\mathcal{E}) which is by adjunction. The last statement follows from Lemma 4.12. ∎

5.2. The analogous projection object for 𝒜X\mathcal{A}_{X}

In this subsection, we state and prove the analogous results as in Subsection 5.1, except for 𝒜X\mathcal{A}_{X} instead of 𝒦u(X)\mathcal{K}u(X). Let π:=𝐑𝐑𝒪X(H):Db(X)𝒜X\pi^{\prime}:=\bm{\mathrm{R}}_{\mathcal{E}}\bm{\mathrm{R}}_{\operatorname{\mathcal{O}}_{X}(-H)}\colon\mathrm{D}^{b}(X)\rightarrow\mathcal{A}_{X} be the right adjoint to the inclusion 𝒜XDb(X)\mathcal{A}_{X}\hookrightarrow\mathrm{D}^{b}(X).

Lemma 5.4.

The projection object π(𝒬)\pi^{\prime}(\mathcal{Q}^{\vee}) is the unique object fits into a non-trivial exact triangle

[1]π(𝒬)𝒬.\mathcal{E}[1]\to\pi^{\prime}(\mathcal{Q}^{\vee})\to\mathcal{Q}^{\vee}. (3)
Proof.

The proof is completely analogous to the proof of Lemma 5.2. By Serre duality, we have the vanishing RHom(𝒬,𝒪X(H))=RHom(𝒪X,𝒬)=0\mathrm{RHom}^{\bullet}(\mathcal{Q}^{\vee},\operatorname{\mathcal{O}}_{X}(-H))=\mathrm{RHom}^{\bullet}(\operatorname{\mathcal{O}}_{X},\mathcal{Q}^{\vee})^{\vee}=0. Thus π(𝒬)=𝐑𝒬\pi^{\prime}(\mathcal{Q}^{\vee})=\bm{\mathrm{R}}_{\mathcal{E}}\mathcal{Q}^{\vee}. Then the result follows from Lemma 5.1 (4). Finally, the uniqueness also follows from Lemma 5.1 (4) and (3) is non-trivial since RHom(𝒬,π(𝒬))=0\mathrm{RHom}^{\bullet}(\mathcal{Q}^{\vee},\pi^{\prime}(\mathcal{Q}^{\vee}))=0. ∎

Remark 5.5.

Later in Section 7, we will see that we have π(𝒬)pr(IC)[1]\pi^{\prime}(\mathcal{Q}^{\vee})\cong\mathrm{pr}(I_{C})[1] where CXC\subset X is a conic such that IC𝒜XI_{C}\not\in\mathcal{A}_{X}.

Lemma 5.6.

Let XX be a GM threefold. Then

  • RHom(π(𝒬),π(𝒬))=kk2[1]\operatorname{RHom}^{\bullet}(\pi^{\prime}(\mathcal{Q}^{\vee}),\pi^{\prime}(\mathcal{Q}^{\vee}))=k\oplus k^{2}[-1] when XX is ordinary.

  • RHom(π(𝒬),π(𝒬))=kk3[1]k[2]\operatorname{RHom}^{\bullet}(\pi^{\prime}(\mathcal{Q}^{\vee}),\pi^{\prime}(\mathcal{Q}^{\vee}))=k\oplus k^{3}[-1]\oplus k[-2] when XX is special.

Hence π(𝒬)\pi^{\prime}(\mathcal{Q}^{\vee}) is stable with respect to every Serre-invariant stability condition on 𝒜X\mathcal{A}_{X}.

Proof.

It is not hard to check that Ξ(π())π(𝒬)[1]\Xi(\pi(\mathcal{E}))\cong\pi^{\prime}(\mathcal{Q}^{\vee})[1], where Ξ\Xi is the equivalence 𝒦u(X)𝒜X\mathcal{K}u(X)\simeq\mathcal{A}_{X} in Lemma 3.7. Then the result follows from Lemma 5.3. ∎

6. Conics on GM threefolds

In this section, we collect some useful results regarding the birational geometry of GM threefolds and their Hilbert schemes of conics. The results in this section are all from [13], [39], and [22].

Recall that a conic means a closed subscheme CXC\subset X with Hilbert polynomial pC(t)=1+2tp_{C}(t)=1+2t, and a line means a closed subscheme LXL\subset X with Hilbert polynomial pL(t)=1+tp_{L}(t)=1+t. Denote their Hilbert schemes by 𝒞(X)\mathcal{C}(X) and Γ(X)\Gamma(X), respectively.

6.1. Conics on ordinary GM threefolds

Let XX be an ordinary GM threefold. Recall that it is a quadric section of a linear section of codimension 22 of the Grassmannian Gr(2,5)=Gr(2,V5)\operatorname{Gr}(2,5)=\operatorname{Gr}(2,V_{5}), where V5V_{5} is a 55-dimensional complex vector space. Let ViV_{i} be an ii-dimensional vector subspace of V5V_{5}. There are two types of 22-planes in Gr(2,5)\operatorname{Gr}(2,5); σ\sigma-planes are given set-theoretically as {[V2]V1V2V4}\{[V_{2}]\mid V_{1}\subset V_{2}\subset V_{4}\}, and ρ\rho-planes are given by {[V2]V2V3}\{[V_{2}]\mid V_{2}\subset V_{3}\}.

Remark 6.1.

In [13, Section 3.1], the σ\sigma-planes and ρ\rho-planes are called α\alpha-planes and β\beta-planes, respectively.

By [13, Section 3.1] and [24, Section 3.1], we have the following classification of conics on XX.

Definition 6.2 ([13, p. 5]).
  • A conic CXC\subset X is called a τ\tau-conic if the 22-plane C\langle C\rangle is not contained in Gr(2,V5)\operatorname{Gr}(2,V_{5}), there is a unique V4V5V_{4}\subset V_{5} such that CGr(2,V4)C\subset\operatorname{Gr}(2,V_{4}), the conic CC is reduced and if it is smooth, the union of corresponding lines in (V5)\mathbb{P}(V_{5}) is a smooth quadric surface in (V4)\mathbb{P}(V_{4}).

  • A conic CXC\subset X is called a σ\sigma-conic if the 22-plane C\langle C\rangle spanned by CC is an σ\sigma-plane, and if there is a unique hyperplane V4V5V_{4}\subset V_{5} such that CGr(2,V4)C\subset\operatorname{Gr}(2,V_{4}) and the union of the corresponding lines in (V5)\mathbb{P}(V_{5}) is a quadric cone in (V4)\mathbb{P}(V_{4}).

  • A conic CXC\subset X is called a ρ\rho-conic if the 22-plane C\langle C\rangle spanned by CC is a ρ\rho-plane, and the union of corresponding lines in (V5)\mathbb{P}(V_{5}) is this 22-plane.

The following lemma is very useful for computations:

Lemma 6.3.

Let XX be an ordinary GM threefold and CC be a conic on XX.

  1. (1)

    If CC is a τ\tau-conic, then we have RHom(,IC)=k\operatorname{RHom}^{\bullet}(\mathcal{E},I_{C})=k and RHom(,IC)=0\operatorname{RHom}^{\bullet}(\mathcal{E}^{\vee},I_{C})=0.

  2. (2)

    If CC is a ρ\rho-conic, then we have RHom(,IC)=k2k[1]\operatorname{RHom}^{\bullet}(\mathcal{E},I_{C})=k^{2}\oplus k[-1] and RHom(,IC)=0\operatorname{RHom}^{\bullet}(\mathcal{E}^{\vee},I_{C})=0.

  3. (3)

    If CC is a σ\sigma-conic, then we have RHom(,IC)=k\operatorname{RHom}^{\bullet}(\mathcal{E},I_{C})=k and RHom(,IC)=k[1]k[2]\operatorname{RHom}^{\bullet}(\mathcal{E}^{\vee},I_{C})=k[-1]\oplus k[-2].

Proof.

Note that if Hom(,IC)=ka\operatorname{Hom}(\mathcal{E},I_{C})=k^{a}, then CGr(2,5a)XC\subset\mathrm{Gr}(2,5-a)\cap X. Since for any conic CC, there is some V4V_{4} such that CGr(2,V4)C\subset\mathrm{Gr}(2,V_{4}), then we have hom(,IC)1\hom(\mathcal{E},I_{C})\geq 1 for any conic CC. Now if hom(,IC)2\hom(\mathcal{E},I_{C})\geq 2, we know that CC is contained in a ρ\rho-plane Gr(2,3)\mathrm{Gr}(2,3). Since C\langle C\rangle is not in Gr(2,5)\mathrm{Gr}(2,5) for a τ\tau-conic CC, and C\langle C\rangle is a σ\sigma-plane {V2|V1V2V4}\{V_{2}|V_{1}\subset V_{2}\subset V_{4}\} for a σ\sigma-conic, for these two types of conics we have Hom(,IC)=k\operatorname{Hom}(\mathcal{E},I_{C})=k. Also, for a ρ\rho-conic CC, since C=Gr(2,3)\langle C\rangle=\mathrm{Gr}(2,3), we have hom(,IC)2\hom(\mathcal{E},I_{C})\geq 2. But if hom(,IC)3\hom(\mathcal{E},I_{C})\geq 3, we know that CGr(2,2)C\subset\mathrm{Gr}(2,2) which is impossible. Hence for a ρ\rho-conic CC we have Hom(,IC)=k2\operatorname{Hom}(\mathcal{E},I_{C})=k^{2}. Now the result for Ext groups follows from applying Hom(,)\operatorname{Hom}(\mathcal{E},-) to the short exact sequence 0IC𝒪X𝒪C00\to I_{C}\to\operatorname{\mathcal{O}}_{X}\to\operatorname{\mathcal{O}}_{C}\to 0 and χ(,IC)=1\chi(\mathcal{E},I_{C})=1.

First by stability and Serre duality, we have Hom(,IC)=Ext3(,IC)=0\operatorname{Hom}(\mathcal{E}^{\vee},I_{C})=\operatorname{Ext}^{3}(\mathcal{E}^{\vee},I_{C})=0. From χ(,IC)=0\chi(\mathcal{E}^{\vee},I_{C})=0, we only need to compute Ext1(,IC)\operatorname{Ext}^{1}(\mathcal{E}^{\vee},I_{C}). Since RHom(𝒪X,IC)=0\operatorname{RHom}^{\bullet}(\operatorname{\mathcal{O}}_{X},I_{C})=0, applying Hom(,IC)\operatorname{Hom}(-,I_{C}) to the tautological sequence, we have Hom(𝒬,IC)=Ext1(,IC)\operatorname{Hom}(\mathcal{Q}^{\vee},I_{C})=\operatorname{Ext}^{1}(\mathcal{E}^{\vee},I_{C}). Note that if Hom(𝒬,IC)=ka\operatorname{Hom}(\mathcal{Q}^{\vee},I_{C})=k^{a}, then CGr(2a,5a)XC\subset\mathrm{Gr}(2-a,5-a)\cap X. Thus we have hom(𝒬,IC)1\hom(\mathcal{Q}^{\vee},I_{C})\leq 1 for any conic CC. And since hom(𝒬,IC)=1\hom(\mathcal{Q}^{\vee},I_{C})=1 if and only if CC is contained in the zero locus of a global section of 𝒬\mathcal{Q}, which is a σ\sigma-3-plane in Gr(2,5)\mathrm{Gr}(2,5), we know that Hom(𝒬,IC)=0\operatorname{Hom}(\mathcal{Q}^{\vee},I_{C})=0 for CC of type τ\tau or ρ\rho, and Hom(𝒬,IC)=k\operatorname{Hom}(\mathcal{Q}^{\vee},I_{C})=k for a σ\sigma-conic. Then the result follows. ∎

Now we recall some properties of the Fano surface of conics 𝒞(X)\mathcal{C}(X).

Theorem 6.4 ([39], [13]).

Let XX be an ordinary GM threefold. Then 𝒞(X)\mathcal{C}(X) is an irreducible projective surface. If XX is furthermore general, then 𝒞(X)\mathcal{C}(X) is smooth.

It is a fact that there is a unique ρ\rho-conic on XX, and there is a curve Lσ𝒞(X)L_{\sigma}\subset\mathcal{C}(X) parameterise all σ\sigma-conics on XX (cf. [13, Section 5.1]), and we denote it by cXc_{X}. Furthermore, we have the following result which is a corollary of Logachev’s Tangent Bundle Theorem ([39, Section 4]).

Lemma 6.5 ([13, p. 16]).

The only rational curve in 𝒞(X)\mathcal{C}(X) is LσL_{\sigma}. Furthermore, there exists a surface 𝒞m(X)\mathcal{C}_{m}(X) and a map 𝒞(X)𝒞m(X)\mathcal{C}(X)\rightarrow\mathcal{C}_{m}(X) which contracts LσL_{\sigma} to a point [π][\pi]. If XX is general, then 𝒞m(X)\mathcal{C}_{m}(X) is the minimal surface of 𝒞(X)\mathcal{C}(X).

Theorem 6.6 ([13, Section 5.2]).

Let XX be a general ordinary GM threefold. Then there is a natural involution ι\iota on 𝒞m(X)\mathcal{C}_{m}(X), switching the points [cX][c_{X}] and [π][\pi].

Another important result that we require is Logachev’s Reconstruction Theorem. This was originally proved in [39, Theorem 7.7], and then reproved later in [13, Theorem 9.1].

Theorem 6.7 (Logachev’s Reconstruction Theorem).

Let XX and XX^{\prime} be general ordinary GM threefolds. If 𝒞(X)𝒞(X)\mathcal{C}(X)\cong\mathcal{C}(X^{\prime}), then XXX\cong X^{\prime}.

6.2. Conic and line transforms

For this section, we follow [13, Section 6.1]. Let XX be a general ordinary GM threefold, and let CC be a conic. Then in [13, § 6.1, Theorem 6.4], the authors construct a new GM threefold XCX_{C} and a birational map ψC:XXC\psi_{C}\colon X\dashrightarrow X_{C}, called the conic transform. Similarly, for any line LXL\subset X, a new GM threefold XLX_{L} and a birational morphism ψL:XXL\psi_{L}\colon X\to X_{L} are constructed in [13, Section 6.2], called the line transform.

Note that in [14], such an XCX_{C} is called the period partner of XX, and the line transforms are called the period duals. We now list some important results about conic and line transforms below.

Theorem 6.8 ([13, Theorem 6.4]).

Let XX be a general ordinary GM threefold, and let CXC\subset X be a conic. Then 𝒞(XC)\mathcal{C}(X_{C}) is isomorphic to 𝒞m(X)\mathcal{C}_{m}(X) blown up at the point [C]𝒞m(X)[C]\in\mathcal{C}_{m}(X), where 𝒞m(X)\mathcal{C}_{m}(X) is the minimal surface of 𝒞(X)\mathcal{C}(X).

Proposition 6.9 ([13, Theorem 6.4, Remark 7.2]).

Let XX be a general ordinary GM threefold. Then the isomorphism classes of conic transforms of XX are parametrized by the surface 𝒞m(X)/ι\mathcal{C}_{m}(X)/\iota.

Theorem 6.10 ([30, Theorem 1.6]).

Let XX be a general ordinary GM threefold. Then the Kuznetsov components of all conic transforms and line transforms of XX are equivalent to 𝒜X\mathcal{A}_{X}.

6.3. Conics on special GM threefolds

Let XX be a special GM threefold. Recall that XX is a double cover XYX\to Y of a degree five del Pezzo threefold YY with branch locus a quadric hypersurface Y\mathcal{B}\subset Y. When XX is general, \mathcal{B} is a smooth K3 surface of Picard number 11 and degree 1010. Recall that YY is a codimension 33 linear section of Gr(2,5)\mathrm{Gr}(2,5). Let 𝒱\mathcal{V} be the tautological quotient bundle on YY. We recall some properties of 𝒞(X)\mathcal{C}(X) from [22].

Theorem 6.11 ([22]).

Let XX be a special GM threefold. Then 𝒞(X)\mathcal{C}(X) has two components 𝒞1\mathcal{C}_{1} and 𝒞2\mathcal{C}_{2}. One of the components 𝒞2Σ(Y)2\mathcal{C}_{2}\cong\Sigma(Y)\cong\mathbb{P}^{2} parametrizes the preimage of lines on YY. Moreover, when XX is general, 𝒞(X)\mathcal{C}(X) is smooth away from 𝒞1𝒞2\mathcal{C}_{1}\cap\mathcal{C}_{2}.

The following lemma will be useful in computations; it is similar to Lemma 6.3.

Lemma 6.12.

Let XX be a special GM threefold and CC a conic on XX. Then RHom(,IC)0\operatorname{RHom}^{\bullet}(\mathcal{E}^{\vee},I_{C})\neq 0 if and only if CC is the preimage of a line on YY. In this case RHom(,IC)=k[1]k[2]\operatorname{RHom}^{\bullet}(\mathcal{E}^{\vee},I_{C})=k[-1]\oplus k[-2], and such a family of conics is parametrized by the Hilbert scheme of lines Σ(Y)2\Sigma(Y)\cong\mathbb{P}^{2} on YY.

Proof.

The proof is almost the same as the second part of the proof of Lemma 6.3. The same argument shows that RHom(,IC)0\operatorname{RHom}^{\bullet}(\mathcal{E}^{\vee},I_{C})\neq 0 if and only if Hom(𝒬,IC)0\operatorname{Hom}(\mathcal{Q}^{\vee},I_{C})\neq 0. The image of a non-trivial map 𝒬IC\mathcal{Q}^{\vee}\to I_{C} is the ideal sheaf of the zero locus of a section ss of 𝒬\mathcal{Q}, which is the preimage of the zero locus of a section of 𝒱\mathcal{V}. By [53, Lemma 2.18], the zero locus of a section of 𝒱\mathcal{V} is either a line or a point. Thus the zero locus of a section of 𝒬\mathcal{Q} is either the preimage of a line on YY which is a conic on XX, or a zero-dimensional closed subscheme of length two. But this zero locus contains a conic CXC\subset X, so C=Z(s)C=Z(s) is the preimage of a line on YY and the map 𝒬IC\mathcal{Q}^{\vee}\to I_{C} is surjective. In particular, such conics are exactly the preimages of lines on YY and are parametrized by Σ(Y)2\Sigma(Y)\cong\mathbb{P}^{2}. ∎

7. Conics and Bridgeland moduli spaces

In this section, we study the moduli space of σ\sigma-stable objects of the (1)(-1)-class x-x in the alternative Kuznetsov component 𝒜X\mathcal{A}_{X} of a GM threefold XX and its relation to 𝒞(X)\mathcal{C}(X). Our main result in this section is Theorem 7.12, which realizes the Bridgeland moduli space as a contraction of 𝒞(X)\mathcal{C}(X).

First, we study those conics CC such that IC𝒜XI_{C}\notin\mathcal{A}_{X}.

Proposition 7.1.

Let CXC\subset X be a conic on a GM threefold XX. Then IC𝒜XI_{C}\not\in\mathcal{A}_{X} if and only if

  1. (1)

    CC is a σ\sigma-conic when XX is ordinary. In particular, such a family of conics is parametrized by the line LσL_{\sigma}.

  2. (2)

    CC is the preimage of a line on YY when XX is special. In particular, such a family of conics is parametrized by the Hilbert scheme of lines Σ(Y)2\Sigma(Y)\cong\mathbb{P}^{2} on YY.

Moreover, we have an exact sequence

0𝒬IC0.0\to\mathcal{E}\to\mathcal{Q}^{\vee}\to I_{C}\to 0.
Proof.

Note that IC𝒜XI_{C}\notin\mathcal{A}_{X} if and only if RHom(,IC)0\mathrm{RHom}^{\bullet}(\mathcal{E}^{\vee},I_{C})\neq 0. When XX is ordinary, (1) follows from Lemma 6.3. When XX is special, we deduce (2) from Lemma 6.12. Note that since IC𝒜XI_{C}\notin\mathcal{A}_{X}, we have Hom(𝒬,IC)0\operatorname{Hom}(\mathcal{Q}^{\vee},I_{C})\neq 0. The non-trivial map 𝒬IC\mathcal{Q}^{\vee}\to I_{C} is surjective by the arguments in Lemma 6.3 and 6.12. Note that by the stability of 𝒬\mathcal{Q}^{\vee}, the kernel of 𝒬IC\mathcal{Q}^{\vee}\twoheadrightarrow I_{C} is μ\mu-stable with the same Chern character as \mathcal{E}, hence we have ker(𝒬IC)\ker(\mathcal{Q}^{\vee}\twoheadrightarrow I_{C})\cong\mathcal{E} by [13, Proposition 4.1]. ∎

Proposition 7.2.

Let XX be a GM threefold and CXC\subset X a conic on XX. If IC𝒜XI_{C}\not\in\mathcal{A}_{X}, then we have the exact triangle

[1]pr(IC)𝒬\mathcal{E}[1]\rightarrow\mathrm{pr}(I_{C})\rightarrow\mathcal{Q}^{\vee}

and pr(IC)π(𝒬)\mathrm{pr}(I_{C})\cong\pi^{\prime}(\mathcal{Q}^{\vee})

Proof.

By Proposition 7.1, ICI_{C} fits into the short exact sequence

0𝒬IC0.0\rightarrow\mathcal{E}\rightarrow\mathcal{Q}^{\vee}\rightarrow I_{C}\rightarrow 0.

Applying the projection functor to this exact sequence, and note that applying the functor pr\mathrm{pr} to the dual exact sequence of (1) gives pr(𝒬)=0\mathrm{pr}(\mathcal{Q}^{\vee})=0. Then we have pr(IC)pr()[1]\mathrm{pr}(I_{C})\cong\mathrm{pr}(\mathcal{E})[1]. Now we compute the projection pr()\mathrm{pr}(\mathcal{E}). Since RHom(,)k[3]\mathrm{RHom}^{\bullet}(\mathcal{E}^{\vee},\mathcal{E})\cong k[-3], we get an exact triangle [3]𝐋\mathcal{E}^{\vee}[-3]\rightarrow\mathcal{E}\rightarrow\bm{\mathrm{L}}_{\mathcal{E}^{\vee}}\mathcal{E}. Now applying 𝐋𝒪X\bm{\mathrm{L}}_{\mathcal{O}_{X}} to this triangle and using 𝐋𝒪X=𝒬[1]\bm{\mathrm{L}}_{\operatorname{\mathcal{O}}_{X}}\mathcal{E}^{\vee}=\mathcal{Q}^{\vee}[1], we get

𝒬[2]pr().\mathcal{Q}^{\vee}[-2]\rightarrow\mathcal{E}\rightarrow\mathrm{pr}(\mathcal{E}).

Therefore we obtain the triangle

[1]pr()[1]𝒬\mathcal{E}[1]\rightarrow\mathrm{pr}(\mathcal{E})[1]\rightarrow\mathcal{Q}^{\vee}

and the desired result follows from Lemma 5.4. ∎

Now the following two results follow from Proposition 7.2 and Lemma 5.6.

Lemma 7.3.

Let XX be a GM threefold. If CXC\subset X is a conic such that IC𝒜XI_{C}\not\in\mathcal{A}_{X}, then

  • RHom(pr(IC),pr(IC))=kk2[1]\operatorname{RHom}^{\bullet}(\mathrm{pr}(I_{C}),\mathrm{pr}(I_{C}))=k\oplus k^{2}[-1] when XX is ordinary.

  • RHom(pr(IC),pr(IC))=kk3[1]k[2]\operatorname{RHom}^{\bullet}(\mathrm{pr}(I_{C}),\mathrm{pr}(I_{C}))=k\oplus k^{3}[-1]\oplus k[-2] when XX is special.

Lemma 7.4.

Let XX be a GM threefold. If IC𝒜XI_{C}\not\in\mathcal{A}_{X}, the projection pr(IC)[1]\mathrm{pr}(I_{C})[1] is stable with respect to every Serre-invariant stability condition on 𝒜X\mathcal{A}_{X}.

When IC𝒜XI_{C}\in\mathcal{A}_{X}, we cannot use Proposition 4.12 to prove the Bridgeland stability of ICI_{C}, since 𝒞(X)\mathcal{C}(X) can be singular and Ext1(IC,IC)\operatorname{Ext}^{1}(I_{C},I_{C}) may have large dimension. Instead, we use a wall-crossing argument and the uniqueness of Serre-invariant stability conditions (Theorem A.10).

Lemma 7.5.

Let XX be a GM threefold. Let FF be an object with ch2(F)=(1,0,2L)\mathrm{ch}_{\leq 2}(F)=(1,0,-2L). Then there are no walls for FF in the range 12β<0-\frac{1}{2}\leq\beta<0 and α>0\alpha>0.

Proof.

Recall that by [3, Theorem 4.13], β=0\beta=0 is the unique vertical wall of FF. Any other wall is a semicircle centered along the β\beta-axis, and its apex lies on the hyperbola μα,β(F)=0\mu_{\alpha,\beta}(F)=0. Moreover, no two walls intersect.

Note that when μα,β(F)=0\mu_{\alpha,\beta}(F)=0 holds, we have β<25<12\beta<-\sqrt{\frac{2}{5}}<-\frac{1}{2}, thus we know that there is no semicircular wall centered in the interval 12β<0-\frac{1}{2}\leq\beta<0. Therefore, any semicircular wall in the range 12β<0-\frac{1}{2}\leq\beta<0 will intersect β=12\beta=-\frac{1}{2}. To prove the statement, we only need to show that there are no walls when β=12\beta=-\frac{1}{2}. This follows from the fact that ch112(F)\mathrm{ch}^{-\frac{1}{2}}_{1}(F) is minimal. ∎

Lemma 7.6.

Let CXC\subset X be a conic on a GM threefold XX such that IC𝒜XI_{C}\in\mathcal{A}_{X}. Then IC[1]I_{C}[1] is stable with respect to every Serre-invariant stability condition on 𝒜X\mathcal{A}_{X}.

Proof.

By Lemma 4.5 and Lemma 7.5, we know that ICI_{C} is σα,β\sigma_{\alpha,\beta}-semistable for every (α,β)V(\alpha,\beta)\in V. Since ICI_{C} is torsion-free, IC[1]Cohα,β0(X)I_{C}[1]\in\operatorname{\mathrm{Coh}}^{0}_{\alpha,\beta}(X) is σα,β0\sigma^{0}_{\alpha,\beta}-semistable. Thus IC[1]𝒜(α,β)I_{C}[1]\in\mathcal{A}(\alpha,\beta) is σ(α,β)\sigma(\alpha,\beta)-semistable. Then stability with respect to every Serre-invariant stability condition follows from Theorem 4.11 and Theorem A.10. ∎

7.1. The Bridgeland moduli space of class x-x

In this subsection, we are going to describe the Bridgeland moduli space σ(𝒜X,x)\mathcal{M}_{\sigma}(\mathcal{A}_{X},-x) in Theorem 7.12.

The proofs in this section seem technical. However, the only results in this section that will be used in other sections are Proposition 7.11 in the proof of Theorem 7.12, so there is no harm for readers in skipping this whole section and assuming Proposition 7.11 and Theorem 7.12.

We start with two lemmas.

Lemma 7.7.

Let XX be a GM threefold and EE a μ\mu-semistable sheaf on XX with truncated Chern character ch2(E)=(2,H,aL)\mathrm{ch}_{\leq 2}(E)=(2,-H,aL). If a1a\geq 1 and c3(E)0c_{3}(E)\geq 0 and then we have EE\cong\mathcal{E}.

Proof.

By Lemma 4.6, we have a1a\leq 1 which means a=1a=1 by our assumption. Then c1(E)=1c_{1}(E)=-1 and c2(E)=4c_{2}(E)=4. Since c3(E)0c_{3}(E)\geq 0, by [4, Proposition 3.5] we have χ(E)=0\chi(E)=0. This implies c3(E)=0c_{3}(E)=0. Moreover, EE^{\vee\vee} also satisfies the assumptions above. Hence by the previous argument, we have c1(E)=1,c2(E)=4c_{1}(E^{\vee\vee})=-1,c_{2}(E^{\vee\vee})=4 and c3(E)=0c_{3}(E^{\vee\vee})=0 as well. In other words, E=EE=E^{\vee\vee}. Since c3(E)=0c_{3}(E)=0 and EE is reflexive of rank two, it is a vector bundle. Moreover, EE is a globally generated bundle by [4, Proposition 3.5]. Thus EE\cong\mathcal{E} by [13, Proposition 4.1]. ∎

Lemma 7.8.

Let XX be a GM threefold and EE a μ\mu-semistable sheaf on XX with ch(E)=ch(𝒬)\mathrm{ch}(E)=\mathrm{ch}(\mathcal{Q}). Then we have E𝒬E\cong\mathcal{Q}.

Proof.

First we show that h2(E)=0h^{2}(E)=0; then from χ(E)=5\chi(E)=5 we have h0(E)5h^{0}(E)\geq 5. Indeed, if h2(E)0h^{2}(E)\neq 0, then Hom(E,𝒪X(H)[1])0\operatorname{Hom}(E,\operatorname{\mathcal{O}}_{X}(-H)[1])\neq 0 by Serre duality. Therefore, we have a non-trivial extension

0𝒪X(H)FE0.0\to\operatorname{\mathcal{O}}_{X}(-H)\to F\to E\to 0.

If FF is not μ\mu-semistable, then by the stability of 𝒪X(H)\operatorname{\mathcal{O}}_{X}(-H) and EE, the minimal destabilizing quotient sheaf FF^{\prime} of FF has ch1(F)=(1,H)\mathrm{ch}_{\leq 1}(F^{\prime})=(1,-H). Thus F𝒪X(H)F^{\prime\vee\vee}\cong\operatorname{\mathcal{O}}_{X}(-H). But if we apply Hom(,𝒪X(H))\operatorname{Hom}(-,\operatorname{\mathcal{O}}_{X}(-H)) to the exact sequence above, we obtain Hom(F,𝒪X(H))=0\operatorname{Hom}(F,\operatorname{\mathcal{O}}_{X}(-H))=0 since this extension is non-trivial, which gives a contradiction. Then FF is μ\mu-semistable with ch2(F)=(4,0,4L)\mathrm{ch}_{\leq 2}(F)=(4,0,4L), which is impossible since Δ(F)<0\Delta(F)<0.

Now we can take five linearly independent elements in H0(E)H^{0}(E) and obtain a map t:𝒪X5Et\colon\operatorname{\mathcal{O}}^{\oplus 5}_{X}\to E. From the stability of 𝒪X\operatorname{\mathcal{O}}_{X} and EE, we have μ(Im(t))=0\mu(\operatorname{Im}(t))=0 or μ(Im(t))=13\mu(\operatorname{Im}(t))=\frac{1}{3}. But the first case cannot happen, since then Im(t)\operatorname{Im}(t) is the direct sum of a number of copies of 𝒪X\operatorname{\mathcal{O}}_{X}, and this contradicts the construction of tt. Thus μ(Im(t))=13\mu(\operatorname{Im}(t))=\frac{1}{3} and ch1(Im(t))=(3,H)\mathrm{ch}_{\leq 1}(\operatorname{Im}(t))=(3,H). Also ch2(ker(t))=(2,H,xL)\mathrm{ch}_{\leq 2}(\ker(t))=(2,-H,xL), where x1x\geq 1. Note that ker(t)\ker(t) is reflexive, thus we have c3(ker(t))0c_{3}(\ker(t))\geq 0 since ker(t)\ker(t) has rank two. Then by stability of 𝒪X\operatorname{\mathcal{O}}_{X} and Hom(𝒪X,ker(t))=0\operatorname{Hom}(\operatorname{\mathcal{O}}_{X},\ker(t))=0, it is not hard to see that ker(t)\ker(t) is μ\mu-semistable. Thus by Lemma 7.7 we have ker(t)\ker(t)\cong\mathcal{E}. Therefore ch(Im(t))=ch(E)\mathrm{ch}(\operatorname{Im}(t))=\mathrm{ch}(E) and thus tt is surjective.

Now applying Hom(𝒬,)\operatorname{Hom}(\mathcal{Q},-) to the exact sequence

0𝒪X5E0,0\to\mathcal{E}\to\operatorname{\mathcal{O}}^{\oplus 5}_{X}\to E\to 0,

from RHom(𝒬,𝒪X)=0\operatorname{RHom}^{\bullet}(\mathcal{Q},\operatorname{\mathcal{O}}_{X})=0 and Ext1(𝒬,)=k\operatorname{Ext}^{1}(\mathcal{Q},\mathcal{E})=k we have Hom(𝒬,E)=k\operatorname{Hom}(\mathcal{Q},E)=k. Thus from the stability of EE and 𝒬\mathcal{Q}, we have E𝒬E\cong\mathcal{Q} and the result follows. ∎

Now we introduce some notations. Let α>0\alpha>0 and β<0\beta<0. For an object EDb(X)E\in\mathrm{D}^{b}(X), the limit central charge Z0,00(E)Z^{0}_{0,0}(E) is defined as the limit of Zα,β0(E)Z^{0}_{\alpha,\beta}(E) when (α,β)(0,0)(\alpha,\beta)\to(0,0). Note that Zα,β0(E)Z^{0}_{\alpha,\beta}(E) is given by \mathbb{Q}-linear combinations of α,β,α2,β2\alpha,\beta,\alpha^{2},\beta^{2}, thus such a limit Z0,00(E)Z^{0}_{0,0}(E) always exists. For Z0,00(E)0Z^{0}_{0,0}(E)\neq 0, we can also define the limit slope μ0,00(E)\mu^{0}_{0,0}(E) as follows:

  • If Im(Z0,00(E))0\operatorname{Im}(Z^{0}_{0,0}(E))\neq 0, then we define μ0,00(E):=Re(Z0,00(E))Im(Z0,00(E))\mu^{0}_{0,0}(E):=-\frac{\operatorname{Re}(Z^{0}_{0,0}(E))}{\operatorname{Im}(Z^{0}_{0,0}(E))}.

  • If Im(Z0,00(E))=0\operatorname{Im}(Z^{0}_{0,0}(E))=0 and Re(Z0,00(E))>0\operatorname{Re}(Z^{0}_{0,0}(E))>0, then we define μ0,00(E):=\mu^{0}_{0,0}(E):=-\infty.

  • If Im(Z0,00(E))=0\operatorname{Im}(Z^{0}_{0,0}(E))=0 and Re(Z0,00(E))<0\operatorname{Re}(Z^{0}_{0,0}(E))<0, then we define μ0,00(E):=+\mu^{0}_{0,0}(E):=+\infty.

Note that Z0,00(E)=0Z^{0}_{0,0}(E)=0 if and only if ch2(E)\mathrm{ch}_{\leq 2}(E) is a multiple of ch2(𝒪X)\mathrm{ch}_{\leq 2}(\operatorname{\mathcal{O}}_{X}).

Let ECohα,β0(X)E\in\operatorname{\mathrm{Coh}}^{0}_{\alpha,\beta}(X). By continuity, we can find a neighborhood UEU_{E} of the origin such that for any (α,β)UE(\alpha,\beta)\in U_{E}, the slopes μα,β0(E)\mu^{0}_{\alpha,\beta}(E) and μ0,00(E)\mu^{0}_{0,0}(E) are both negative or positive. Let FCohα,β0(X)F\in\operatorname{\mathrm{Coh}}^{0}_{\alpha,\beta}(X) be another object such that E,FE,F are both σα,β0\sigma^{0}_{\alpha,\beta}-semistable in a neighborhood UE,FU_{E,F} of the origin. If μ0,00(E)>μ0,00(F)\mu^{0}_{0,0}(E)>\mu^{0}_{0,0}(F), then by continuity, we can find a smaller neighborhood UE,FU^{\prime}_{E,F} such that μα,β0(E)>μα,β0(F)\mu^{0}_{\alpha,\beta}(E)>\mu^{0}_{\alpha,\beta}(F) holds for every (α,β)UE,F(\alpha,\beta)\in U^{\prime}_{E,F}. Thus we have Hom(E,F)=0\operatorname{Hom}(E,F)=0. We will use these two elementary facts repeatedly.

Proposition 7.9.

If F𝒜(α,β)F\in\mathcal{A}(\alpha,\beta) is σ(α,β)\sigma(\alpha,\beta)-stable such that [F]=x[F]=-x and FF is σα,β0\sigma^{0}_{\alpha,\beta}-semistable for some (α,β)V(\alpha,\beta)\in V, then FIC[1]F\cong I_{C}[1] for a conic CC on XX.

Proof.

Since FF is σα,β0\sigma^{0}_{\alpha,\beta}-semistable and μα,β0(F)>0\mu^{0}_{\alpha,\beta}(F)>0, as in [52, Proposition 4.6] there is a triangle

F1[1]FF2F_{1}[1]\to F\to F_{2}

where F1Cohβ(X)F_{1}\in\operatorname{\mathrm{Coh}}^{\beta}(X) with μα,β+(F1)<0\mu^{+}_{\alpha,\beta}(F_{1})<0 and F2F_{2} is supported on points. Thus ch(F1)=(1,0,2L,mP)\mathrm{ch}(F_{1})=(1,0,-2L,mP), where mm is the length of F2F_{2}. By Lemmas 7.5 and 4.5, F1F_{1} is a rank one torsion-free sheaf, hence it is the ideal sheaf of a closed subscheme. Thus by [53, Corollary 1.38], we have m0m\leq 0, which means F2=0F_{2}=0 and F1F[1]F_{1}\cong F[-1]. Thus by Lemma 7.5 again, F[1]F[-1] is a μ\mu-semistable torsion free sheaf, which is of the form F[1]ICF[-1]\cong I_{C} for a conic CC on XX since Pic(X)=H\operatorname{Pic}(X)=\mathbb{Z}\cdot H. ∎

When FF is not σα,β0\sigma^{0}_{\alpha,\beta}-semistable for (α,β)V(\alpha,\beta)\in V, the argument is more complicated. Our main tools are the inequalities in [49], [52, Proposition 4.1], Lemma 4.6 and Theorem 4.7, which allow us to bound the rank and first two Chern characters ch1,ch2\mathrm{ch}_{1},\mathrm{ch}_{2} of the destabilizing objects and their cohomology objects. Since F𝒜XF\in\mathcal{A}_{X}, by using the Euler characteristics χ(𝒪X,)\chi(\mathcal{O}_{X},-) and χ(,)\chi(\mathcal{E}^{\vee},-) we can obtain a bound on ch3\mathrm{ch}_{3}. Finally, via a similar argument as in Lemma 7.7 we deduce that the Harder–Narasimhan factors of FF are the ones we expect.

Proposition 7.10.

If F𝒜(α,β)F\in\mathcal{A}(\alpha,\beta) is σ(α,β)\sigma(\alpha,\beta)-stable such that [F]=x[F]=-x and FF is not σα,β0\sigma^{0}_{\alpha,\beta}-semistable for every (α,β)V(\alpha,\beta)\in V, then FF fits into a triangle

[2]F𝒬[1].\mathcal{E}[2]\to F\to\mathcal{Q}^{\vee}[1].
Proof.

Since there are no walls for FF tangent to the wall β=0\beta=0, by the local finiteness of walls and [9, Proposition 2.2.2] we can find an open neighborhood UU^{\prime} of the origin such that the Harder–Narasimhan filtration with respect to σα,β0\sigma^{0}_{\alpha,\beta} is constant for every (α,β)U:=UV(\alpha,\beta)\in U:=U^{\prime}\cap V. In the following we will only consider σα,β0\sigma^{0}_{\alpha,\beta} for (α,β)U(\alpha,\beta)\in U.

Let BB be the minimal destabilizing quotient object of FF and 0AFB00\to A\to F\to B\to 0 be the destabilizing short exact sequence of FF in Cohα,β0(X)\operatorname{\mathrm{Coh}}^{0}_{\alpha,\beta}(X). Hence we know that A,BCohα,β0(X)A,B\in\mathrm{Coh}^{0}_{\alpha,\beta}(X) and BB is σα,β0\sigma^{0}_{\alpha,\beta}-semistable with μα,β0,(A)>μα,β0(F)>μα,β0(B)\mu^{0,-}_{\alpha,\beta}(A)>\mu^{0}_{\alpha,\beta}(F)>\mu^{0}_{\alpha,\beta}(B) for all (α,β)U(\alpha,\beta)\in U. By [5, Remark 5.12], we have μα,β0(B)\mu^{0}_{\alpha,\beta}(B) min{μα,β0(F),μα,β0(𝒪X),μα,β0()}\geq\min\{\mu^{0}_{\alpha,\beta}(F),\mu^{0}_{\alpha,\beta}(\mathcal{O}_{X}),\mu^{0}_{\alpha,\beta}(\mathcal{E}^{\vee})\}. Hence the following relations hold for all (α,β)U(\alpha,\beta)\in U:

  1. (a)

    μα,β0(A)>μα,β0(F)>μα,β0(B)\mu^{0}_{\alpha,\beta}(A)>\mu^{0}_{\alpha,\beta}(F)>\mu^{0}_{\alpha,\beta}(B),

  2. (b)

    Im(Zα,β0(A))0\mathrm{Im}(Z^{0}_{\alpha,\beta}(A))\geq 0, Im(Zα,β0(B))>0\mathrm{Im}(Z^{0}_{\alpha,\beta}(B))>0,

  3. (c)

    μα,β0(B)min{μα,β0(F),μα,β0(𝒪X),μα,β0()}\mu^{0}_{\alpha,\beta}(B)\geq\min\{\mu^{0}_{\alpha,\beta}(F),\mu^{0}_{\alpha,\beta}(\mathcal{O}_{X}),\mu^{0}_{\alpha,\beta}(\mathcal{E}^{\vee})\},

  4. (d)

    Δ(B)0\Delta(B)\geq 0.

By continuity and taking (α,β)(0,0)(\alpha,\beta)\to(0,0), we have:

  1. (1)

    μ0,00(A)μ0,00(F)=0μ0,00(B)\mu^{0}_{0,0}(A)\geq\mu^{0}_{0,0}(F)=0\geq\mu^{0}_{0,0}(B),

  2. (2)

    Im(Z0,00(A))0\mathrm{Im}(Z^{0}_{0,0}(A))\geq 0, Im(Z0,00(B))0\mathrm{Im}(Z^{0}_{0,0}(B))\geq 0,

  3. (3)

    μ0,00(B)min{μ0,00(F),μ0,00(𝒪X),μ0,00()}\mu^{0}_{0,0}(B)\geq\min\{\mu^{0}_{0,0}(F),\mu^{0}_{0,0}(\mathcal{O}_{X}),\mu^{0}_{0,0}(\mathcal{E}^{\vee})\},

  4. (4)

    Δ(B)0\Delta(B)\geq 0.

Assume that [A]=a[𝒪X]+b[𝒪H]+c[𝒪L]+d[𝒪P][A]=a[\mathcal{O}_{X}]+b[\mathcal{O}_{H}]+c[\mathcal{O}_{L}]+d[\mathcal{O}_{P}] for integers a,b,c,da,b,c,d\in\mathbb{Z}. Then we have [B]=(1a)[𝒪X]b[𝒪H]+(2c)[𝒪L](1+d)[𝒪P][B]=(-1-a)[\mathcal{O}_{X}]-b[\mathcal{O}_{H}]+(2-c)[\mathcal{O}_{L}]-(1+d)[\mathcal{O}_{P}]. Then we see

  • ch(A)=(a,bH,c5b10H2,53b+c2+d10H3)\mathrm{ch}(A)=(a,bH,\frac{c-5b}{10}H^{2},\frac{\frac{5}{3}b+\frac{c}{2}+d}{10}H^{3})

  • Z0,00(A)=bH3+(c5b10H3)𝔦Z^{0}_{0,0}(A)=bH^{3}+(\frac{c-5b}{10}H^{3})\mathfrak{i}, Z0,00(B)=bH3+(2c+5b10H3)𝔦Z^{0}_{0,0}(B)=-bH^{3}+(\frac{2-c+5b}{10}H^{3})\mathfrak{i}

  • μ0,00(A)=10b5bc,μ0,00(B)=10bc5b2\mu^{0}_{0,0}(A)=\frac{10b}{5b-c},\mu^{0}_{0,0}(B)=\frac{-10b}{c-5b-2}.

Note that [F]=[𝒪X]+2[𝒪L][𝒪P][F]=-[\mathcal{O}_{X}]+2[\mathcal{O}_{L}]-[\mathcal{O}_{P}]. From (2)(2) we know c5b=0c-5b=0, 11 or 22. But when c5b=2c-5b=2, it is not hard to see that (c)(c) fails near the origin. Thus c5b=0c-5b=0 or 11.

We begin with two claims.

Claim 1: We have RHom(𝒪X,B)=Hom(𝒪X,B)\operatorname{RHom}^{\bullet}(\operatorname{\mathcal{O}}_{X},B)=\operatorname{Hom}(\operatorname{\mathcal{O}}_{X},B) and RHom(𝒪X,A)=Ext1(𝒪X,A)[1]\operatorname{RHom}^{\bullet}(\operatorname{\mathcal{O}}_{X},A)=\operatorname{Ext}^{1}(\operatorname{\mathcal{O}}_{X},A)[-1].

Since F𝒜XF\in\mathcal{A}_{X}, we only need to prove that Exti(𝒪X,A)=0\operatorname{Ext}^{i}(\operatorname{\mathcal{O}}_{X},A)=0 for i1i\neq 1. Indeed, since 𝒪XCohα,β0(X)\operatorname{\mathcal{O}}_{X}\in\operatorname{\mathrm{Coh}}^{0}_{\alpha,\beta}(X) and F𝒜XF\in\mathcal{A}_{X}, we have Exti(𝒪X,A)=0\operatorname{Ext}^{i}(\operatorname{\mathcal{O}}_{X},A)=0 for all i0i\leq 0. Also, by Serre duality we have Exti(𝒪X,A)=Hom(A,𝒪X(H)[3i])\operatorname{Ext}^{i}(\operatorname{\mathcal{O}}_{X},A)=\operatorname{Hom}(A,\operatorname{\mathcal{O}}_{X}(-H)[3-i]). Thus from the fact that 𝒪X(H)Cohα,β0(X)\operatorname{\mathcal{O}}_{X}(-H)\in\operatorname{\mathrm{Coh}}^{0}_{\alpha,\beta}(X), we obtain Hom(A,𝒪X(H)[3i])=0\operatorname{Hom}(A,\operatorname{\mathcal{O}}_{X}(-H)[3-i])=0 for i2i\geq 2. Therefore we have Exti(𝒪X,A)=0\operatorname{Ext}^{i}(\operatorname{\mathcal{O}}_{X},A)=0 for i1i\neq 1.

Claim 2: We have RHom(,B)=Hom(,B)\operatorname{RHom}^{\bullet}(\mathcal{E}^{\vee},B)=\operatorname{Hom}(\mathcal{E}^{\vee},B) and RHom(,A)=Ext1(,A)[1]\operatorname{RHom}^{\bullet}(\mathcal{E}^{\vee},A)=\operatorname{Ext}^{1}(\mathcal{E}^{\vee},A)[-1].

Since \mathcal{E}^{\vee} and [2]Cohα,β0(X)\mathcal{E}[2]\in\operatorname{\mathrm{Coh}}^{0}_{\alpha,\beta}(X), the argument is the same as Claim 1.

Now we deal with the cases c5b=0c-5b=0 and c5b=1c-5b=1 separately.

Case 1 (c5b=0c-5b=0):

First, we assume that c5b=0c-5b=0. By 7.1, we have:

  1. (1)

    2b0-2\leq b\leq 0,

  2. (2)

    b2+2a+250b^{2}+\frac{2a+2}{5}\geq 0.

case 1.1 (b=0b=0): If b=0b=0, then c=0c=0 and a1a\geq-1. In this case we have ch2(B)=(1a,0,2L)\mathrm{ch}_{\leq 2}(B)=(-1-a,0,2L). If a=1a=-1, then ch2(A)=ch2(𝒪X[1])=(1,0,0)\mathrm{ch}_{\leq 2}(A)=\mathrm{ch}_{\leq 2}(\operatorname{\mathcal{O}}_{X}[1])=(-1,0,0), which is impossible since Im(Zα,β0(A))<0\operatorname{Im}(Z^{0}_{\alpha,\beta}(A))<0 for (α,β)V(\alpha,\beta)\in V. Thus a0a\geq 0, and a0a\neq 0 otherwise μα,β0(F)=μα,β0(B)\mu^{0}_{\alpha,\beta}(F)=\mu^{0}_{\alpha,\beta}(B) for any (α,β)V(\alpha,\beta)\in V. But then we have μα,β0(F)<μα,β0(B)\mu^{0}_{\alpha,\beta}(F)<\mu^{0}_{\alpha,\beta}(B) when (α,β)U(\alpha,\beta)\in U is sufficiently close to the origin. This contradicts our assumption on BB.

case 1.2 (b=1b=-1): If b=1b=-1, we have c=5c=-5. In this case ch2(A)=(a,H,0)\mathrm{ch}_{\leq 2}(A)=(a,-H,0). Since ACohα,β0(X)A\in\operatorname{\mathrm{Coh}}^{0}_{\alpha,\beta}(X), we have Im(Zα,β0(A))0\mathrm{Im}(Z^{0}_{\alpha,\beta}(A))\geq 0 for every (α,β)U(\alpha,\beta)\in U. Note that Im(Zα,β0(A))=(β+a(β2α2)2)H3\mathrm{Im}(Z^{0}_{\alpha,\beta}(A))=(\beta+\frac{a(\beta^{2}-\alpha^{2})}{2})H^{3} and 0<α<β0<\alpha<-\beta, and we have a2ββ2α2a\geq\frac{-2\beta}{\beta^{2}-\alpha^{2}}. But note that when α=β2\alpha=\frac{-\beta}{2} and β0\beta\to 0, we have 2ββ2α2+\frac{-2\beta}{\beta^{2}-\alpha^{2}}\to+\infty, thus we get a contradiction since aa is a finite number.

case 1.3 (b=2b=-2): If b=2b=-2, we have c=10c=-10. In this case we have ch2(A)=(a,2H,0)\mathrm{ch}_{\leq 2}(A)=(a,-2H,0). Similarly to case 1.2, we have Im(Zα,β0(A))0\mathrm{Im}(Z^{0}_{\alpha,\beta}(A))\geq 0 for every (α,β)U(\alpha,\beta)\in U. Note that Im(Zα,β0(A))=(2β+a(β2α2)2)H3\mathrm{Im}(Z^{0}_{\alpha,\beta}(A))=(2\beta+\frac{a(\beta^{2}-\alpha^{2})}{2})H^{3} and α<β\alpha<-\beta, and we have a4ββ2α2a\geq\frac{-4\beta}{\beta^{2}-\alpha^{2}}. Then as in case 1.2, we get a contradiction.

Case 2 (c5b=1c-5b=1): Now we assume that c5b=1c-5b=1. Then by 7.1, we have:

  1. (1)

    1b0-1\leq b\leq 0,

  2. (2)

    b2+a+150b^{2}+\frac{a+1}{5}\geq 0.

case 2.1 (b=0b=0): If b=0b=0, then c=1c=1. Therefore 1a-1\leq a. If a=1a=-1, since BB is σα,β0\sigma^{0}_{\alpha,\beta}-semistable, we know Cohβ(X)0(B)\mathcal{H}^{0}_{\mathrm{Coh}^{\beta}(X)}(B) is either 0 or supported on points. Thus ch2(Cohβ(X)1(B))=(0,0,L)\mathrm{ch}_{\leq 2}(\mathcal{H}^{-1}_{\mathrm{Coh}^{\beta}(X)}(B))=(0,0,-L). But Re(Zα,β(Cohβ(X)1(B)))>0\mathrm{Re}(Z_{\alpha,\beta}(\mathcal{H}^{-1}_{\mathrm{Coh}^{\beta}(X)}(B)))>0 which is impossible since Cohβ(X)1(B)Cohβ(X)\mathcal{H}^{-1}_{\mathrm{Coh}^{\beta}(X)}(B)\in\mathrm{Coh}^{\beta}(X) with Im(Zα,β(Cohβ(X)1(B)))=0\mathrm{Im}(Z_{\alpha,\beta}(\mathcal{H}^{-1}_{\mathrm{Coh}^{\beta}(X)}(B)))=0.

Therefore we have a0a\geq 0. Hence ch2(B)=(a+1,0,L)\mathrm{ch}_{\leq 2}(B)=-(a+1,0,-L), where a+11a+1\geq 1. This is also impossible since when (α,β)U(\alpha,\beta)\in U is sufficiently close to the origin, we have μα,β0(B)>μα,β0(F)\mu^{0}_{\alpha,\beta}(B)>\mu^{0}_{\alpha,\beta}(F).

case 2.2 (b=1b=-1): We have b=1b=-1 and c=4c=-4. Hence 6a-6\leq a. In this case ch2(B)=(1a,H,L)\mathrm{ch}_{\leq 2}(B)=(-1-a,H,L) and we have μα,β0(B)<0\mu^{0}_{\alpha,\beta}(B)<0 and ch1β(B)>0\mathrm{ch}_{1}^{\beta}(B)>0 for when (α,β)U(\alpha,\beta)\in U is sufficiently close to the origin. Thus BCohβ(X)B\in\operatorname{\mathrm{Coh}}^{\beta}(X) is σα,β\sigma_{\alpha,\beta}-semistable. Applying Lemma 4.6 to BB, we have a3a\geq-3.

We first prove a claim.

Claim 3: In the situation of case 2.2, we have AA is σα,β0\sigma^{0}_{\alpha,\beta}-semistable. Hence RHom(𝒪X,A)=0\operatorname{RHom}^{\bullet}(\operatorname{\mathcal{O}}_{X},A)=0, ch(A)=(a,H,L,(73a)P)\mathrm{ch}(A)=(a,-H,L,(\frac{7}{3}-a)P) and χ(,A)=32a\chi(\mathcal{E}^{\vee},A)=3-2a.

Assume AA is not σα,β0\sigma^{0}_{\alpha,\beta}-semistable for some (α,β)U(\alpha,\beta)\in U. Then we can take a neighborhood UAU^{\prime}_{A} of the origin such that AA has constant Harder–Narasimhan factors for any (α,β)UA:=UUAV(\alpha,\beta)\in U_{A}:=U\cap U^{\prime}_{A}\cap V. Let CC be the minimal destabilizing quotient object of AA with respect to σα,β0\sigma^{0}_{\alpha,\beta} for (α,β)UA(\alpha,\beta)\in U_{A}. In this case we have ch2(A)=(a,H,L)\mathrm{ch}_{\leq 2}(A)=(a,-H,L). Since Im(Z0,00(A))=110H3\mathrm{Im}(Z^{0}_{0,0}(A))=\frac{1}{10}H^{3}, we know that Im(Z0,00(C))=0\mathrm{Im}(Z^{0}_{0,0}(C))=0 or 110H3\frac{1}{10}H^{3}. If Im(Z0,00(C))=0\mathrm{Im}(Z^{0}_{0,0}(C))=0, then μ0,00(C)=+\mu^{0}_{0,0}(C)=+\infty or -\infty. But the previous case contradicts μα,β0(A)>μα,β0(C)\mu^{0}_{\alpha,\beta}(A)>\mu^{0}_{\alpha,\beta}(C) and the latter case contradicts μα,β0(C)>μα,β0(F)\mu^{0}_{\alpha,\beta}(C)>\mu^{0}_{\alpha,\beta}(F). Therefore we have Im(Z0,00(C))=110H3\mathrm{Im}(Z^{0}_{0,0}(C))=\frac{1}{10}H^{3} and we can assume that ch2(C)=(e,fH,L)\mathrm{ch}_{\leq 2}(C)=(e,fH,L) where e,fe,f\in\mathbb{Z}. Since μ0,00(A)μ0,00(C)μ0,00(F)=0\mu^{0}_{0,0}(A)\geq\mu^{0}_{0,0}(C)\geq\mu^{0}_{0,0}(F)=0, we have 1010f010\geq-10f\geq 0. If f=0f=0, then ch2(C)=(e,0,L)\mathrm{ch}_{\leq 2}(C)=(e,0,L) and ch2(D)=(ae,H,0)\mathrm{ch}_{\leq 2}(D)=(a-e,-H,0), where D=cone(AC)[1]D=\mathrm{cone}(A\to C)[-1]. Then μα,β0(D)>μα,β0(A)\mu^{0-}_{\alpha,\beta}(D)>\mu^{0}_{\alpha,\beta}(A) for any (α,β)UA(\alpha,\beta)\in U_{A}. Hence

μα,β0(D)=1+(ae)ββ+ae2(β2α2)>μα,β0(A)>μα,β0(F).\mu^{0}_{\alpha,\beta}(D)=\frac{1+(a-e)\beta}{\beta+\frac{a-e}{2}(\beta^{2}-\alpha^{2})}>\mu^{0}_{\alpha,\beta}(A)>\mu^{0}_{\alpha,\beta}(F).

But note that μα,β0(D)<0\mu^{0}_{\alpha,\beta}(D)<0 for (α,β)UA(\alpha,\beta)\in U_{A} that sufficiently closed to the origin, which gives a contradiction since μα,β0(D)>μα0,β00(F)\mu^{0}_{\alpha,\beta}(D)>\mu^{0}_{\alpha_{0},\beta_{0}}(F) holds for any (α,β)UA(\alpha,\beta)\in U_{A}.

Therefore the only possible case is f=1f=-1, and hence μ0,00(C)=10\mu^{0}_{0,0}(C)=10. Since μα,β0(A)>μα,β0(C)\mu^{0}_{\alpha,\beta}(A)>\mu^{0}_{\alpha,\beta}(C) for (α,β)UA(\alpha,\beta)\in U_{A}, we have rkC>a\operatorname{rk}C>a. But this is impossible since D,𝒪XCohα,β0(X)D,\operatorname{\mathcal{O}}_{X}\in\operatorname{\mathrm{Coh}}^{0}_{\alpha,\beta}(X) but ch2(D)=(s,0,0)=sch2(𝒪X)\mathrm{ch}_{\leq 2}(D)=(s,0,0)=s\cdot\mathrm{ch}_{\leq 2}(\operatorname{\mathcal{O}}_{X}) where s=arkC<0s=a-\operatorname{rk}C<0. Now for the last statement, note that 𝒪X(H)[2]Cohα,β0(X)\operatorname{\mathcal{O}}_{X}(-H)[2]\in\operatorname{\mathrm{Coh}}^{0}_{\alpha,\beta}(X) is σα,β0\sigma^{0}_{\alpha,\beta}-semistable with μ0,00(𝒪X(H)[2])=2\mu^{0}_{0,0}(\operatorname{\mathcal{O}}_{X}(-H)[2])=2, hence we have Hom(A,𝒪X(H)[2])=Hom(𝒪X,A[1])=0\operatorname{Hom}(A,\operatorname{\mathcal{O}}_{X}(-H)[2])=\operatorname{Hom}(\operatorname{\mathcal{O}}_{X},A[1])=0. Now combined with Claim 1, this proves our claim.

Now we deal with the three cases a=3a=-3, 2a1-2\leq a\leq 1 and a2a\geq 2 separately.

When a=3a=-3, we have ch2(B)=ch2()\mathrm{ch}_{\leq 2}(B)=\mathrm{ch}_{\leq 2}(\mathcal{E}^{\vee}). Then since ch2(B)\mathrm{ch}_{\leq 2}(B) is on the boundary of Lemma 4.6, by a standard argument we know that BB is σα,β\sigma_{\alpha,\beta}-semistable for every α>0\alpha>0 and β<0\beta<0, as explained in [48, Proposition 3.2]. Thus by Lemma 4.5, BB is a μ\mu-semistable sheaf. From Claim 3 we have χ(𝒪X,B)=0\chi(\operatorname{\mathcal{O}}_{X},B)=0, hence ch(B)=ch()\mathrm{ch}(B)=\mathrm{ch}(\mathcal{E}^{\vee}) and by Lemma 7.7 we have BB\cong\mathcal{E}^{\vee}. But this implies Hom(𝒪X,A[1])=k5\operatorname{Hom}(\operatorname{\mathcal{O}}_{X},A[1])=k^{5} since F𝒜XF\in\mathcal{A}_{X}, which contradicts Claim 3.

When 2a1-2\leq a\leq 1, we have μα,β0(A)>μα,β0([2])\mu^{0}_{\alpha,\beta}(A)>\mu^{0}_{\alpha,\beta}(\mathcal{E}[2]). Since AA is σα,β0\sigma^{0}_{\alpha,\beta}-semistable, we have Hom(A,[2])=Hom(,A[1])=0\operatorname{Hom}(A,\mathcal{E}[2])=\operatorname{Hom}(\mathcal{E}^{\vee},A[1])=0. Thus RHom(,A)=0\operatorname{RHom}^{\bullet}(\mathcal{E}^{\vee},A)=0 by Claim 2. But this contradicts Claim 3 since χ(,A)=32a\chi(\mathcal{E}^{\vee},A)=3-2a.

When a2a\geq 2, applying Theorem 4.7 to BB, we have a=2a=2. Thus ch2(B)=ch2(𝒬[1])\mathrm{ch}_{\leq 2}(B)=\mathrm{ch}_{\leq 2}(\mathcal{Q}^{\vee}[1]). By Claim 3, we know that RHom(𝒪X,B)=0\operatorname{RHom}^{\bullet}(\operatorname{\mathcal{O}}_{X},B)=0 and we get ch(B)=ch(𝒬[1])\mathrm{ch}(B)=\mathrm{ch}(\mathcal{Q}^{\vee}[1]). Thus χ(,B)=hom(,B)>0\chi(\mathcal{E}^{\vee},B)=\hom(\mathcal{E}^{\vee},B)>0. Therefore, if we apply Hom(,B)\operatorname{Hom}(-,B) to the exact sequence 0𝒬𝒪X500\to\mathcal{Q}^{\vee}\to\operatorname{\mathcal{O}}_{X}^{\oplus 5}\to\mathcal{E}^{\vee}\to 0, we obtain hom(𝒬[1],B)>0\hom(\mathcal{Q}^{\vee}[1],B)>0. Now by stability, we have B𝒬[1]B\cong\mathcal{Q}^{\vee}[1]. Now ch(A)=ch([2])\mathrm{ch}(A)=\mathrm{ch}(\mathcal{E}[2]). By Claim 2 and Claim 3, we have ext1(,A)=hom(A,[2])=1\operatorname{ext}^{1}(\mathcal{E}^{\vee},A)=\hom(A,\mathcal{E}[2])=1. Since AA is σα,β0\sigma^{0}_{\alpha,\beta}-semistable and [2]\mathcal{E}[2] is σα,β0\sigma^{0}_{\alpha,\beta}-stable, we have A[2]A\cong\mathcal{E}[2]. ∎

Proposition 7.11.

Let XX be a GM threefold. Then every object in the moduli space σ(𝒜X,x)\mathcal{M}_{\sigma}(\mathcal{A}_{X},-x) is of form pr(IC)[1]\mathrm{pr}(I_{C})[1] for a conic CXC\subset X.

Proof.

Note that hom(𝒬[1],[2])=1\mathrm{hom}(\mathcal{Q}^{\vee}[1],\mathcal{E}[2])=1. Then the result follows from Proposition 7.9 and Proposition 7.10. ∎

Now we are ready to realize the Bridgeland moduli space σ(𝒜X,x)\mathcal{M}_{\sigma}(\mathcal{A}_{X},-x) as the contraction 𝒞m(X)\mathcal{C}_{m}(X) of the Fano surface 𝒞(X)\mathcal{C}(X):

Theorem 7.12.

Let XX be a GM threefold and σ\sigma a Serre-invariant stability condition on 𝒜X\mathcal{A}_{X}. The projection functor pr:Db(X)𝒜X\mathrm{pr}\colon\mathrm{D}^{b}(X)\rightarrow\mathcal{A}_{X} induces a surjective morphism p:𝒞(X)σ(𝒜X,x)p\colon\mathcal{C}(X)\rightarrow\mathcal{M}_{\sigma}(\mathcal{A}_{X},-x), where pp is

  • a blow-down morphism to a smooth point when XX is ordinary;

  • a contraction of the component 2\mathbb{P}^{2} to a singular point when XX is special.

In particular, when XX is general and ordinary, σ(𝒜X,x)\mathcal{M}_{\sigma}(\mathcal{A}_{X},-x) is isomorphic to the minimal model 𝒞m(X)\mathcal{C}_{m}(X) of the Fano surface of conics on XX. When XX is general and special, the moduli space σ(𝒜X,x)\mathcal{M}_{\sigma}(\mathcal{A}_{X},-x) has only one singular point.

Proof.

By Lemma 7.4 and Lemma 7.6, pr(IC)[1]\mathrm{pr}(I_{C})[1] is σ\sigma-stable for any conic CXC\subset X. Then we obtain a morphism p:𝒞(X)σ(𝒜X,x)p\colon\mathcal{C}(X)\to\mathcal{M}_{\sigma}(\mathcal{A}_{X},-x). Moreover, Proposition 7.11 implies that pp is surjective.

Now according to Proposition 7.1, the family of conics CXC\subset X with the property that IC𝒜XI_{C}\not\in\mathcal{A}_{X} is parametrized by the line LσL_{\sigma} when XX is ordinary, and the component 2\mathbb{P}^{2} when XX is special. Since pr(IC)[1]π(𝒬)[1]\mathrm{pr}(I_{C})[1]\cong\pi^{\prime}(\mathcal{Q}^{\vee})[1] for IC𝒜XI_{C}\notin\mathcal{A}_{X} by Proposition 7.2, we know that p(Lσ)=[π(𝒬)[1]]p(L_{\sigma})=[\pi^{\prime}(\mathcal{Q}^{\vee})[1]] when XX is ordinary, and p(2)=[π(𝒬)[1]]p(\mathbb{P}^{2})=[\pi^{\prime}(\mathcal{Q}^{\vee})[1]] when XX is special, where [π(𝒬)[1]]σ(𝒜X,x)[\pi^{\prime}(\mathcal{Q}^{\vee})[1]]\in\mathcal{M}_{\sigma}(\mathcal{A}_{X},-x) is the point represented by the object π(𝒬)[1]\pi^{\prime}(\mathcal{Q}^{\vee})[1]. Thus pp is a blow-down morphism to a smooth point when XX is ordinary and a contraction of the component 2\mathbb{P}^{2} to a singular point when XX is special by Lemma 5.6.

When XX is general and ordinary, the Fano surface 𝒞(X)\mathcal{C}(X) is smooth by Theorem 6.4. Thus σ(𝒜X,x)\mathcal{M}_{\sigma}(\mathcal{A}_{X},-x) is a smooth surface obtained by blowing down a smooth rational curve LσL_{\sigma} on the smooth irreducible projective surface 𝒞(X)\mathcal{C}(X). This implies that σ(𝒜X,x)\mathcal{M}_{\sigma}(\mathcal{A}_{X},-x) is also a smooth irreducible projective surface. On the other hand, it is known that there is a unique rational curve Lσ𝒞(X)L_{\sigma}\subset\mathcal{C}(X) and it is the unique exceptional curve by Lemma 6.5. Thus σ(𝒜X,x)\mathcal{M}_{\sigma}(\mathcal{A}_{X},-x) is isomorphic to the minimal model 𝒞m(X)\mathcal{C}_{m}(X) of Fano surface of conics on XX.

When XX is general and special, the last statement follows from Theorem 6.11 and Lemma 7.3. ∎

7.2. Involutions on 𝒞m(X)\mathcal{C}_{m}(X)

In this section, we are going to describe the involution ι\iota on 𝒞m(X)\mathcal{C}_{m}(X) in Theorem 6.6, described in [13, Section 5.2] using the involution on 𝒜X\mathcal{A}_{X}. Recall that there is a natural involutive autoequivalence functor of 𝒜X\mathcal{A}_{X}, denoted by τ𝒜\tau_{\mathcal{A}} (cf. Remark 3.2). When XX is special, it is induced by the natural involution τ\tau on XX, which comes from the double cover XYX\to Y. In this case it is easy to see that τ𝒜(pr(IC))pr(Iτ(C))\tau_{\mathcal{A}}(\mathrm{pr}(I_{C}))\cong\mathrm{pr}(I_{\tau(C)}).

When XX is ordinary, the situation is more subtle. In the following, we describe the action of τ𝒜\tau_{\mathcal{A}} on the projection into 𝒜X\mathcal{A}_{X} of an ideal sheaf of a conic pr(IC)\mathrm{pr}(I_{C}) in this case, and compare with the involution ι\iota on 𝒞m(X)\mathcal{C}_{m}(X) described in [13, Section 5.2].

Proposition 7.13.

Let XX be an ordinary GM threefold and CC a conic on XX.

  1. (1)

    If IC𝒜XI_{C}\in\mathcal{A}_{X}, then τ𝒜(IC)\mathcal{\tau}_{\mathcal{A}}(I_{C}) is either

    1. (a).

      ICI_{C^{\prime}} such that CC=Z(s)C\cup C^{\prime}=Z(s) for sH0()s\in H^{0}(\mathcal{E}^{\vee}), where Z(s)Z(s) is the zero locus of the section ss;

    2. (b).

      or π(𝒬)\pi^{\prime}(\mathcal{Q}^{\vee}), and in this case CC is the ρ\rho-conic

  2. (2)

    If IC𝒜XI_{C}\notin\mathcal{A}_{X}, then τ𝒜(pr(IC))IC′′\tau_{\mathcal{A}}(\mathrm{pr}(I_{C}))\cong I_{C^{\prime\prime}} for the ρ\rho-conic C′′XC^{\prime\prime}\subset X.

Therefore, the involution induced by τ𝒜\tau_{\mathcal{A}} on 𝒞m(X)\mathcal{C}_{m}(X) is the same as ι\iota in Theorem 6.6.

Remark 7.14.

We can define a birational involution on 𝒞(X)\mathcal{C}(X) for any GM threefold XX as in Proposition 7.13 (1)(a), which is regular on the locus of conics CC with hom(,IC)=1\hom(\mathcal{E},I_{C})=1.

We first state some lemmas which we require for the proof of the proposition above.

Lemma 7.15.

Let XX be an ordinary GM threefold and CC be the ρ\rho-conic on XX. Then the natural morphism s:2ICs^{\prime}\colon\mathcal{E}^{\oplus 2}\to I_{C} is surjective and there is a short exact sequence

0𝒬(H)2sIC0.0\rightarrow\mathcal{Q}(-H)\rightarrow\mathcal{E}^{\oplus 2}\xrightarrow{s^{\prime}}I_{C}\rightarrow 0.
Proof.

By Lemma 6.3, we have Hom(,IC)=k2\operatorname{Hom}(\mathcal{E},I_{C})=k^{2}. Thus, taking two linearly independent elements in Hom(,IC)\operatorname{Hom}(\mathcal{E},I_{C}), we have a natural map s:2ICs^{\prime}\colon\mathcal{E}^{\oplus 2}\to I_{C}. Moreover, since C=Gr(2,3)\langle C\rangle=\mathrm{Gr}(2,3) and CX=C\langle C\rangle\cap X=C, we know that ss^{\prime} is surjective. Let K:=ker(s)K:=\ker(s^{\prime}). Then it is not hard to see that ch(K)=ch(𝒬(H))\mathrm{ch}(K)=\mathrm{ch}(\mathcal{Q}(-H)). Note that Hom(,K)=0\operatorname{Hom}(\mathcal{E},K)=0 and KK is reflexive.

We claim that KK is μ\mu-semistable. Indeed, suppose KK is not μ\mu-semistable and let KK^{\prime} be its maximal destabilizing subsheaf. Then KK^{\prime} is also reflexive. Since Hom(,K)=0\operatorname{Hom}(\mathcal{E},K)=0, we have KK^{\prime}\neq\mathcal{E}. By the stability of \mathcal{E} and the fact that K2K\subset\mathcal{E}^{\oplus 2}, we know that μ(K)=12\mu(K^{\prime})=-\frac{1}{2}. Since Hom(K,)0\operatorname{Hom}(K^{\prime},\mathcal{E})\neq 0, by the stability of KK^{\prime} and \mathcal{E} we have KK^{\prime}\subset\mathcal{E}. Thus from ch1(K)=ch1()\mathrm{ch}_{\leq 1}(K^{\prime})=\mathrm{ch}_{\leq 1}(\mathcal{E}) we know that /K\mathcal{E}/K^{\prime} is supported in codimension 2\geq 2, which gives a contradiction since \mathcal{E} and KK^{\prime} are both reflexive.

Now the result follows from Lemma 7.8, since K(H)K(H) is μ\mu-semistable with ch(K(H))=ch(𝒬)\mathrm{ch}(K(H))=\mathrm{ch}(\mathcal{Q}). ∎

Lemma 7.16.

Let XX be an ordinary GM threefold. Let CC be a conic on XX. Then

𝐋(IC)={𝔻(IC)𝒪X(H)[1],RHom(,IC)=kπ(),RHom(,IC)=k2k[1]\bm{\mathrm{L}}_{\mathcal{E}}(I_{C})=\begin{cases}\mathbb{D}(I_{C^{\prime}})\otimes\mathcal{O}_{X}(-H)[1],&\mathrm{RHom}^{\bullet}(\mathcal{E},I_{C})=k\\ \pi(\mathcal{E}),&\mathrm{RHom}^{\bullet}(\mathcal{E},I_{C})=k^{2}\oplus k[-1]\end{cases}

such that CC=Z(s)C\cup C^{\prime}=Z(s) for sH0()s\in H^{0}(\mathcal{E}^{\vee}), where Z(s)Z(s) is the zero locus of the section ss

Proof.

By Lemma 6.3, we have that RHom(,IC)\mathrm{RHom}^{\bullet}(\mathcal{E},I_{C}) is either kk or k2k[1]k^{2}\oplus k[-1]. If RHom(,IC)=k\mathrm{RHom}^{\bullet}(\mathcal{E},I_{C})=k, then we have the triangle

IC𝐋(IC).\mathcal{E}\rightarrow I_{C}\rightarrow\bm{\mathrm{L}}_{\mathcal{E}}(I_{C}).

Taking cohomology with respect to the standard heart we get

01(𝐋(IC))𝑠IC0(𝐋(IC))0.0\rightarrow\mathcal{H}^{-1}(\bm{\mathrm{L}}_{\mathcal{E}}(I_{C}))\rightarrow\mathcal{E}\xrightarrow{s}I_{C}\rightarrow\mathcal{H}^{0}(\bm{\mathrm{L}}_{\mathcal{E}}(I_{C}))\rightarrow 0.

The image of the map ss is the ideal sheaf of an elliptic quartic D=Z(s)D=Z(s) for sH0()s\in H^{0}(\mathcal{E}^{\vee}), thus we have following two short exact sequences: 01(𝐋(IC))ID00\rightarrow\mathcal{H}^{-1}(\bm{\mathrm{L}}_{\mathcal{E}}(I_{C}))\rightarrow\mathcal{E}\rightarrow I_{D}\rightarrow 0 and 0IDIC0(𝐋(IC))00\rightarrow I_{D}\rightarrow I_{C}\rightarrow\mathcal{H}^{0}(\bm{\mathrm{L}}_{\mathcal{E}}(I_{C}))\rightarrow 0. Then 1(𝐋(IC))\mathcal{H}^{-1}(\bm{\mathrm{L}}_{\mathcal{E}}(I_{C})) is a torsion-free sheaf of rank 11 with the same Chern character as 𝒪X(H)\mathcal{O}_{X}(-H). It is easy to show that it must be 𝒪X(H)\mathcal{O}_{X}(-H). On the other hand 0(𝐋(IC))\mathcal{H}^{0}(\bm{\mathrm{L}}_{\mathcal{E}}(I_{C})) is supported on the residual curve CC^{\prime} of CC in DD and 0(𝐋(IC))𝒪C(H)\mathcal{H}^{0}(\bm{\mathrm{L}}_{\mathcal{E}}(I_{C}))\cong\mathcal{O}_{C^{\prime}}(-H). Thus we have the triangle

𝒪X(H)[1]𝐋(IC)𝒪C(H)\mathcal{O}_{X}(-H)[1]\rightarrow\bm{\mathrm{L}}_{\mathcal{E}}(I_{C})\rightarrow\mathcal{O}_{C^{\prime}}(-H)

and we observe that 𝐋(IC)\bm{\mathrm{L}}_{\mathcal{E}}(I_{C}) is exactly the twisted derived dual of the ideal sheaf ICI_{C^{\prime}} of a conic CXC^{\prime}\subset X, i.e. 𝐋(IC)𝔻(IC)𝒪X(H)[1]\bm{\mathrm{L}}_{\mathcal{E}}(I_{C})\cong\mathbb{D}(I_{C^{\prime}})\otimes\mathcal{O}_{X}(-H)[1].

If RHom(,IC)=k2k[1]\mathrm{RHom}^{\bullet}(\mathcal{E},I_{C})=k^{2}\oplus k[-1], then we have the triangle

2[1]IC𝐋(IC).\mathcal{E}^{2}\oplus\mathcal{E}[-1]\rightarrow I_{C}\rightarrow\bm{\mathrm{L}}_{\mathcal{E}}(I_{C}).

Taking the long exact sequence in cohomology with respect to the standard heart, we get

01(𝐋(IC))2sIC0(𝐋(IC))0.0\rightarrow\mathcal{H}^{-1}(\bm{\mathrm{L}}_{\mathcal{E}}(I_{C}))\rightarrow\mathcal{E}^{2}\xrightarrow{s^{\prime}}I_{C}\rightarrow\mathcal{H}^{0}(\bm{\mathrm{L}}_{\mathcal{E}}(I_{C}))\rightarrow\mathcal{E}\rightarrow 0.

Now by Lemma 7.15, ss^{\prime} is surjective and the cohomology objects are given by 1(𝐋(IC))𝒬(H)\mathcal{H}^{-1}(\bm{\mathrm{L}}_{\mathcal{E}}(I_{C}))\cong\mathcal{Q}(-H) and 0(𝐋(IC))\mathcal{H}^{0}(\bm{\mathrm{L}}_{\mathcal{E}}(I_{C}))\cong\mathcal{E}, which implies that 𝐋(IC)π()\bm{\mathrm{L}}_{\mathcal{E}}(I_{C})\cong\pi(\mathcal{E}). ∎

Proof of Proposition 7.13.

Since τ𝒜τ𝒜id\tau_{\mathcal{A}}\circ\tau_{\mathcal{A}}\cong\operatorname{id}, we have τ𝒜τ𝒜1\tau_{\mathcal{A}}\cong\tau^{-1}_{\mathcal{A}}. By Proposition 2.6, we have τ𝒜τ𝒜1𝐋𝒪X𝐋(𝒪X(H))[1]\tau_{\mathcal{A}}\cong\tau_{\mathcal{A}}^{-1}\cong\bm{\mathrm{L}}_{\mathcal{O}_{X}}\circ\bm{\mathrm{L}}_{\mathcal{E}^{\vee}}(-\otimes\mathcal{O}_{X}(H))[-1]. Then

τ𝒜(IC)\displaystyle\tau_{\mathcal{A}}(I_{C}) 𝐋𝒪X𝐋(IC𝒪X(H))[1]\displaystyle\cong\bm{\mathrm{L}}_{\mathcal{O}_{X}}\circ\bm{\mathrm{L}}_{\mathcal{E}^{\vee}}(I_{C}\otimes\mathcal{O}_{X}(H))[-1]
𝐋𝒪X(𝐋(IC)𝒪X(H))[1].\displaystyle\cong\bm{\mathrm{L}}_{\mathcal{O}_{X}}(\bm{\mathrm{L}}_{\mathcal{E}}(I_{C})\otimes\mathcal{O}_{X}(H))[-1].

The left mutation 𝐋(IC)\bm{\mathrm{L}}_{\mathcal{E}}(I_{C}) is given by

RHom(,IC)IC𝐋(IC).\mathrm{RHom}^{\bullet}(\mathcal{E},I_{C})\otimes\mathcal{E}\rightarrow I_{C}\rightarrow\bm{\mathrm{L}}_{\mathcal{E}}(I_{C}).

Note that by Lemma 6.3, RHom(,IC)\mathrm{RHom}^{\bullet}(\mathcal{E},I_{C}) is either kk or k2k[1]k^{2}\oplus k[-1], and in the latter case CC is the unique ρ\rho-conic. Then by Lemma 7.16,

𝐋(IC)={𝔻(IC)𝒪X(H)[1],RHom(,IC)=kπ(),RHom(,IC)=k2k[1]\bm{\mathrm{L}}_{\mathcal{E}}(I_{C})=\begin{cases}\mathbb{D}(I_{C^{\prime}})\otimes\mathcal{O}_{X}(-H)[1],&\mathrm{RHom}^{\bullet}(\mathcal{E},I_{C})=k\\ \pi(\mathcal{E}),&\mathrm{RHom}^{\bullet}(\mathcal{E},I_{C})=k^{2}\oplus k[-1]\end{cases}

If RHom(,IC)=k\mathrm{RHom}^{\bullet}(\mathcal{E},I_{C})=k, then τ𝒜(IC)𝐋𝒪X(𝔻(IC))\tau_{\mathcal{A}}(I_{C})\cong\bm{\mathrm{L}}_{\mathcal{O}_{X}}(\mathbb{D}(I_{C^{\prime}})). We have the triangle

RHom(𝒪X,𝔻(IC))𝒪X𝔻(IC)𝐋𝒪X(𝔻(IC)).\mathrm{RHom}^{\bullet}(\mathcal{O}_{X},\mathbb{D}(I_{C^{\prime}}))\otimes\mathcal{O}_{X}\rightarrow\mathbb{D}(I_{C^{\prime}})\rightarrow\bm{\mathrm{L}}_{\mathcal{O}_{X}}(\mathbb{D}(I_{C^{\prime}})).

Note that RHom(𝒪X,𝔻(IC))RHom(IC,𝒪X)=kk[1]\mathrm{RHom}^{\bullet}(\mathcal{O}_{X},\mathbb{D}(I_{C^{\prime}}))\cong\mathrm{RHom}^{\bullet}(I_{C^{\prime}},\mathcal{O}_{X})=k\oplus k[-1]. Then we have the triangle

𝒪X𝒪X[1]𝔻(IC)𝐋𝒪X(𝔻(IC)).\mathcal{O}_{X}\oplus\mathcal{O}_{X}[-1]\rightarrow\mathbb{D}(I_{C^{\prime}})\rightarrow\bm{\mathrm{L}}_{\mathcal{O}_{X}}(\mathbb{D}(I_{C^{\prime}})). (4)

The derived dual 𝔻(IC)\mathbb{D}(I_{C^{\prime}}) is given by the triangle 𝒪X𝔻(IC)𝒪C[1]\mathcal{O}_{X}\rightarrow\mathbb{D}(I_{C^{\prime}})\rightarrow\mathcal{O}_{C^{\prime}}[-1]. Then taking cohomology with respect to the standard heart of triangle (4) we have the long exact sequence

0=1(𝔻(IC))1(𝐋𝒪X(𝔻(IC)))𝒪X𝒪X0=\mathcal{H}^{-1}(\mathbb{D}(I_{C^{\prime}}))\rightarrow\mathcal{H}^{-1}(\bm{\mathrm{L}}_{\mathcal{O}_{X}}(\mathbb{D}(I_{C^{\prime}})))\rightarrow\mathcal{O}_{X}\rightarrow\mathcal{O}_{X}
0(𝐋𝒪X(𝔻(IC)))𝒪X𝒪C1(𝐋𝒪X(𝔻(IC)))0.\rightarrow\mathcal{H}^{0}(\bm{\mathrm{L}}_{\mathcal{O}_{X}}(\mathbb{D}(I_{C^{\prime}})))\rightarrow\mathcal{O}_{X}\rightarrow\mathcal{O}_{C^{\prime}}\rightarrow\mathcal{H}^{1}(\bm{\mathrm{L}}_{\mathcal{O}_{X}}(\mathbb{D}(I_{C^{\prime}})))\rightarrow 0.

Thus we have 1(𝐋𝒪X(𝔻(IC)))=0\mathcal{H}^{-1}(\bm{\mathrm{L}}_{\mathcal{O}_{X}}(\mathbb{D}(I_{C^{\prime}})))=0, 1(𝐋𝒪X(𝔻(IC)))=0\mathcal{H}^{1}(\bm{\mathrm{L}}_{\mathcal{O}_{X}}(\mathbb{D}(I_{C^{\prime}})))=0 and 0(𝐋𝒪X(𝔻(IC)))IC\mathcal{H}^{0}(\bm{\mathrm{L}}_{\mathcal{O}_{X}}(\mathbb{D}(I_{C^{\prime}})))\cong I_{C^{\prime}}. Hence τ𝒜(IC)𝐋𝒪X(𝔻(IC))IC\tau_{\mathcal{A}}(I_{C})\cong\bm{\mathrm{L}}_{\mathcal{O}_{X}}(\mathbb{D}(I_{C^{\prime}}))\cong I_{C^{\prime}}.

If RHom(,IC)=k2k[1]\mathrm{RHom}^{\bullet}(\mathcal{E},I_{C})=k^{2}\oplus k[-1], then τ𝒜(IC)𝐋𝒪X𝐋(IC𝒪X(H))[1]𝐋𝒪X(π()𝒪X(H)[1])\tau_{\mathcal{A}}(I_{C})\cong\bm{\mathrm{L}}_{\mathcal{O}_{X}}\circ\bm{\mathrm{L}}_{\mathcal{E}^{\vee}}(I_{C}\otimes\mathcal{O}_{X}(H))[-1]\cong\bm{\mathrm{L}}_{\mathcal{O}_{X}}(\pi(\mathcal{E})\otimes\mathcal{O}_{X}(H)[-1]) by Lemma 7.16. Then using the triangle (2), we have τ𝒜(IC)π(𝒬)\tau_{\mathcal{A}}(I_{C})\cong\pi^{\prime}(\mathcal{Q}^{\vee}). Then (2)(2) follows from τ𝒜τ𝒜1\tau_{\mathcal{A}}\cong\tau^{-1}_{\mathcal{A}}.

Now since τ𝒜=S𝒜X[2]\tau_{\mathcal{A}}=S_{\mathcal{A}_{X}}[-2] and τ𝒜\tau_{\mathcal{A}} acts trivially on 𝒩(𝒜X)\mathcal{N}(\mathcal{A}_{X}), it induces an involution on the Bridgeland moduli space of any class with respect to any Serre-invariant stability condition. In particular, τ𝒜\tau_{\mathcal{A}} induces an involution on 𝒞m(X)σ(𝒜X,x)\mathcal{C}_{m}(X)\cong\mathcal{M}_{\sigma}(\mathcal{A}_{X},-x) by Theorem 7.12. By (1) and (2), this induced involution coincides with ι\iota in Theorem 6.6, described in [13, Section 5.2]. ∎

Remark 7.17.

Smooth τ\tau-conics form an open subscheme UU of 𝒞(X)\mathcal{C}(X). Therefore, the open subscheme Uι(U)U\cap\iota(U) parameterizes smooth τ\tau-conics CC such that their involutive conics in Proposition 7.13 are smooth as well. The same also works for special GM threefolds, but replace τ\tau-conics with conics with hom(,IC)=1\hom(\mathcal{E},I_{C})=1 and IC𝒜XI_{C}\in\mathcal{A}_{X}, which are parametrized by 𝒞(X)2\mathcal{C}(X)\setminus\mathbb{P}^{2}. In other words, for any GM threefold XX, there is a two-dimensional open subscheme 𝒞1𝒞(X)\mathcal{C}_{1}\subset\mathcal{C}(X) parameterizing smooth conics CC with hom(,IC)=1\hom(\mathcal{E},I_{C})=1 such that their involutive conics are smooth.

8. The moduli space MG(2,1,5)M_{G}(2,1,5) for GM threefolds

In this section, we investigate the moduli space of rank 22 Gieseker-semistable sheaves on a GM threefold XX with Chern classes c1=H,c2=5Lc_{1}=H,c_{2}=5L and c3=0c_{3}=0, denoted MGX(2,1,5)M_{G}^{X}(2,1,5). We drop XX from the notation when it is clear from context on which threefold we work. Note that if FMG(2,1,5)F\in M_{G}(2,1,5), then

ch(F)=(2,H,0,56P).\mathrm{ch}(F)=(2,H,0,-\frac{5}{6}P).

We are interested in MG(2,1,5)M_{G}(2,1,5) since it naturally appears in the description of the period fiber in [13]. Our main theorem in Section is Theorem 8.9, which realizes MG(2,1,5)M_{G}(2,1,5) as the Bridgeland moduli space σ(𝒜X,y2x)\mathcal{M}_{\sigma}(\mathcal{A}_{X},y-2x).

First, we prove a classification result of sheaves in MG(2,1,5)M_{G}(2,1,5).

Proposition 8.1.

Let XX be a GM threefold and FMG(2,1,5)F\in M_{G}(2,1,5). Then we have RHom(𝒪X,F)=k4\operatorname{RHom}^{\bullet}(\operatorname{\mathcal{O}}_{X},F)=k^{4} and RHom(𝒪X,F(H))=0\operatorname{RHom}^{\bullet}(\operatorname{\mathcal{O}}_{X},F(-H))=0. Moreover, FF is either a

  1. (1)

    globally generated bundle which fits into a short exact sequence

    0𝒪XFIZ(H)00\rightarrow\operatorname{\mathcal{O}}_{X}\rightarrow F\rightarrow I_{Z}(H)\rightarrow 0

    where ZZ is a projective normal smooth elliptic quintic curve;

  2. (2)

    non-locally free sheaf with a short exact sequence

    0F𝒪L00\rightarrow F\rightarrow\mathcal{E}^{\vee}\rightarrow\operatorname{\mathcal{O}}_{L}\rightarrow 0

    where LL is a line on XX. Moreover, FF is uniquely determined by LL.

Remark 8.2.

In [13, Section 8], they also did computations for non-globally generated bundles in MGX(2,1,5)M^{X}_{G}(2,1,5). However, in the following proof, we will show such sheaves do not exist.

Proof.

The first statement follows from [4, Proposition 3.5 (1)] and the fact χ(F)=4\chi(F)=4. (1) and (2) also follow from [4, Proposition 3.5] or the argument in [13, Section 8]. So we only need to prove the non-existence of non-globally generated bundles in MGX(2,1,5)M^{X}_{G}(2,1,5). If FMGX(2,1,5)F\in M^{X}_{G}(2,1,5) is a non-globally generated bundle, then as showed in [13, Section 8], we have an exact sequence

0F𝒪X4𝑎𝒪L0.0\to F^{\vee}\to\operatorname{\mathcal{O}}_{X}^{\oplus 4}\to\mathcal{E}^{\vee}\xrightarrow{a}\operatorname{\mathcal{O}}_{L}\to 0.

By (2), we know that E:=ker(a)E:=\ker(a) is a non-locally free stable sheaf and EMGX(2,1,5)E\in M^{X}_{G}(2,1,5). Thus we have an exact sequence 0F𝒪X4E00\to F^{\vee}\to\operatorname{\mathcal{O}}_{X}^{\oplus 4}\to E\to 0. In particular, EE is generated by global sections.

However, we also have the following commutative diagram of exact sequences:

0{0}𝒪X4{{\operatorname{\mathcal{O}}_{X}^{\oplus 4}}}𝒪X5{{\operatorname{\mathcal{O}}_{X}^{\oplus 5}}}𝒪X{{\operatorname{\mathcal{O}}_{X}}}0{0}0{0}E{E}{{\mathcal{E}^{\vee}}}𝒪L{{\operatorname{\mathcal{O}}_{L}}}0{0}ev\scriptstyle{\mathrm{ev}}

where ev:𝒪X4E\mathrm{ev}\colon\operatorname{\mathcal{O}}_{X}^{\oplus 4}\to E is the evaluation map. Then using Snake Lemma, we have an exact sequence

0ker(ev)𝒬𝑠ILcok(ev)0.0\to\ker(\mathrm{ev})\to\mathcal{Q}^{\vee}\xrightarrow{s}I_{L}\to\mathrm{cok}(\mathrm{ev})\to 0.

As shown in Lemma 6.3 and Proposition 7.1, the image of ss is the zero locus of a non-zero section of 𝒬\mathcal{Q}. It is a σ\sigma-conic when XX is ordinary, and a preimage of a line on YY when XX is special. Hence in both cases, im(s)\mathrm{im}(s) is an ideal sheaf of a conic, and ss is not surjective. Therefore, ev\mathrm{ev} is not surjective as well and we get a contradiction. ∎

A natural question to ask is what Bridgeland moduli space we get after projecting a sheaf in MG(2,1,5)M_{G}(2,1,5) into the Kuznetsov component. Since it is easier in this setting, we will work with the alternative Kuznetsov component 𝒜X\mathcal{A}_{X} in this section. Our analysis of the projections of objects in MG(2,1,5)M_{G}(2,1,5) is based on the three cases listed in Proposition 8.1. We begin with a Hom-vanishing result.

Lemma 8.3.

Let XX be a GM threefold and FMGX(2,1,5)F\in M^{X}_{G}(2,1,5). Then we have RHom(,F)=0\operatorname{RHom}^{\bullet}(\mathcal{E}^{\vee},F)=0.

Proof.

By Serre duality and the stability of \mathcal{E}^{\vee} and FF, we have Hom(,F)=Ext3(,F)=0\operatorname{Hom}(\mathcal{E}^{\vee},F)=\operatorname{Ext}^{3}(\mathcal{E}^{\vee},F)=0. Since χ(,F)=0\chi(\mathcal{E}^{\vee},F)=0, we only need to show that Ext2(,F)=0\operatorname{Ext}^{2}(\mathcal{E}^{\vee},F)=0 or Ext2(,F)=0\operatorname{Ext}^{2}(\mathcal{E}^{\vee},F)=0. By Serre duality, we have Ext2(,F)=Hom(F,[1])\operatorname{Ext}^{2}(\mathcal{E}^{\vee},F)=\operatorname{Hom}(F,\mathcal{E}[1]). Since ch10(F)=ch10([1])=1\mathrm{ch}^{0}_{1}(F)=\mathrm{ch}^{0}_{1}(\mathcal{E}[1])=1, by Lemma 4.5 we know that FF and [1]\mathcal{E}[1] are both σα,0\sigma_{\alpha,0}-stable for any α>0\alpha>0. Then Hom(F,[1])=0\operatorname{Hom}(F,\mathcal{E}[1])=0 since μα,0(F)>μα,0([1])\mu_{\alpha,0}(F)>\mu_{\alpha,0}(\mathcal{E}[1]) when 0<α0<\alpha is sufficiently small. ∎

We are now ready to give an explicit description of pr(F)\mathrm{pr}(F), for all objects FMG(2,1,5)F\in M_{G}(2,1,5). Recall that for any line LXL\subset X, we have 𝒬|L𝒪L2𝒪L(1)\mathcal{Q}|_{L}\cong\operatorname{\mathcal{O}}_{L}^{\oplus 2}\oplus\operatorname{\mathcal{O}}_{L}(1). Hence LL is contained in a unique σ\sigma-conic CC. We define the residue line of LL to be the support of 𝒪C𝒪L\operatorname{\mathcal{O}}_{C}\twoheadrightarrow\operatorname{\mathcal{O}}_{L}. Note that when CC is a double line, we have L=LL^{\prime}=L.

Lemma 8.4.

Let XX be a GM threefold and FMG(2,1,5)F\in M_{G}(2,1,5).

  • If FF is a globally generated bundle, then

    pr(F)ker(ev)[1],\mathrm{pr}(F)\cong\ker(\mathrm{ev})[1],

    where ev:𝒪X4F\mathrm{ev}\colon\operatorname{\mathcal{O}}_{X}^{\oplus 4}\twoheadrightarrow F is the evaluation map.

  • If FF is a non-locally free sheaf determined by a line LXL\subset X, then pr(F)\mathrm{pr}(F) is the unique object fits into a non-trivial exact triangle

    [1]pr(F)𝒪L(1),\mathcal{E}[1]\to\mathrm{pr}(F)\to\operatorname{\mathcal{O}}_{L^{\prime}}(-1),

    where LL^{\prime} is the residue line of LL.

Proof.

As a result of Lemma 8.3, 𝐋F=F\bm{\mathrm{L}}_{\mathcal{E}^{\vee}}F=F, so pr(F)=𝐋𝒪XF\mathrm{pr}(F)=\bm{\mathrm{L}}_{\operatorname{\mathcal{O}}_{X}}F. By Proposition 8.1 we have RHom(𝒪X,F)=k4\operatorname{RHom}^{\bullet}(\operatorname{\mathcal{O}}_{X},F)=k^{4}, and the triangle defining the left mutation is

𝒪X4evFpr(F).\operatorname{\mathcal{O}}^{\oplus 4}_{X}\xrightarrow{\mathrm{ev}}F\rightarrow\mathrm{pr}(F). (5)

In the cases where FF is globally generated, the evaluation map ev\mathrm{ev} is surjective, so pr(F)=ker(ev)[1]\mathrm{pr}(F)=\ker(\mathrm{ev})[1].

When FF is non-locally free, as in Proposition 8.1, we have an exact sequence

0ker(ev)𝒬𝑠ILcok(ev)0.0\to\ker(\mathrm{ev})\to\mathcal{Q}^{\vee}\xrightarrow{s}I_{L}\to\mathrm{cok}(\mathrm{ev})\to 0.

As shown in Lemma 6.3 and Proposition 7.1, the image of ss is the zero locus of a non-zero section of 𝒬\mathcal{Q}, which is a σ\sigma-conic. Hence by Proposition 7.1, we obtain ker(ev)=\ker(\mathrm{ev})=\mathcal{E} and cok(ev)=𝒪L(1)\mathrm{cok}(\mathrm{ev})=\operatorname{\mathcal{O}}_{L^{\prime}}(-1). Since pr(F)𝒜X\mathrm{pr}(F)\in\mathcal{A}_{X}, by Serre duality we have RHom(,pr(F))=RHom(pr(F),)[3]=0\mathrm{RHom}^{\bullet}(\mathcal{E}^{\vee},\mathrm{pr}(F))=\mathrm{RHom}^{\bullet}(\mathrm{pr}(F),\mathcal{E})^{\vee}[-3]=0, which implies such triangle is non-trivial. And the uniqueness follows from Ext2(𝒪L(1),)=H1((1)|L)=k\operatorname{Ext}^{2}(\operatorname{\mathcal{O}}_{L^{\prime}}(-1),\mathcal{E})=H^{1}(\mathcal{E}(-1)|_{L})=k. ∎

8.1. Stability of projection objects

In the following, we prove the stability of pr(F)\mathrm{pr}(F) for any FMGX(2,1,5)F\in M_{G}^{X}(2,1,5).

Lemma 8.5.

The functor pr:Db(X)𝒜X\mathrm{pr}\colon\mathrm{D}^{b}(X)\rightarrow\mathcal{A}_{X} induces isomorphisms of Extk(pr(F1),pr(F2))\mathrm{Ext}^{k}(\mathrm{pr}(F_{1}),\mathrm{pr}(F_{2})) and Extk(F1,F2)\mathrm{Ext}^{k}(F_{1},F_{2}) for all kk and for all F1,F2MG(2,1,5)F_{1},F_{2}\in M_{G}(2,1,5).

Proof.

We apply Hom(F1,)\mathrm{Hom}(F_{1},-) to the exact triangle 𝒪X4F2pr(F2)\mathcal{O}_{X}^{\oplus 4}\rightarrow F_{2}\rightarrow\mathrm{pr}(F_{2}). By adjunction of pr\mathrm{pr} and the inclusion 𝒜XDb(X)\mathcal{A}_{X}\hookrightarrow\mathrm{D}^{b}(X), we have Extk(F1,pr(F2))=Extk(pr(F1),pr(F2))\operatorname{Ext}^{k}(F_{1},\mathrm{pr}(F_{2}))=\operatorname{Ext}^{k}(\mathrm{pr}(F_{1}),\mathrm{pr}(F_{2})) for all kk. Thus we get a long exact sequence

\displaystyle\cdots Extk(F1,𝒪X)4Extk(F1,F2)Extk(pr(F1),pr(F2))Extk+1(F1,𝒪X)4.\displaystyle\rightarrow\mathrm{Ext}^{k}(F_{1},\operatorname{\mathcal{O}}_{X})^{\oplus 4}\rightarrow\mathrm{Ext}^{k}(F_{1},F_{2})\rightarrow\mathrm{Ext}^{k}(\mathrm{pr}(F_{1}),\mathrm{pr}(F_{2}))\rightarrow\mathrm{Ext}^{k+1}(F_{1},\operatorname{\mathcal{O}}_{X})^{\oplus 4}\rightarrow\cdots.

Note that Extk(F1,𝒪X)=Ext3k(𝒪X,F1(H))=0\mathrm{Ext}^{k}(F_{1},\operatorname{\mathcal{O}}_{X})=\operatorname{Ext}^{3-k}(\operatorname{\mathcal{O}}_{X},F_{1}(-H))=0 for all kk by Proposition 8.1. Thus the desired result follows. ∎

Before we show the stability of projection objects, let us recall a classical result:

Proposition 8.6.

Let XX be an ordinary GM threefold and LXL\subset X be a line. Then RHom(𝒪L,𝒪L)=kk[1]\mathrm{RHom}^{\bullet}(\operatorname{\mathcal{O}}_{L},\operatorname{\mathcal{O}}_{L})=k\oplus k[-1] or kk2[1]k[2]k\oplus k^{2}[-1]\oplus k[-2]. Moreover, when XX is general, we always have RHom(𝒪L,𝒪L)=kk[1]\mathrm{RHom}^{\bullet}(\operatorname{\mathcal{O}}_{L},\operatorname{\mathcal{O}}_{L})=k\oplus k[-1].

Proof.

The first statement follows from [50, Lemma 4.2.1] and the second one follows from [50, Theorem 4.2.7]. ∎

Now we are ready to prove the stability of pr(F)\mathrm{pr}(F).

Proposition 8.7.

Let XX be a GM threefold and FMGX(2,1,5)F\in M^{X}_{G}(2,1,5). Then we have RHom(F,F)=kk2[1]\operatorname{RHom}^{\bullet}(F,F)=k\oplus k^{2}[-1] or RHom(F,F)=kk3[1]k[2]\operatorname{RHom}^{\bullet}(F,F)=k\oplus k^{3}[-1]\oplus k[-2]. Hence pr(F)\mathrm{pr}(F) is stable with respect to every Serre-invariant stability condition on 𝒜X\mathcal{A}_{X}.

Moreover,

  1. (1)

    when XX is ordinary, if RHom(F,F)=kk3[1]k[2]\operatorname{RHom}^{\bullet}(F,F)=k\oplus k^{3}[-1]\oplus k[-2] then FF is a non-globally generated bundle or a non-locally free sheaf determined by a line LL, and [L]Γ(X)[L]\in\Gamma(X) is a singular point. In particular, we always have RHom(F,F)=kk2[1]\operatorname{RHom}^{\bullet}(F,F)=k\oplus k^{2}[-1] when XXis general;

  2. (2)

    when XX is special, RHom(F,F)=kk3[1]k[2]\operatorname{RHom}^{\bullet}(F,F)=k\oplus k^{3}[-1]\oplus k[-2] if and only if τFF\tau^{*}F\cong F, where τ\tau is the natural involution on XX.

Proof.

First, we assume that XX is ordinary. We have hom(F,F)=1\hom(F,F)=1 and ext3(F,F)=0\operatorname{ext}^{3}(F,F)=0 by Serre duality and the stability of FF. Since χ(F,F)=1\chi(F,F)=-1, we need to prove ext2(F,F)=0\operatorname{ext}^{2}(F,F)=0 or 11.

When FF is a globally generated bundle, by the proof of [13, Theorem 8.2], we have ext1(F,F)=2\mathrm{ext}^{1}(F,F)=2 and ext2(F,F)=0\mathrm{ext}^{2}(F,F)=0. When FF is non-locally free, there is a mistake made in the proof of [13, Theorem 8.2] and we fix it here. From Proposition 8.1, we have an exact sequence 0F𝒪L00\to F\to\mathcal{E}^{\vee}\to\operatorname{\mathcal{O}}_{L}\to 0. Since RHom(,F)=0\mathrm{RHom}^{\bullet}(\mathcal{E}^{\vee},F)=0 by Lemma 8.3, applying Hom(,F)\operatorname{Hom}(-,F) to this exact sequence, we get Extk(F,F)=Extk+1(𝒪L,F)\operatorname{Ext}^{k}(F,F)=\operatorname{Ext}^{k+1}(\operatorname{\mathcal{O}}_{L},F) for any kk. Now applying Hom(𝒪L,)\operatorname{Hom}(\operatorname{\mathcal{O}}_{L},-) to this exact sequence, we get a long exact sequence

Ext2(𝒪L,𝒪L)Ext3(𝒪L,F)Ext3(𝒪L,)0.\cdots\to\operatorname{Ext}^{2}(\operatorname{\mathcal{O}}_{L},\operatorname{\mathcal{O}}_{L})\to\operatorname{Ext}^{3}(\operatorname{\mathcal{O}}_{L},F)\to\operatorname{Ext}^{3}(\operatorname{\mathcal{O}}_{L},\mathcal{E}^{\vee})\to 0.

By Serre duality, we have Ext3(𝒪L,)=H0((1)|L)=0\operatorname{Ext}^{3}(\operatorname{\mathcal{O}}_{L},\mathcal{E}^{\vee})=H^{0}(\mathcal{E}(-1)|_{L})=0. Then from Proposition 8.6, we have ext2(F,F)=ext3(𝒪L,F)ext2(𝒪L,𝒪L)1\operatorname{ext}^{2}(F,F)=\operatorname{ext}^{3}(\operatorname{\mathcal{O}}_{L},F)\leq\operatorname{ext}^{2}(\operatorname{\mathcal{O}}_{L},\operatorname{\mathcal{O}}_{L})\leq 1. Moreover, if ext2(F,F)=1\operatorname{ext}^{2}(F,F)=1, then ext2(𝒪L,𝒪L)=1\operatorname{ext}^{2}(\operatorname{\mathcal{O}}_{L},\operatorname{\mathcal{O}}_{L})=1. In other words, [L]Σ(X)[L]\in\Sigma(X) is a singular point. This proves (1).

Now we assume that XX is special. Then by Lemma 8.5 and Serre duality in 𝒦u(X)\mathcal{K}u(X), we have

Ext2(F,F)\displaystyle\mathrm{Ext}^{2}(F,F) Ext2(pr(F),pr(F))\displaystyle\cong\mathrm{Ext}^{2}(\mathrm{pr}(F),\mathrm{pr}(F))
Hom(pr(F),τ𝒜(pr(F)))\displaystyle\cong\mathrm{Hom}(\mathrm{pr}(F),\tau_{\mathcal{A}}(\mathrm{pr}(F)))
Hom(pr(F),pr(τF))Hom(F,τF),\displaystyle\cong\mathrm{Hom}(\mathrm{pr}(F),\mathrm{pr}(\tau^{*}F))\cong\operatorname{Hom}(F,\tau^{*}F),

where τ\tau is the involution on XX induced by the double cover. Thus when FτFF\cong\tau^{*}F, we have Ext2(F,F)=k\mathrm{Ext}^{2}(F,F)=k, and Ext2(F,F)=0\mathrm{Ext}^{2}(F,F)=0 otherwise. Since Ext3(F,F)=0\operatorname{Ext}^{3}(F,F)=0 and Hom(F,F)=k\operatorname{Hom}(F,F)=k, the result follows from χ(F,F)=1\chi(F,F)=-1.

Finally, the stability of pr(F)\mathrm{pr}(F) follows from Lemma 8.5 and Proposition 4.12. ∎

8.2. Involutions on MG(2,1,5)M_{G}(2,1,5)

In this subsection, we briefly recall the involutions that exist on MG(2,1,5)M_{G}(2,1,5) and compare it with the one induced by τ𝒜\tau_{\mathcal{A}}. Let FF be a globally generated vector bundle, and consider the short exact sequence

0ker(ev)H0(X,F)𝒪XevF0.0\rightarrow\ker(\mathrm{ev})\rightarrow H^{0}(X,F)\otimes\operatorname{\mathcal{O}}_{X}\xrightarrow{\mathrm{ev}}F\rightarrow 0.

Note that ker(ev)\ker(\mathrm{ev}) is a rank 22 vector bundle with c1=Hc_{1}=-H and c2=5Lc_{2}=5L and no global sections, hence ker(ev)MG(2,1,5)\ker(\mathrm{ev})^{\vee}\in M_{G}(2,1,5). Define ιF:=ker(ev)\iota F:=\ker(\mathrm{ev})^{\vee}. This bundle ιF\iota F is globally generated, and we have H0(X,ιF)H0(X,F)H^{0}(X,\iota F)\cong H^{0}(X,F)^{\vee} [13, p. 29]. This defines a birational involution on MGX(2,1,5)M_{G}^{X}(2,1,5).

Note that there is no non-globally generated bundle in MGX(2,1,5)M^{X}_{G}(2,1,5) by Proposition 8.1, then the definition of ι\iota on the non-locally free locus in [13, Theorem 8.2] does not work. However, we can fix this issue as follows: for any non-locally free FMGX(2,1,5)F\in M^{X}_{G}(2,1,5) determined by a line LL, we define ι(F):=F\iota(F):=F^{\prime}, where F:=ker(𝒪L)F^{\prime}:=\ker(\mathcal{E}^{\vee}\twoheadrightarrow\operatorname{\mathcal{O}}_{L^{\prime}}) is a non-locally free stable sheaf determined by the residue line LL^{\prime} of LL. This extends ι\iota to be a regular involution on MGX(2,1,5)M^{X}_{G}(2,1,5).

Note that for a special GM threefold, there is another involution on MG(2,1,5)M_{G}(2,1,5) induced by the involution τ\tau on XX,

τ:MG(2,1,5)MG(2,1,5),FτF.\tau^{*}\colon M_{G}(2,1,5)\to M_{G}(2,1,5),\,\,F\mapsto\tau^{*}F.

And it is clear that τ𝒜(pr(F))pr(τF)\tau_{\mathcal{A}}(\mathrm{pr}(F))\cong\mathrm{pr}(\tau^{*}F).

Now let XX be an ordinary GM threefold, τ𝒜\tau_{\mathcal{A}} be the involution of 𝒜X\mathcal{A}_{X}, and ι\iota be the geometric involution of MG(2,1,5)M_{G}(2,1,5) defined above. Then τ𝒜\tau_{\mathcal{A}} induces involutions of the Bridgeland moduli spaces of σ\sigma-stable objects σ(𝒜X,x)\mathcal{M}_{\sigma}(\mathcal{A}_{X},-x) and σ(𝒜X,y2x)\mathcal{M}_{\sigma}(\mathcal{A}_{X},y-2x). In Proposition 7.13, we already showed that the action of τ𝒜\tau_{\mathcal{A}} on σ(𝒜X,x)\mathcal{M}_{\sigma}(\mathcal{A}_{X},-x) induces a geometric involution on 𝒞m(X)\mathcal{C}_{m}(X). In this section, we show that the involution induced by τ𝒜\tau_{\mathcal{A}} is also compatible with ι\iota on MG(2,1,5)M_{G}(2,1,5).

Proposition 8.8.

Let XX be an ordinary GM threefold and FMGX(2,1,5)F\in M^{X}_{G}(2,1,5). Then τ𝒜pr(F)pr(ι(F))\tau_{\mathcal{A}}\mathrm{pr}(F)\cong\mathrm{pr}(\iota(F)).

Proof.
  1. (1)

    If FF is a non-locally free sheaf determined by a line LL, then by Lemma 8.4 we have the triangle

    [1]pr(F)𝒪L(1).\mathcal{E}[1]\rightarrow\mathrm{pr}(F)\rightarrow\mathcal{O}_{L^{\prime}}(-1).

    Then since τ𝒜τ𝒜1𝐋𝒪X𝐋(𝒪X(H))[1]\tau_{\mathcal{A}}\cong\tau_{\mathcal{A}}^{-1}\cong\bm{\mathrm{L}}_{\mathcal{O}_{X}}\circ\bm{\mathrm{L}}_{\mathcal{E}^{\vee}}(-\otimes\mathcal{O}_{X}(H))[-1], τ𝒜(pr(F))\tau_{\mathcal{A}}(\mathrm{pr}(F)) is given by a triangle

    𝐋𝒪X𝐋()τ𝒜(pr(F))𝐋𝒪X𝐋(𝒪L)[1].\bm{\mathrm{L}}_{\mathcal{O}_{X}}\bm{\mathrm{L}}_{\mathcal{E}^{\vee}}(\mathcal{E}^{\vee})\rightarrow\tau_{\mathcal{A}}(\mathrm{pr}(F))\rightarrow\bm{\mathrm{L}}_{\mathcal{O}_{X}}\bm{\mathrm{L}}_{\mathcal{E}^{\vee}}(\mathcal{O}_{L^{\prime}})[-1].

    Note that 𝐋()=0\bm{\mathrm{L}}_{\mathcal{E}^{\vee}}(\mathcal{E}^{\vee})=0, hence τA(pr(F))𝐋𝒪X𝐋(𝒪L)[1]\tau_{A}(\mathrm{pr}(F))\cong\bm{\mathrm{L}}_{\mathcal{O}_{X}}\bm{\mathrm{L}}_{\mathcal{E}^{\vee}}(\mathcal{O}_{L^{\prime}})[-1]. It is easy to see RHom(,𝒪L)=k\mathrm{RHom}^{\bullet}(\mathcal{E}^{\vee},\mathcal{O}_{L^{\prime}})=k, therefore we have 𝒪L𝐋𝒪L\mathcal{E}^{\vee}\rightarrow\mathcal{O}_{L^{\prime}}\rightarrow\bm{\mathrm{L}}_{\mathcal{E}^{\vee}}\mathcal{O}_{L^{\prime}}. Also, since 𝒪L\mathcal{E}^{\vee}\rightarrow\mathcal{O}_{L^{\prime}} is surjective, we have 𝐋𝒪LF[1]\bm{\mathrm{L}}_{\mathcal{E}^{\vee}}\mathcal{O}_{L^{\prime}}\cong F^{\prime}[1], where F:=ker(𝒪L)F^{\prime}:=\ker(\mathcal{E}^{\vee}\rightarrow\mathcal{O}_{L^{\prime}}) is a non-locally free sheaf in MG(2,1,5)M_{G}(2,1,5) determined by LL^{\prime} as in Proposition 8.1. Thus τ𝒜(pr(F))𝐋𝒪XFpr(F)=pr(ι(F))\tau_{\mathcal{A}}(\mathrm{pr}(F))\cong\bm{\mathrm{L}}_{\operatorname{\mathcal{O}}_{X}}F^{\prime}\cong\mathrm{pr}(F^{\prime})=\mathrm{pr}(\iota(F)).

  2. (2)

    If FF is a globally generated vector bundle, consider the standard short exact exact sequence

    0ker(ev)H0(X,F)𝒪XevF0.0\rightarrow\ker(\mathrm{ev})\rightarrow H^{0}(X,F)\otimes\mathcal{O}_{X}\xrightarrow{\mathrm{ev}}F\rightarrow 0.

    Dualizing the sequence and applying pr\mathrm{pr}, we get the triangle

    pr(F)pr(𝒪X4)pr(ker(ev))pr(ιF).\mathrm{pr}(F^{\vee})\rightarrow\mathrm{pr}(\mathcal{O}_{X}^{\oplus 4})\rightarrow\mathrm{pr}(\ker(\mathrm{ev})^{\vee})\cong\mathrm{pr}(\iota F).

    Note that F𝒜XF^{\vee}\in\mathcal{A}_{X} and pr(𝒪X)=0\mathrm{pr}(\mathcal{O}_{X})=0, thus we get pr(ιF)F[1]\mathrm{pr}(\iota F)\cong F^{\vee}[1]. Since FMG(2,1,5)F\in M_{G}(2,1,5) is a globally generated vector bundle, we have FιEF\cong\iota E for some globally generated vector bundle EE. Then pr(F)=pr(ιE)E[1]E𝒪X(H)[1]\mathrm{pr}(F)=\mathrm{pr}(\iota E)\cong E^{\vee}[1]\cong E\otimes\mathcal{O}_{X}(-H)[1], hence τ𝒜(pr(F))τ𝒜(E𝒪X(H))[1]pr(E)pr(ιF)\tau_{\mathcal{A}}(\mathrm{pr}(F))\cong\tau_{\mathcal{A}}(E\otimes\mathcal{O}_{X}(-H))[1]\cong\mathrm{pr}(E)\cong\mathrm{pr}(\iota F).

8.3. The Bridgeland moduli space of class y2xy-2x

In this subsection, we show that MG(2,1,5)σ(𝒜X,y2x)M_{G}(2,1,5)\cong\mathcal{M}_{\sigma}(\mathcal{A}_{X},y-2x).

Theorem 8.9.

Let XX be a GM threefold and σ\sigma be a Serre-invariant stability condition on 𝒜X\mathcal{A}_{X}. Then the projection functor pr:Db(X)𝒜X\mathrm{pr}\colon\mathrm{D}^{b}(X)\rightarrow\mathcal{A}_{X} induces an isomorphism MGX(2,1,5)σ(𝒜X,y2x)M^{X}_{G}(2,1,5)\cong\mathcal{M}_{\sigma}(\mathcal{A}_{X},y-2x).

We split the proof of this theorem into a series of lemmas and propositions. Recall that in 4.9 we defined

V:={(α,β):110<β<0,0<α<β}.V:=\{(\alpha,\beta)\colon-\frac{1}{10}<\beta<0,0<\alpha<-\beta\}.
Proposition 8.10.

Let F𝒜(α,β)F\in\mathcal{A}(\alpha,\beta) be a σ(α,β)\sigma(\alpha,\beta)-stable object with numerical class y2xy-2x for every (α,β)V(\alpha,\beta)\in V. Then F=pr(E)F=\mathrm{pr}(E) for some EMG(2,1,5)E\in M_{G}(2,1,5).

Proof.

First, we argue as in [52, Proposition 4.6]. When (α0,β0)=(0,0)(\alpha_{0},\beta_{0})=(0,0), we have μα0,β00(F)=\mu^{0}_{\alpha_{0},\beta_{0}}(F)=-\infty. Since there are no walls intersecting with β=0\beta=0 as in [52, Proposition 4.6], we know that FF is σα,00\sigma^{0}_{\alpha,0}-semistable for all α>0\alpha>0. By the definition of the double-tilted heart, we have a triangle

A[1]FBA[1]\to F\to B

such that AA (respectively BB) is in Coh0(X)\mathrm{Coh}^{0}(X) with its σα,0\sigma_{\alpha,0}-semistable factors having slope μα,00\mu_{\alpha,0}\leq 0 (respectively μα,0>0\mu_{\alpha,0}>0). Since FF is σα,00\sigma^{0}_{\alpha,0}-semistable and μα,00(F)<0\mu^{0}_{\alpha,0}(F)<0, we have that A[1]=0A[1]=0 and BFB\cong F. Since ch10(F)\mathrm{ch}^{0}_{1}(F) is minimal, there are no walls on β=0\beta=0, and we know that FF is σα,0\sigma_{\alpha,0}-semistable for every α>0\alpha>0. Thus by Lemma 4.5, 1(F)\mathcal{H}^{-1}(F) is a μ\mu-semistable reflexive sheaf and 0(F)\mathcal{H}^{0}(F) is 0 or supported in dimension 1\leq 1.

If 0(F)\mathcal{H}^{0}(F) is supported in dimension 0, then ch(0(F))=bP\mathrm{ch}(\mathcal{H}^{0}(F))=bP for b1b\geq 1. But this is impossible since then c3(1(F))>0c_{3}(\mathcal{H}^{-1}(F))>0 and by [4, Proposition 3.5] we have χ(1(F))=0\chi(\mathcal{H}^{-1}(F))=0, which implies b=0b=0.

If 0(F)\mathcal{H}^{0}(F) is supported in dimension 1, we can assume ch(0(F))=aL+b2P\mathrm{ch}(\mathcal{H}^{0}(F))=aL+\frac{b}{2}P where a1a\geq 1 and bb are integers. Thus ch(1(F))=2H+aL+(56+b2)P\mathrm{ch}(\mathcal{H}^{-1}(F))=2-H+aL+(\frac{5}{6}+\frac{b}{2})P. Now from Lemma 7.7, we know 1(F)\mathcal{H}^{-1}(F)\cong\mathcal{E} and ch(0(F))=LP2\mathrm{ch}(\mathcal{H}^{0}(F))=L-\frac{P}{2}. Thus 0(F)𝒪L(1)\mathcal{H}^{0}(F)\cong\mathcal{O}_{L}(-1) for some line LL on XX. Therefore we have a triangle

[1]F𝒪L(1).\mathcal{E}[1]\to F\to\mathcal{O}_{L}(-1).

In this case we have Hom(𝒪L(1),[2])=Hom((1),𝒪L[1])=H1(L,(1)|L)=H1(L,𝒪L(1)𝒪L(2))=k\mathrm{Hom}(\mathcal{O}_{L}(-1),\mathcal{E}[2])=\mathrm{Hom}(\mathcal{E}^{\vee}(1),\mathcal{O}_{L}[1])=H^{1}(L,\mathcal{E}(-1)|_{L})=H^{1}(L,\mathcal{O}_{L}(-1)\oplus\mathcal{O}_{L}(-2))=k. Hence by Lemma 8.4, Fpr(E)F\cong\mathrm{pr}(E) for some EMG(2,1,5)E\in M_{G}(2,1,5) such that EE is locally free but not globally generated.

If 0(F)=0\mathcal{H}^{0}(F)=0, we have F[1]1(F)F[-1]\cong\mathcal{H}^{-1}(F). Then F[1]F[-1] is a μ\mu-semistable sheaf. Since F[1]F[-1] is reflexive and c3(F[1])=0c_{3}(F[-1])=0, F[1]MG(2,1,5)F[-1]\in M_{G}(2,-1,5) is a stable vector bundle. Thus by Lemma 8.4, we know F[1]=pr(E)F[-1]=\mathrm{pr}(E) for some EMG(2,1,5)E\in M_{G}(2,1,5) such that EE is a globally generated vector bundle. ∎

Lemma 8.11.

The functor pr:Db(X)𝒜X\mathrm{pr}\colon\mathrm{D}^{b}(X)\rightarrow\mathcal{A}_{X} is injective on all objects in MG(2,1,5)M_{G}(2,1,5), i.e. if pr(F1)pr(F2)\mathrm{pr}(F_{1})\cong\mathrm{pr}(F_{2}), then F1F2F_{1}\cong F_{2}.

Proof.

For the case of globally generated vector bundles, by Corollary 8.4, pr(F1)pr(F2)\mathrm{pr}(F_{1})\cong\mathrm{pr}(F_{2}) implies that

(ιF1)(ιF2).(\iota F_{1})^{\vee}\cong(\iota F_{2})^{\vee}.

Note that (ιFi)ιFi𝒪X(H)(\iota F_{i})^{\vee}\cong\iota F_{i}\otimes\mathcal{O}_{X}(-H) for i=1,2i=1,2. Then we get ιF1ιF2\iota F_{1}\cong\iota F_{2}. Finally, we apply ι\iota to both sides. Since it is an involution ι2=id\iota^{2}=\mathrm{id}, so F1F2F_{1}\cong F_{2} as required.

For the case of non-locally free sheaves FF, recall that from Lemma 8.4 we have 1(pr(F))=\mathcal{H}^{-1}(\mathrm{pr}(F))=\mathcal{E} and 0(pr(F))=𝒪L(H)\mathcal{H}^{0}(\mathrm{pr}(F))=\operatorname{\mathcal{O}}_{L}(-H). Since FF is uniquely determined by the line LL, and Hom(𝒪L(H),[2])=k\operatorname{Hom}(\operatorname{\mathcal{O}}_{L}(-H),\mathcal{E}[2])=k, the object pr(F)\mathrm{pr}(F) is also uniquely determined by the line LL. Thus pr(F1)pr(F2)\mathrm{pr}(F_{1})\cong\mathrm{pr}(F_{2}) implies F1F2F_{1}\cong F_{2}, as required. ∎

Proof of Theorem 8.9.

Using Proposition 8.7, we know that the projection functor pr\mathrm{pr} induces a morphism

p:MG(2,1,5)σ(𝒜X,y2x)p\colon M_{G}(2,1,5)\to\mathcal{M}_{\sigma}(\mathcal{A}_{X},y-2x)

which is bijective on points by Proposition 8.10 and Lemma 8.11, and bijective on tangent spaces by Lemma 8.5. Hence it is an isomorphism. ∎

9. Refined and birational categorical Torelli theorems for GM threefolds

In this section, we will prove several refined/birational categorical Torelli theorems for GM threefolds, using results from the previous sections.

9.1. The universal family for 𝒞m(X)\mathcal{C}_{m}(X)

In this subsection, we show that 𝒞m(X)σ(𝒜X,x)\mathcal{C}_{m}(X)\cong\mathcal{M}_{\sigma}(\mathcal{A}_{X},-x) admits a universal family, which thus gives a fine moduli space. Let \mathcal{I} be the universal ideal sheaf of conics on X×𝒞(X)X\times\mathcal{C}(X) and Lσ\mathcal{I}_{L_{\sigma}} be the universal ideal sheaf of conics restricted to X×LσX\times L_{\sigma}. Let q:X×𝒞(X)Xq\colon X\times\mathcal{C}(X)\rightarrow X and π:X×𝒞(X)𝒞(X)\pi\colon X\times\mathcal{C}(X)\rightarrow\mathcal{C}(X) be the projection maps on the first and second factors, respectively. Let 𝒢:=pr(Lσ)\mathcal{G}^{\prime}:=\mathrm{pr}(\mathcal{I}_{L_{\sigma}}) be the projected family in 𝒜X×Lσ\mathcal{A}_{X\times L_{\sigma}}. Let tLσ1t\in L_{\sigma}\cong\mathbb{P}^{1} be any point. Then jtpr(Lσ)Aj_{t}^{*}\mathrm{pr}(\mathcal{I}_{L_{\sigma}})\cong A, where jt:XtXt×Lσj_{t}\colon X_{t}\rightarrow X_{t}\times L_{\sigma} and A𝒜XA\in\mathcal{A}_{X} is Apr(IC)A\cong\mathrm{pr}(I_{C}) for IC𝒜XI_{C}\notin\mathcal{A}_{X} by Proposition 7.2. Then 𝒢q(A)π𝒪Lσ(k)\mathcal{G}^{\prime}\cong q^{*}(A)\otimes\pi^{*}\mathcal{O}_{L_{\sigma}}(k) for some kk\in\mathbb{Z}. Now let 𝒢:=pr()π𝒪𝒞(X)(kE)\mathcal{G}:=\mathrm{pr}(\mathcal{I})\otimes\pi^{*}\mathcal{O}_{\mathcal{C}(X)}(kE), where ELσ1E\cong L_{\sigma}\cong\mathbb{P}^{1} is the unique exceptional curve on 𝒞(X)\mathcal{C}(X).

Proposition 9.1.

The object (pX)𝒢(p_{X})_{*}\mathcal{G} is the universal family of 𝒞m(X)\mathcal{C}_{m}(X), where pX=idX×p:X×𝒞(X)X×𝒞m(X)p_{X}=\mathrm{id}_{X}\times p\colon X\times\mathcal{C}(X)\rightarrow X\times\mathcal{C}_{m}(X).

Proof.
  1. (1)

    If s=[A]=π𝒞m(X)s=[A]=\pi\in\mathcal{C}_{m}(X), ss is contracted from the unique rational curve Lσ1𝒞(X)L_{\sigma}\cong\mathbb{P}^{1}\subset\mathcal{C}(X). Note that in this case pX|Lσ=qp_{X}|_{L_{\sigma}}=q. Then

    is(pX)𝒢\displaystyle i_{s}^{*}(p_{X})_{*}\mathcal{G} is(pX)(𝒢π𝒪𝒞(X)(kE))\displaystyle\cong i_{s}^{*}(p_{X})_{*}(\mathcal{G}^{\prime}\otimes\pi^{*}\mathcal{O}_{\mathcal{C}(X)}(kE))
    isq(q(A)π𝒪Lσ(k)π𝒪𝒞(X)(kE))\displaystyle\cong i_{s}^{*}q_{*}(q^{*}(A)\otimes\pi^{*}\mathcal{O}_{L_{\sigma}}(k)\otimes\pi^{*}\mathcal{O}_{\mathcal{C}(X)}(kE))
    isq(q(A)(π𝒪Lσ(k)𝒪Lσ(kE)))\displaystyle\cong i_{s}^{*}q_{*}(q^{*}(A)\otimes(\pi^{*}\mathcal{O}_{L_{\sigma}}(k)\otimes\mathcal{O}_{L_{\sigma}}(kE)))
    isq(q(A)π(𝒪Lσ(k)𝒪Lσ(k)))\displaystyle\cong i_{s}^{*}q_{*}(q^{*}(A)\otimes\pi^{*}(\mathcal{O}_{L_{\sigma}}(k)\otimes\mathcal{O}_{L_{\sigma}}(-k)))
    isq(q(A))is(A)A.\displaystyle\cong i_{s}^{*}q_{*}(q^{*}(A))\cong i_{s}^{*}(A)\cong A.
  2. (2)

    If s=[IC]s=[I_{C}], then 𝒞m(X)\mathcal{C}_{m}(X) and 𝒞(X)\mathcal{C}(X) are isomorphic outside LσL_{\sigma}. Note that pp restricts to id\mathrm{id} on 𝒞(X)Lσ\mathcal{C}(X)\smallsetminus L_{\sigma}. Then

    is(pX)𝒢\displaystyle i_{s}^{*}(p_{X})_{*}\mathcal{G} is(pX)(pr()π𝒪𝒞(X)(kE))\displaystyle\cong i_{s}^{*}(p_{X})_{*}(\mathrm{pr}(\mathcal{I})\otimes\pi^{*}\mathcal{O}_{\mathcal{C}(X)}(kE))
    js(pr())jsπ𝒪𝒞(X)(kE)\displaystyle\cong j_{s}^{*}(\mathrm{pr}(\mathcal{I}))\otimes j_{s}^{*}\pi^{*}\mathcal{O}_{\mathcal{C}(X)}(kE)
    IC(πjs)𝒪𝒞(X)(kE)\displaystyle\cong I_{C}\otimes(\pi\circ j_{s})^{*}\mathcal{O}_{\mathcal{C}(X)}(kE)
    IC(isπs)𝒪𝒞(X)(kE)IC.\displaystyle\cong I_{C}\otimes(i_{s}\circ\pi_{s})^{*}\mathcal{O}_{\mathcal{C}(X)}(kE)\cong I_{C}.

See below for the commutative diagrams which summarise the maps in the proof:

Xs{X_{s}}Xs{X_{s}}{s}{\{s\}}X×𝒞(X){X\times\mathcal{C}(X)}X×𝒞m(X){X\times\mathcal{C}_{m}(X)}𝒞m(X){\mathcal{C}_{m}(X)}\scriptstyle{\cong}js\scriptstyle{j_{s}}πs\scriptstyle{\pi_{s}}is\scriptstyle{i_{s}}pX\scriptstyle{p_{X}}
Xs{X_{s}}X×𝒞(X){X\times\mathcal{C}(X)}{s}{\{s\}}𝒞(X){\mathcal{C}(X)}js\scriptstyle{j_{s}}πs\scriptstyle{\pi_{s}}π\scriptstyle{\pi}is\scriptstyle{i_{s}}

9.2. A refined categorical Torelli theorem for ordinary GM threefolds

We now prove a refined categorical Torelli theorem for ordinary GM threefolds.

Theorem 9.2.

Let XX and XX^{\prime} be general ordinary GM threefolds such that Φ:𝒦u(X)𝒦u(X)\Phi\colon\mathcal{K}u(X)\simeq\mathcal{K}u(X^{\prime}) is an equivalence and Φ(π())π()\Phi(\pi(\mathcal{E}))\cong\pi(\mathcal{E}^{\prime}). Then XXX\cong X^{\prime}.

Proof.

Since Φ\Phi commutes with Serre functors, it preserves the stability of an object with respect to any Serre-invariant stability condition. Then the existence of the universal family on 𝒞m(X)σ(𝒜X,x)\mathcal{C}_{m}(X)\cong\mathcal{M}_{\sigma}(\mathcal{A}_{X},-x) guarantees a morphism from 𝒞m(X)\mathcal{C}_{m}(X) to 𝒞m(X)\mathcal{C}_{m}(X^{\prime}), denoted by ψ\psi, which is induced by Ψ\Psi (for more details on the construction of the morphism ψ\psi, see [7, 1]). Since Φ\Phi is an equivalence, ψ\psi is an isomorphism. On the other hand, we have ψ([πX])=[πX]\psi([\pi_{X}])=[\pi_{X^{\prime}}] by the assumption, where πX:=π(𝒬)\pi_{X}:=\pi^{\prime}(\mathcal{Q}^{\vee}) and πX:=π(𝒬)\pi_{X^{\prime}}:=\pi^{\prime}(\mathcal{Q}^{\vee}). Then ψ\psi induces an isomorphism ϕ:𝒞(X)𝒞(X)\phi\colon\mathcal{C}(X)\cong\mathcal{C}(X^{\prime}) by blowing up [πX]𝒞m(X)[\pi_{X}]\in\mathcal{C}_{m}(X) and [πX]𝒞m(X)[\pi_{X^{\prime}}]\in\mathcal{C}_{m}(X^{\prime}), respectively. Then we have XXX\cong X^{\prime} by Logachev’s Reconstruction Theorem 6.7. ∎

9.3. Birational categorical Torelli theorem for ordinary GM threefolds

In this subsection, we show a birational categorical Torelli theorem for ordinary GM threefolds, i.e. assuming the Kuznetsov components are equivalent leads to a birational equivalence of the ordinary GM threefolds.

Theorem 9.3.

Let XX and XX^{\prime} be general ordinary GM threefolds such that 𝒜X𝒜X\mathcal{A}_{X}\simeq\mathcal{A}_{X^{\prime}}. Then XX^{\prime} is a conic transform or a conic transform of a line transform of XX. In particular, we have XXX\simeq X^{\prime}.

Proof.

The equivalence Φ:𝒜X𝒜X\Phi\colon\mathcal{A}_{X}\xrightarrow{\sim}\mathcal{A}_{X^{\prime}} sends x-x to either itself or y2xy-2x in 𝒩(𝒜X)\mathcal{N}(\mathcal{A}_{X^{\prime}}) up to sign, since they are only (1)(-1)-class. By the same argument as in Theorem 9.2 and [7, 1], we thus get two possible induced isomorphisms between Bridgeland moduli spaces

𝒞m(X)σ(𝒜X,x){{\mathcal{C}_{m}(X)\cong\mathcal{M}_{\sigma}(\mathcal{A}_{X},-x)}}𝒞m(X)Φ(σ)(𝒜X,x){{\mathcal{C}_{m}(X^{\prime})\cong\mathcal{M}_{\Phi(\sigma)}(\mathcal{A}_{X^{\prime}},-x)}}MGX(2,1,5)Φ(σ)(𝒜X,y2x){{M_{G}^{X^{\prime}}(2,1,5)\cong\mathcal{M}_{\Phi(\sigma)}(\mathcal{A}_{X^{\prime}},y-2x)}}γ\scriptstyle{\gamma}γ\scriptstyle{\gamma^{\prime}}

If we have the isomorphism γ\gamma, then we blow up 𝒞m(X)\mathcal{C}_{m}(X) at the distinguished point [πX]:=[Ξ(π())][\pi_{X}]:=[\Xi(\pi(\mathcal{E}))], and blow up 𝒞m(X)\mathcal{C}_{m}(X^{\prime}) at the point [C]:=[Φ(πX)]=γ([πX])[C]:=[\Phi(\pi_{X})]=\gamma([\pi_{X}]). We have

𝒞(X)Bl[πX]𝒞m(X)Bl[C]𝒞m(X),\mathcal{C}(X)\cong\mathrm{Bl}_{[\pi_{X}]}\mathcal{C}_{m}(X)\cong\mathrm{Bl}_{[C]}\mathcal{C}_{m}(X^{\prime}),

and Bl[C]𝒞m(X)𝒞(XC)\mathrm{Bl}_{[C]}\mathcal{C}_{m}(X^{\prime})\cong\mathcal{C}(X^{\prime}_{C}) by Theorem 6.8, so 𝒞(X)𝒞(XC)\mathcal{C}(X)\cong\mathcal{C}(X_{C}^{\prime}). Therefore by Logachev’s Reconstruction Theorem 6.7 we have XXCX\cong X^{\prime}_{C}.

For the second case, we get 𝒞m(X)MGX(2,1,5)\mathcal{C}_{m}(X)\cong M_{G}^{X^{\prime}}(2,1,5). And by [13, Proposition 8.1] we have a birational equivalence MGX(2,1,5)𝒞(XL)M_{G}^{X^{\prime}}(2,1,5)\simeq\mathcal{C}(X_{L}^{\prime}) of surfaces, where LXL\subset X^{\prime} is a line. Then we see 𝒞m(X)\mathcal{C}_{m}(X) is birationally equivalent to 𝒞(XL)\mathcal{C}(X^{\prime}_{L}). Let 𝒞m(XL)\mathcal{C}_{m}(X^{\prime}_{L}) be the minimal surface of 𝒞(XL)\mathcal{C}(X^{\prime}_{L}). Note that the surfaces here are all smooth surfaces of general type. By the uniqueness of minimal models of surfaces of general type, we get 𝒞m(X)𝒞m(XL)\mathcal{C}_{m}(X)\cong\mathcal{C}_{m}(X^{\prime}_{L}), which implies X(XL)CXX\cong(X^{\prime}_{L})_{C}\simeq X^{\prime} for a conic CXLC\subset X^{\prime}_{L} as in the first case. ∎

Remark 9.4.

Theorem 9.3 proves a conjecture [30, Conjecture 1.7] of Kuznetsov–Perry for general ordinary GM varieties of dimension 33.

In [13], the authors proved that 𝒞m(XL)\mathcal{C}_{m}(X_{L}) is birational to MGX(2,1,5)M^{X}_{G}(2,1,5). The following corollary shows that they are indeed isomorphic.

Corollary 9.5.

Let XX be a general ordinary GM threefold, and XLX_{L} be a line transform of XX. Then we have 𝒞m(XL)MGX(2,1,5)\mathcal{C}_{m}(X_{L})\cong M^{X}_{G}(2,1,5). Moreover, this isomorphism commutes with involutions ι\iota and ι\iota^{\prime} on both sides, thus giving an isomorphism 𝒞m(XL)/ιMGX(2,1,5)/ι\mathcal{C}_{m}(X_{L})/\iota\cong M^{X}_{G}(2,1,5)/\iota^{\prime}.

Proof.

By the same argument as in the proof of Theorem 9.3, we have 𝒞m(XL)𝒞m(X)\mathcal{C}_{m}(X_{L})\cong\mathcal{C}_{m}(X) or 𝒞m(XL)MGX(2,1,5)\mathcal{C}_{m}(X_{L})\cong M^{X}_{G}(2,1,5). Note that 𝒞m(XL)𝒞m(X)\mathcal{C}_{m}(X_{L})\cong\mathcal{C}_{m}(X) implies that XLXCX_{L}\cong X_{C} for some conic CXC\subset X as in Theorem 9.3. But this is impossible by [13, Remark 7.3]. Thus we always have 𝒞m(XL)MGX(2,1,5)\mathcal{C}_{m}(X_{L})\cong M^{X}_{G}(2,1,5). The last statement follows from the fact that any equivalence between Kuznetsov components commutes with Serre functors, and the involutions on 𝒞m(XL)\mathcal{C}_{m}(X_{L}) and MGX(2,1,5)M^{X}_{G}(2,1,5) can be induced by Serre functors up to shift by Propositions 7.13 and Proposition 8.8. ∎

Since the intermediate Jacobian J(X)J(X) is invariant under conic and line transforms, as a corollary we have

Corollary 9.6.

Let XX and XX^{\prime} be general ordinary GM threefolds. If 𝒦u(X)𝒦u(X)\mathcal{K}u(X)\simeq\mathcal{K}u(X^{\prime}), then we have J(X)J(X)J(X)\cong J(X^{\prime}).

In fact, we can relax the assumptions on XX by looking at the singularities of Bridgeland moduli spaces.

Theorem 9.7.

Let XX and XX^{\prime} be general GM threefolds (they can be either general ordinary or general special) and suppose their Kuznetsov components 𝒜X𝒜X\mathcal{A}_{X}\simeq\mathcal{A}_{X^{\prime}} are equivalent. Then XX is birationally equivalent to XX^{\prime}.

Proof.

First, we claim that if XX and XX^{\prime} are general GM threefolds such that Φ:𝒜X𝒜X\Phi\colon\mathcal{A}_{X}\simeq\mathcal{A}_{X^{\prime}}, then both XX and XX^{\prime} are ordinary or special simultaneously. Indeed, we may assume XX^{\prime} is ordinary and XX is special. Then the equivalence would identify the moduli space 𝒞m(X)σ(𝒜X,x)\mathcal{C}_{m}(X)\cong\mathcal{M}_{\sigma}(\mathcal{A}_{X},-x) of stable objects of class x-x in 𝒜X\mathcal{A}_{X} with either the moduli space 𝒞m(X)σ(𝒜X,x)\mathcal{C}_{m}(X^{\prime})\cong\mathcal{M}_{\sigma^{\prime}}(\mathcal{A}_{X^{\prime}},-x) or MGX(2,1,5)σ(𝒜X,y2x)M_{G}^{X^{\prime}}(2,1,5)\cong\mathcal{M}_{\sigma^{\prime}}(\mathcal{A}_{X^{\prime}},y-2x). But 𝒞m(X)\mathcal{C}_{m}(X) has a unique singular point by Theorem 7.12, and both 𝒞m(X)\mathcal{C}_{m}(X^{\prime}) and MGX(2,1,5)M_{G}^{X^{\prime}}(2,1,5) are smooth for XX^{\prime} general by Theorem 7.12 and Theorem 8.9. This means that neither identification is possible, so the claim follows.

Now XX and XX^{\prime} are both general ordinary or general special, hence the result follows from Theorem 9.3 and 9.4. ∎

Corollary 9.8.

Let XX and XX^{\prime} be general GM threefolds such that one of them is ordinary and their Kuznetsov components 𝒜X𝒜X\mathcal{A}_{X}\simeq\mathcal{A}_{X^{\prime}} are equivalent. Then they are both general ordinary and XX is birationally equivalent to XX^{\prime}.

9.4. A categorical Torelli theorem for special GM threefolds

In this subsection, we show that the Kuznetsov component of a general special GM threefold XX determines the isomorphism class of XX.

Recall from Section 3 that every special GM threefold XX is a double cover of a degree 55 index 22 prime Fano threefold YY branched over a quadric hypersurface \mathcal{B} in YY. Since XX is smooth and general, (,h)(\mathcal{B},h) is a smooth degree h2=10h^{2}=10 K3 surface with Picard number 11. There is a natural geometric involution τ\tau on XX induced by the double cover. The Serre functor on 𝒦u(X)\mathcal{K}u(X) is given by S𝒦u(X)=τ[2]S_{\mathcal{K}u(X)}=\tau\circ[2].

Theorem 9.9.

Let XX and XX^{\prime} be general special GM threefolds with Φ:𝒦u(X)𝒦u(X)\Phi\colon\mathcal{K}u(X)\simeq\mathcal{K}u(X^{\prime}). Then XXX\cong X^{\prime}.

Proof.

By [28, Theorem 1.1, Section 8.2], the equivariant triangulated category 𝒦u(X)μ2\mathcal{K}u(X)^{\mu_{2}} is equivalent to Db()\mathrm{D}^{b}(\mathcal{B}), where μ2\mu_{2} is the group of square roots of 11 generated by the involution τ\tau acting on 𝒦u(X)\mathcal{K}u(X). Assume there is an equivalence Φ:𝒦u(X)𝒦u(X)\Phi\colon\mathcal{K}u(X)\simeq\mathcal{K}u(X^{\prime}). Since S𝒦u(X)τ[2]S_{\mathcal{K}u(X)}\cong\tau[2] and S𝒦u(X)τ[2]S_{\mathcal{K}u(X^{\prime})}\cong\tau^{\prime}[2], Φ\Phi commutes with the involutions τ\tau and τ\tau^{\prime} on 𝒦u(X)\mathcal{K}u(X) and 𝒦u(X)\mathcal{K}u(X^{\prime}), respectively. Then we get an induced equivalence

Ψ:𝒦u(X)μ2𝒦u(X)μ2\Psi:\mathcal{K}u(X)^{\mu_{2}}\simeq\mathcal{K}u(X^{\prime})^{\mu_{2}^{\prime}}

where μ2=τ\mu_{2}=\langle\tau\rangle, μ2=ΦτΦ1=τ\mu_{2}^{\prime}=\langle\Phi\circ\tau\circ\Phi^{-1}=\tau^{\prime}\rangle and μ2μ2\mu_{2}\cong\mu_{2}^{\prime}. Thus we have Ψ:Db()Db()\Psi\colon\mathrm{D}^{b}(\mathcal{B})\simeq\mathrm{D}^{b}(\mathcal{B}^{\prime}). We know that \mathcal{B} and \mathcal{B}^{\prime} are smooth projective surfaces with polarizations hh and hh^{\prime}, respectively, so Ψ\Psi is a Fourier–Mukai functor by Orlov’s Representability Theorem [45, Theorem 2.2]. Moreover, (,h)(\mathcal{B},h) and (,h)(\mathcal{B}^{\prime},h^{\prime}) are both Picard number 1 smooth projective K3 surfaces of degree h2=h2=10=2×5h^{2}=h^{\prime 2}=10=2\times 5. Then by [44, Theorem 1.10] and [19, Corollary 1.7], there is an isomorphism ϕ:\phi\colon\mathcal{B}\cong\mathcal{B}^{\prime}. Since they both have Picard number one, we obtain ϕ(h)=h\phi^{*}(h^{\prime})=h. On the other hand Y5Y_{5} is rigid [33, § 4.1], which implies XXX\cong X^{\prime}. ∎

Remark 9.10.

Theorem 9.9 can also be proved via Bridgeland moduli spaces with respect to the Kuznetsov component 𝒜X\mathcal{A}_{X}. The details are contained in another paper of authors [26].

10. The Debarre–Iliev–Manivel conjecture

Let 𝒳10\mathcal{X}_{10} be the moduli space of smooth ordinary GM threefolds and 𝒜10\mathcal{A}_{10} be the moduli space of 1010-dimensional principal polarised abelian varieties. In [13, pp. 3-4], the authors make the following conjecture regarding the general fiber of the period map:

Conjecture 10.1 ([13, pp. 3-4]).

A general fiber 𝒫1([J(X)])\mathcal{P}^{-1}([J(X)]) of the period map 𝒫:𝒳10𝒜10\mathcal{P}\colon\mathcal{X}_{10}\rightarrow\mathcal{A}_{10} at the intermediate Jacobian J(X)J(X) of an ordinary GM threefold XX is the union of 𝒞m(X)/ι\mathcal{C}_{m}(X)/\iota and a surface birationally equivalent to MGX(2,1,5)/ιM^{X}_{G}(2,1,5)/\iota^{\prime}, where ι,ι\iota,\iota^{\prime} are geometrically meaningful involutions.

Remark 10.2.

Note that by Corollary 9.5, the surface birationally equivalent to MG(2,1,5)/ιM_{G}(2,1,5)/\iota^{\prime} in [13], parametrizing conic transforms of a line transform of XX, is actually isomorphic to MG(2,1,5)/ιM_{G}(2,1,5)/\iota^{\prime}. Thus this conjecture predicts that a general fiber 𝒫1([J(X)])\mathcal{P}^{-1}([J(X)]) is actually the disjoint union of 𝒞m(X)/ι\mathcal{C}_{m}(X)/\iota and MG(2,1,5)/ιM_{G}(2,1,5)/\iota^{\prime}.

We will prove a categorical analogue of this conjecture. Consider the “categorical period map”

𝒫cat:𝒳10{𝒜X}/,X𝒜X\mathcal{P}_{\mathrm{cat}}\colon\mathcal{X}_{10}\rightarrow\{\mathcal{A}_{X}\}/\simeq,\,\,X\mapsto\mathcal{A}_{X}

where 𝒳10\mathcal{X}_{10} is the moduli space of isomorphism classes of GM threefolds and {𝒜X}/\{\mathcal{A}_{X}\}/\simeq is the set of equivalence classes of Kuznetsov components of GM threefolds. Note that a global description of a “moduli of Kuznetsov components” {𝒜X}/\{\mathcal{A}_{X}\}/\simeq is not known, however, local deformations are controlled by the second Hochschild cohomology HH2(𝒜X)\mathrm{HH}^{2}(\mathcal{A}_{X}). The fiber of the “categorical period map” 𝒫cat\mathcal{P}_{\mathrm{cat}} over 𝒜X\mathcal{A}_{X} for an ordinary GM threefold is defined as the isomorphism classes of all ordinary GM threefolds XX^{\prime} such that 𝒜X𝒜X\mathcal{A}_{X^{\prime}}\simeq\mathcal{A}_{X}.

Theorem 10.3.

The general fiber 𝒫cat1([𝒜X])\mathcal{P}^{-1}_{\mathrm{cat}}([\mathcal{A}_{X}]) of the categorical period map over the alternative Kuznetsov component of an ordinary GM threefold XX is the union of 𝒞m(X)/ι\mathcal{C}_{m}(X)/\iota and MGX(2,1,5)/ιM_{G}^{X}(2,1,5)/\iota^{\prime} where ι,ι\iota,\iota^{\prime} are geometrically meaningful involutions.

Proof.

The general fiber 𝒫cat1([𝒜X])\mathcal{P}^{-1}_{\mathrm{cat}}([\mathcal{A}_{X}]) of the categorical period map consists of GM threefolds XX^{\prime} such that there is an equivalence of Kuznetsov components 𝒜X𝒜X\mathcal{A}_{X^{\prime}}\simeq\mathcal{A}_{X}. Then by Theorem 9.7, XX^{\prime} is also a general ordinary GM threefold. Thus by Theorem 9.3 and Theorem 6.10, we know that 𝒜X𝒜X\mathcal{A}_{X^{\prime}}\simeq\mathcal{A}_{X} if and only if XX^{\prime} is a conic transform of XX, or a conic transform of a line transform of XX. Then the result follows from Proposition 6.9 and Corollary 9.5. ∎

The Kuznetsov components of prime Fano threefolds of index 11 and 22 are often regarded as categorical analogues of the intermediate Jacobians of these threefolds, and it is known that if there is a Fourier–Mukai type equivalence 𝒦u(X)𝒦u(X)\mathcal{K}u(X)\simeq\mathcal{K}u(X^{\prime}) (or 𝒜X𝒜X\mathcal{A}_{X}\simeq\mathcal{A}_{X^{\prime}}), then J(X)J(X)J(X)\cong J(X^{\prime}) by [46]. For the converse, we have the following result.

Theorem 10.4.

For smooth prime Fano threefolds XX, if XX is one of the following:

  • Yd,    2d5Y_{d},\quad\quad\,\,\,\,2\leq d\leq 5

  • X2g2,g=5,7,8,9,10,12X_{2g-2},\quad g=5,7,8,9,10,12,

then the intermediate Jacobian J(X)J(X) uniquely determines the Kuznetsov component 𝒦u(X)\mathcal{K}u(X), i.e. for another prime Fano threefold XX^{\prime} of the same degree, if J(X)J(X)J(X)\cong J(X^{\prime}), then 𝒦u(X)𝒦u(X)\mathcal{K}u(X)\simeq\mathcal{K}u(X^{\prime}).

Proof.

If XX is an index 22 prime Fano threefold YdY_{d} of degree 2d52\leq d\leq 5, then the statement follows from the Torelli theorems for YdY_{d}. Now let XdX_{d} be a degree dd index one prime Fano threefold. If X=X8X=X_{8}, the statement follows from its Torelli theorem. If X=X12,X18,X16X=X_{12},X_{18},X_{16}, their intermediate Jacobians are Jacobians of curves: J(X12)J(C7)J(X_{12})\cong J(C_{7}), J(X16)J(C3)J(X_{16})\cong J(C_{3}), and J(X18)J(C2)J(X_{18})\cong J(C_{2}). But 𝒦u(X12)Db(C7)\mathcal{K}u(X_{12})\simeq\mathrm{D}^{b}(C_{7}), 𝒦u(X16)Db(C3)\mathcal{K}u(X_{16})\simeq\mathrm{D}^{b}(C_{3}) and 𝒦u(X18)Db(C2)\mathcal{K}u(X_{18})\simeq\mathrm{D}^{b}(C_{2}). Thus the statement follows from the classical Torelli theorem for curves. If X=X14X=X_{14}, the statement follows from the Kuznetsov conjecture for the pair (Y3,X14)(Y_{3},X_{14}) [32] and the Torelli theorem for cubic threefolds. If X=X22X=X_{22}, the statement is trivial since 𝒦u(X22)𝒦u(Y5)\mathcal{K}u(X_{22})\cong\mathcal{K}u(Y_{5}) ([31]) and Y5Y_{5} is rigid, so 𝒦u(X)𝒦u(X)\mathcal{K}u(X)\simeq\mathcal{K}u(X^{\prime}) is always true. ∎

Therefore, it is natural to make the following conjecture:

Conjecture 10.5.

Let XX be a prime Fano threefold of index one or two. Then the intermediate Jacobian J(X)J(X) uniquely determines the Kuznetsov component 𝒦u(X)\mathcal{K}u(X), i.e. for another prime Fano threefold XX^{\prime} of the same degree, if J(X)J(X)J(X)\cong J(X^{\prime}), then 𝒦u(X)𝒦u(X)\mathcal{K}u(X)\simeq\mathcal{K}u(X^{\prime}).

Surprisingly, in the case of general ordinary GM threefolds, we can restate the Debarre–Iliev–Manivel Conjecture 10.1 as Conjecture 10.5:

Proposition 10.6.

The Debarre–Iliev–Manivel Conjecture 10.1 is equivalent to Conjecture 10.5 for general ordinary GM threefolds.

Proof.

First, we assume that Conjecture 10.5 holds. Then by Corollary 9.6 and Theorem 10.3, the Debarre–Iliev–Manivel Conjecture 10.1 holds.

On the other hand, we assume the Debarre–Iliev–Manivel Conjecture 10.1 holds. Then for any XX and XX^{\prime} such that J(X)J(X)J(X)\cong J(X^{\prime}), XX is either a conic transform of XX^{\prime}, or XX is a conic transform of a line transform of XX^{\prime}. In both cases, we have 𝒦u(X)𝒦u(X)\mathcal{K}u(X)\simeq\mathcal{K}u(X^{\prime}) by the Duality Conjecture Theorem 6.10. Thus Conjecture 10.5 holds. ∎

Appendix A Uniqueness of Serre-invariant stability conditions

In this appendix, we aim to prove the uniqueness of Serre-invariant stability conditions on 𝒦u(X)\mathcal{K}u(X) for several prime Fano threefolds XX (Theorem A.10). We start with a general criterion for when two numerical stability conditions with the same central charge are equal. We always assume that any triangulated category is kk-linear and of finite type, i.e. iexti(A,B)<+\sum_{i}\operatorname{ext}^{i}(A,B)<+\infty for any two objects A,BA,B. Therefore, the Euler form and the numerical Grothendieck group are well-defined.

Theorem A.1.

Let 𝒟\mathcal{D} be a kk-linear triangulated category of finite type. Assume that

  1. (A).

    χ(x,x)1n\chi(x,x)\leq 1-n for a positive integer nn and any non-zero x𝒩(𝒟)x\in\mathcal{N}(\mathcal{D}),

  2. (B).

    there exists an object object DD satisfies

    next1(D,D)<2n,n\leq\operatorname{ext}^{1}(D,D)<2n,

Let σ1=(𝒜1,Z)\sigma_{1}=(\mathcal{A}_{1},Z) be a numerical stability condition on 𝒟\mathcal{D} and D1,D2𝒜1D_{1},D_{2}\in\mathcal{A}_{1} be two σ1\sigma_{1}-stable objects satisfying:

  1. (C).

    for any two objects A,B𝒟A,B\in\mathcal{D}, if ϕσ1+(B)<ϕσ1(A)\phi^{+}_{\sigma_{1}}(B)<\phi^{-}_{\sigma_{1}}(A), then Hom(B,A[2])=0\operatorname{Hom}(B,A[2])=0,

  1. (D).

    if EE is a σ1\sigma_{1}-semistable object with χ(E,D1)0\chi(E,D_{1})\geq 0 and χ(E,D2)0\chi(E,D_{2})\geq 0, then there exist k{1,2}k\in\{1,2\} such that χ(E,Dk)>0\chi(E,D_{k})>0 and μσ1(E)<μσ1(Dk)\mu_{\sigma_{1}}(E)<\mu_{\sigma_{1}}(D_{k}), and

  1. (E).

    if EE is a σ1\sigma_{1}-semistable object with χ(D1,E)0\chi(D_{1},E)\geq 0 and χ(D2,E)0\chi(D_{2},E)\geq 0, then there exist k{1,2}k\in\{1,2\} such that χ(Dk,E)>0\chi(D_{k},E)>0 and μσ1(E)>μσ1(Dk)\mu_{\sigma_{1}}(E)>\mu_{\sigma_{1}}(D_{k}).

If σ2=(𝒜2,Z)\sigma_{2}=(\mathcal{A}_{2},Z) is a numerical stability condition on 𝒟\mathcal{D} satisfies (C), (D) and (E) such that D1D_{1} and D2D_{2} are σ2\sigma_{2}-stable with ϕσ2(D1)=ϕσ1(D1)\phi_{\sigma_{2}}(D_{1})=\phi_{\sigma_{1}}(D_{1}) and ϕσ2(D2)=ϕσ1(D2)\phi_{\sigma_{2}}(D_{2})=\phi_{\sigma_{1}}(D_{2}), then σ1=σ2\sigma_{1}=\sigma_{2}.

We first prove several lemmas. By the same proof as in [2, Lemma 2.5], we have the following generalized version of Weak Mukai Lemma:

Lemma A.2.

Let 𝒟\mathcal{D} be a kk-linear triangulated category with finite-dimensional Hom\operatorname{Hom}-space. Then for any exact triangle AEBA\to E\to B with Hom(A,B)=Hom(B,A[2])=0\operatorname{Hom}(A,B)=\operatorname{Hom}(B,A[2])=0, we have

ext1(A,A)+ext1(B,B)ext1(E,E).\operatorname{ext}^{1}(A,A)+\operatorname{ext}^{1}(B,B)\leq\operatorname{ext}^{1}(E,E).
Lemma A.3.

Let 𝒟\mathcal{D} be a kk-linear triangulated category of finite type satisfies (A). Assume that there is a stability condition σ=(𝒜,Z)\sigma=(\mathcal{A},Z) on 𝒟\mathcal{D} satisfies (C).

  1. (1)

    The homological dimension of 𝒜\mathcal{A} is at most 22.

  2. (2)

    For any exact triangle AEBA\to E\to B with ϕσ(A)>ϕσ+(B)\phi^{-}_{\sigma}(A)>\phi^{+}_{\sigma}(B), we have

    ext1(A,A)+ext1(B,B)ext1(E,E).\operatorname{ext}^{1}(A,A)+\operatorname{ext}^{1}(B,B)\leq\operatorname{ext}^{1}(E,E).
  3. (3)

    For any non-zero object A𝒟A\in\mathcal{D}, we have ext1(A,A)n\operatorname{ext}^{1}(A,A)\geq n.

  4. (4)

    If a non-zero object EE is not σ\sigma-semistable, then any Harder–Narasimhan factor AA of EE satisfies

    ext1(A,A)<ext1(E,E).\operatorname{ext}^{1}(A,A)<\operatorname{ext}^{1}(E,E).
  5. (5)

    Any object EE with

    next1(E,E)<2nn\leq\operatorname{ext}^{1}(E,E)<2n

    is σ\sigma-semistable.

Proof.

Let A,B𝒜A,B\in\mathcal{A}. Then we have ϕσ+(A)1<ϕσ(B[k])\phi^{+}_{\sigma}(A)\leq 1<\phi^{-}_{\sigma}(B[k]) for any k1k\geq 1. Therefore, by (C) we get Hom(A,B[k+2])=Hom(A,B[k][2])=0\operatorname{Hom}(A,B[k+2])=\operatorname{Hom}(A,B[k][2])=0 for any k0k\geq 0. This proves (1).

Now for (2), note that Hom(A,B)=0\operatorname{Hom}(A,B)=0 and by (C) we have Hom(B,A[2])\operatorname{Hom}(B,A[2]). Then the result follows from Lemma A.2.

Next, we prove (3). If A0𝒜A\neq 0\in\mathcal{A}, then from (1), we get χ(A,A)=hom(A,A)ext1(A,A)+ext2(A,A)\chi(A,A)=\hom(A,A)-\operatorname{ext}^{1}(A,A)+\operatorname{ext}^{2}(A,A). Since χ(A,A)1n\chi(A,A)\leq 1-n by (A), we know that ext1(A,A)n\operatorname{ext}^{1}(A,A)\geq n in this case. Now for a general non-zero object A𝒟A\in\mathcal{D}, if AA is σ\sigma-semistable, then it is in 𝒜\mathcal{A} up to shift and the result follows from the previous argument. So we assume that AA is not σ\sigma-semistable. Let AA^{\prime} be the first Harder–Narasimhan factor of AA with respect to σ\sigma, and A′′:=cone(AA)A^{\prime\prime}:=\operatorname{cone}(A^{\prime}\to A). We have ϕσ(A)>ϕσ+(A′′)\phi_{\sigma}(A^{\prime})>\phi^{+}_{\sigma}(A^{\prime\prime}). Using (2) and σ\sigma-semistability of AA^{\prime}, we obtain next1(A,A)+ext1(A′′,A′′)ext1(A,A)n\leq\operatorname{ext}^{1}(A^{\prime},A^{\prime})+\operatorname{ext}^{1}(A^{\prime\prime},A^{\prime\prime})\leq\operatorname{ext}^{1}(A,A) and hence (3) follows. And (4) follows from the induction on the number of Harder–Narasimhan factors of EE and using (2) and (3).

Finally, if such EE in (5) is not σ\sigma-semistable, then by the existence of Harder–Narasimhan filtration, we can find a triangle AEBA\to E\to B with ϕσ(A)>ϕσ+(B)\phi^{-}_{\sigma}(A)>\phi^{+}_{\sigma}(B). By (2) and (3), this contradicts our assumption on ext1(E,E)\operatorname{ext}^{1}(E,E). Thus EE is σ\sigma-semistable. ∎

Now we are ready to prove our criterion.

Proof of Theorem A.1.

Since σ1\sigma_{1} and σ2\sigma_{2} have the same central charge, it remains to show 𝒜1=𝒜2\mathcal{A}_{1}=\mathcal{A}_{2}. By our assumptions, D1,D2𝒜1𝒜2D_{1},D_{2}\in\mathcal{A}_{1}\cap\mathcal{A}_{2} are both σ1\sigma_{1}-stable and σ2\sigma_{2}-stable with phases in (0,1](0,1].

Step 1. First, we show that if EE is a σi\sigma_{i}-semistable object which is also σj\sigma_{j}-semistable, then ϕσ1(E)=ϕσ2(E)\phi_{\sigma_{1}}(E)=\phi_{\sigma_{2}}(E), where {i,j}={1,2}\{i,j\}=\{1,2\}. Since σ1\sigma_{1} and σ2\sigma_{2} satisfy the same assumptions, in the following, we will take i=2i=2 and j=1j=1. The other case can be deduced from the same argument but exchanges the role of σ1\sigma_{1} and σ2\sigma_{2}.

Up to shift, we can assume that E𝒜2E\in\mathcal{A}_{2}. Since σ1\sigma_{1} and σ2\sigma_{2} have the same central charge, we have ϕσ1(E)ϕσ2(E)=2m\phi_{\sigma_{1}}(E)-\phi_{\sigma_{2}}(E)=2m for an integer mm. Then we see

2m<ϕσ1(E)2m+1.2m<\phi_{\sigma_{1}}(E)\leq 2m+1. (6)
  • Assume that there exist k,l{1,2}k,l\in\{1,2\} such that χ(Dk,E)<0\chi(D_{k},E)<0 and χ(E,Dl)<0\chi(E,D_{l})<0. Then by Lemma A.3 (1) and the fact that E,D1,D2𝒜2E,D_{1},D_{2}\in\mathcal{A}_{2}, we have

    ext1(Dk,E)0,ext1(E,Dl)0,\operatorname{ext}^{1}(D_{k},E)\neq 0,\quad\operatorname{ext}^{1}(E,D_{l})\neq 0,

    which imply

    1<ϕσ1(Dk)1ϕσ1(E)ϕσ1(Dl)+12.-1<\phi_{\sigma_{1}}(D_{k})-1\leq\phi_{\sigma_{1}}(E)\leq\phi_{\sigma_{1}}(D_{l})+1\leq 2.

    Hence, by (6) we get 2m<22m<2 and 1<2m+1-1<2m+1, which means m=0m=0 and we obtain ϕσ1(E)=ϕσ2(E)\phi_{\sigma_{1}}(E)=\phi_{\sigma_{2}}(E) as required.

  • Assume that χ(E,D1)0\chi(E,D_{1})\geq 0 and χ(E,D2)0\chi(E,D_{2})\geq 0. By (D), there is an integer s{1,2}s\in\{1,2\} such that χ(E,Ds)>0\chi(E,D_{s})>0 and μσ1(E)<μσ1(Ds)\mu_{\sigma_{1}}(E)<\mu_{\sigma_{1}}(D_{s}). Thus we have μσ2(E)<μσ2(Ds)\mu_{\sigma_{2}}(E)<\mu_{\sigma_{2}}(D_{s}) as well. Since E,Ds𝒜2E,D_{s}\in\mathcal{A}_{2}, we get ϕσ2(E)<ϕσ2(Ds)\phi_{\sigma_{2}}(E)<\phi_{\sigma_{2}}(D_{s}), which implies Hom(E,Ds[2])=0\operatorname{Hom}(E,D_{s}[2])=0 by (C). Then from χ(E,Ds)>0\chi(E,D_{s})>0 and Lemma A.3 (1), we obtain Hom(E,Ds)0\operatorname{Hom}(E,D_{s})\neq 0, and hence

    ϕσ1(E)ϕσ1(Ds)1.\phi_{\sigma_{1}}(E)\leq\phi_{\sigma_{1}}(D_{s})\leq 1.

    Now if one of χ(D1,E)\chi(D_{1},E) and χ(D2,E)\chi(D_{2},E) is negative, the same argument as in the first case shows that 1<ϕσ1(E)-1<\phi_{\sigma_{1}}(E).

    If χ(D1,E)0\chi(D_{1},E)\geq 0 and χ(D2,E)0\chi(D_{2},E)\geq 0, then by (E), there is an integer t{1,2}t\in\{1,2\} such that χ(Dt,E)>0\chi(D_{t},E)>0 and μσ1(E)=μσ2(E)>μσ1(Dt)=μσ2(Dt)\mu_{\sigma_{1}}(E)=\mu_{\sigma_{2}}(E)>\mu_{\sigma_{1}}(D_{t})=\mu_{\sigma_{2}}(D_{t}). Since E,Dt𝒜2E,D_{t}\in\mathcal{A}_{2}, we get ϕσ2(E)>ϕσ2(Dt)\phi_{\sigma_{2}}(E)>\phi_{\sigma_{2}}(D_{t}), which by (C) implies Hom(Dt,E[2])=0\operatorname{Hom}(D_{t},E[2])=0. Then together with χ(Dt,E)>0\chi(D_{t},E)>0 and Lemma A.3 (1), we see Hom(Dt,E)0\operatorname{Hom}(D_{t},E)\neq 0. Therefore, we have 0<ϕσ1(Dt)ϕσ1(E)0<\phi_{\sigma_{1}}(D_{t})\leq\phi_{\sigma_{1}}(E). In both cases, we always have ϕσ1(E)(1,2]\phi_{\sigma_{1}}(E)\in(-1,2]. By (6), we get m=0m=0 and ϕσ1(E)=ϕσ2(E)\phi_{\sigma_{1}}(E)=\phi_{\sigma_{2}}(E) as required.

  • Assume that χ(D1,E)0\chi(D_{1},E)\geq 0 and χ(D2,E)0\chi(D_{2},E)\geq 0. Then using (E), by a similar argument as the second case, we obtain ϕσ1(E)=ϕσ2(E)\phi_{\sigma_{1}}(E)=\phi_{\sigma_{2}}(E). This completes the first step.

Step 2. Next, we prove that an object EE is σ1\sigma_{1}-semistable if and only if σ2\sigma_{2}-semistable. We show this by induction on ext1(E,E)\operatorname{ext}^{1}(E,E). If ext1(E,E)<2n\operatorname{ext}^{1}(E,E)<2n, then from (B) we know such EE exists. By Lemma A.3 (5), EE is both σ1\sigma_{1}-semistable and σ2\sigma_{2}-semistable.

Now assume that the statement holds for any object FF with ext1(F,F)<N\operatorname{ext}^{1}(F,F)<N for an integer N>2nN>2n. Let EE be an object with ext1(E,E)=N\operatorname{ext}^{1}(E,E)=N. If EE is σi\sigma_{i}-semistable but not σj\sigma_{j}-semistable for {i,j}={1,2}\{i,j\}=\{1,2\}, let AA be the first Harder–Narasimhan factor of EE with respect to σj\sigma_{j} and BB be the last one. Therefore, we see

ϕσj(A)>ϕσj(B).\phi_{\sigma_{j}}(A)>\phi_{\sigma_{j}}(B). (7)

And from Lemma A.3 (4), we have

ext1(A,A)<ext1(E,E),ext1(B,B)<ext1(E,E).\operatorname{ext}^{1}(A,A)<\operatorname{ext}^{1}(E,E),\quad\operatorname{ext}^{1}(B,B)<\operatorname{ext}^{1}(E,E).

Therefore, by the induction hypothesis, AA and BB are σi\sigma_{i}-semistable as well and the first step implies

ϕσ1(A)=ϕσ2(A),ϕσ1(B)=ϕσ2(B).\phi_{\sigma_{1}}(A)=\phi_{\sigma_{2}}(A),\quad\phi_{\sigma_{1}}(B)=\phi_{\sigma_{2}}(B). (8)

But then from Hom(A,E)0\operatorname{Hom}(A,E)\neq 0 and Hom(E,B)0\operatorname{Hom}(E,B)\neq 0, we have ϕσi(A)ϕσi(E)ϕσi(B)\phi_{\sigma_{i}}(A)\leq\phi_{\sigma_{i}}(E)\leq\phi_{\sigma_{i}}(B), which contradicts (7) and (8). Hence EE is σj\sigma_{j}-semistable. This completes our induction argument.

Step 3. Finally, by the previous two steps, we know that an object EE is σ1\sigma_{1}-semistable if and only if σ2\sigma_{2}-semistable with ϕσ1(E)=ϕσ2(E)\phi_{\sigma_{1}}(E)=\phi_{\sigma_{2}}(E). Since every non-zero object in the heart is obtained by extensions of semistable objects with phases in (0,1](0,1], we know that 𝒜1=𝒜2\mathcal{A}_{1}=\mathcal{A}_{2}. This ends the proof of our theorem. ∎

A.1. Applications to Kuznetsov components of Fano threefolds

Let YdY_{d} be smooth index 22 degree d2d\geq 2 prime Fano threefold and X4d+2X_{4d+2} an index 11 degree 4d+24d+2 prime Fano threefold. In this section, we apply Theorem A.1 to show that all Serre-invariant stability conditions on 𝒦u(Yd)\mathcal{K}u(Y_{d}) and 𝒦u(X4d+2)\mathcal{K}u(X_{4d+2}) (or 𝒜X4d+2\mathcal{A}_{X_{4d+2}}) are in the same GL~+(2,)\widetilde{\mathrm{GL}}^{+}(2,\mathbb{R})-orbit for each d2d\geq 2 (Theorem A.10).

Recall that the Kuznetsov component of an index two prime Fano threefold YdY_{d} of degree dd is defined by 𝒦u(Yd):=𝒪Yd,𝒪Yd(H)\mathcal{K}u(Y_{d}):=\langle\operatorname{\mathcal{O}}_{Y_{d}},\operatorname{\mathcal{O}}_{Y_{d}}(H)\rangle^{\perp}. The numerical Grothendieck group 𝒩(𝒦u(Yd))\mathcal{N}(\mathcal{K}u(Y_{d})) is a rank two lattice generated by two classes

v=11dH2,w=H12H2+(161d)H2.v=1-\frac{1}{d}H^{2},\quad w=H-\frac{1}{2}H^{2}+(\frac{1}{6}-\frac{1}{d})H^{2}.

Moreover, under this basis, the Euler form is given by the matrix

(111dd).\begin{pmatrix}-1&-1\\ 1-d&-d\end{pmatrix}.

For index one cases, we assume that d2d\geq 2. Then the Kuznetsov component is defined by 𝒦u(X4d+2):=X4d+2,𝒪X4d+2\mathcal{K}u(X_{4d+2}):=\langle\mathcal{E}_{X_{4d+2}},\operatorname{\mathcal{O}}_{X_{4d+2}}\rangle^{\perp}, where X4d+2\mathcal{E}_{X_{4d+2}} is a certain exceptional bundle pulled back from a Grassmannian (cf. [32]).

By [5], σ(α,β)\sigma(\alpha,\beta) is a stability condition on 𝒦u(Yd)\mathcal{K}u(Y_{d}) and 𝒦u(X4d+2)\mathcal{K}u(X_{4d+2}) for suitable (α,β)(\alpha,\beta). Moreover, according to [52, 48], these stability conditions are all Serre-invariant.

Since for every index one prime Fano threefold X4d+2X_{4d+2} with d3d\geq 3, there is an index two prime Fano threefold YdY_{d} with 𝒦u(Yd)𝒦u(X4d+2)\mathcal{K}u(Y_{d})\simeq\mathcal{K}u(X_{4d+2}) by [32], hence we only need to consider Kuznetsov components of YdY_{d} for d2d\geq 2 and X10X_{10}. Moreover, 𝒦u(Y4)\mathcal{K}u(Y_{4}) is equivalent to the derived category of a smooth curve, and 𝒦u(Y5)\mathcal{K}u(Y_{5}) is equivalent to the derived category of the 33-Kronecker quiver. In these two cases, the result is known by [42] and [15]. So in the following, we mainly focus on 𝒟=𝒦u(Yd)\mathcal{D}=\mathcal{K}u(Y_{d}) for 2d32\leq d\leq 3 or 𝒦u(X10)\mathcal{K}u(X_{10}). We first prove some properties of Serre-invariant stability conditions.

Lemma A.4.

Let 𝒟=𝒦u(Yd)\mathcal{D}=\mathcal{K}u(Y_{d}) for 2d32\leq d\leq 3 or 𝒦u(X10)\mathcal{K}u(X_{10}) and σ=(𝒜,Z)\sigma=(\mathcal{A},Z) be a Serre-invariant stability condition on 𝒟\mathcal{D}. Then 𝒟\mathcal{D} satisfies (A) and (B) in Theorem A.1 and σ\sigma satisfies (C). Moreover, for any σ\sigma-semistable object E𝒟E\in\mathcal{D}, we have :

  1. (1)

    if 𝒟=𝒦u(Y3)\mathcal{D}=\mathcal{K}u(Y_{3}), then

    ϕσ(E)+1ϕσ(S𝒟(E))<ϕσ(E)+2,\phi_{\sigma}(E)+1\leq\phi_{\sigma}(S_{\mathcal{D}}(E))<\phi_{\sigma}(E)+2,
  2. (2)

    if 𝒟=𝒦u(Y2)\mathcal{D}=\mathcal{K}u(Y_{2}) or 𝒦u(X10)\mathcal{K}u(X_{10}), then

    ϕσ(S𝒟(E))=ϕσ(E)+2.\phi_{\sigma}(S_{\mathcal{D}}(E))=\phi_{\sigma}(E)+2.
Proof.

It is clear that 𝒟\mathcal{D} satisfies (A). And by [52, Lemma 5.16] and Proposition 8.7, 𝒟\mathcal{D} also satisfies (B). When 𝒟=𝒦u(Y3)\mathcal{D}=\mathcal{K}u(Y_{3}), by [52, Lemma 5.9] we have ϕσ(S𝒟(E))<ϕσ(E)+2\phi_{\sigma}(S_{\mathcal{D}}(E))<\phi_{\sigma}(E)+2. When 𝒟=𝒦u(Y2)\mathcal{D}=\mathcal{K}u(Y_{2}) or 𝒦u(X10)\mathcal{K}u(X_{10}), recall that S𝒟2[4]S^{2}_{\mathcal{D}}\cong[4]. Then the same argument as in [52, Lemma 5.9] shows that ϕσ(S𝒟(E))ϕσ(E)+2\phi_{\sigma}(S_{\mathcal{D}}(E))\leq\phi_{\sigma}(E)+2. Then for any two objects A,B𝒟A,B\in\mathcal{D} with ϕσ+(B)<ϕσ(A)\phi^{+}_{\sigma}(B)<\phi^{-}_{\sigma}(A), using (1) and (2) we see

ϕσ(A[2])=ϕσ(A)+2>ϕσ+(B)+2ϕσ+(S𝒟(B)).\phi^{-}_{\sigma}(A[2])=\phi^{-}_{\sigma}(A)+2>\phi^{+}_{\sigma}(B)+2\geq\phi^{+}_{\sigma}(S_{\mathcal{D}}(B)).

Hence Hom(B,A[2])=Hom(A[2],S𝒟(B))=0\operatorname{Hom}(B,A[2])=\operatorname{Hom}(A[2],S_{\mathcal{D}}(B))=0 and the condition (C) is satisfied.

When 𝒟=𝒦u(Y3)\mathcal{D}=\mathcal{K}u(Y_{3}), from [52, Lemma 5.11], we get ext1(E,E)0\operatorname{ext}^{1}(E,E)\neq 0, which implies ϕσ(E)+1ϕσ(S𝒟(E))\phi_{\sigma}(E)+1\leq\phi_{\sigma}(S_{\mathcal{D}}(E)). This proves (1).

When 𝒟=𝒦u(Y2)\mathcal{D}=\mathcal{K}u(Y_{2}) or 𝒦u(X10)\mathcal{K}u(X_{10}), since 𝒟\mathcal{D} satisfies (A) and (C), by Lemma A.3 we have ext1(E,E)0\operatorname{ext}^{1}(E,E)\neq 0, which implies ϕσ(E)+1ϕσ(S𝒟(E))\phi_{\sigma}(E)+1\leq\phi_{\sigma}(S_{\mathcal{D}}(E)). Now since [E]=[S𝒟(E)]𝒩(𝒟)[E]=[S_{\mathcal{D}}(E)]\in\mathcal{N}(\mathcal{D}), we have ϕσ(E)ϕσ(S𝒟(E))2\phi_{\sigma}(E)-\phi_{\sigma}(S_{\mathcal{D}}(E))\in 2\mathbb{Z}. Hence we get ϕσ(S𝒟(E))=ϕσ(E)+2\phi_{\sigma}(S_{\mathcal{D}}(E))=\phi_{\sigma}(E)+2.

Before verifying (D) and (E), we need several lemmas.

Lemma A.5.

Let XX be a GM threefold.

  1. (1)

    HilbX3t+m=\mathrm{Hilb}_{X}^{3t+m}=\varnothing for m<1m<1. Thus for any conic CXC\subset X and line LXL\subset X, we have Hom(IC,𝒪L(k))=0\operatorname{Hom}(I_{C},\operatorname{\mathcal{O}}_{L}(-k))=0 for any k>1k>1.

  2. (2)

    If a line LL and a conic CC satisfies LCL\cap C\neq\varnothing, then LCL\cap C is of length one and LCL\cup C is a twisted cubic.

  3. (3)

    Let Γ(X)×𝒞(X)\mathcal{I}\subset\Gamma(X)\times\mathcal{C}(X) be the incidence variety, i.e.

    ={(L,C):LC}.\mathcal{I}=\{(L,C)\leavevmode\nobreak\ \colon\leavevmode\nobreak\ L\cap C\neq\varnothing\}.

    Then the projection maps 𝒞(X)\mathcal{I}\to\mathcal{C}(X) and Γ(X)\mathcal{I}\to\Gamma(X) are surjective.

Proof.

By [53, Corollary 1.38], we have HilbX3t+m=\mathrm{Hilb}_{X}^{3t+m}=\varnothing for m<0m<0. Thus to prove (1), we only need to show HilbX3t=\mathrm{Hilb}_{X}^{3t}=\varnothing. From [53, Corollary 1.38], C2\langle C\rangle\cong\mathbb{P}^{2} for any [C]HilbX3t[C]\in\mathrm{Hilb}_{X}^{3t}. Since XX is an intersection of quadrics, such a CC cannot exist on XX. Hence HilbX3t=\mathrm{Hilb}_{X}^{3t}=\varnothing. Now note that the kernel of any non-zero map IC𝒪L(k)I_{C}\to\operatorname{\mathcal{O}}_{L}(-k) is the ideal sheaf of a closed subscheme with the Hilbert polynomial 3t+m3t+m for m2km\leq 2-k. Therefore, Hom(IC,𝒪L(k))=0\operatorname{Hom}(I_{C},\operatorname{\mathcal{O}}_{L}(-k))=0 when k>1k>1. This proves (1). For (2), note that χ(𝒪LC)=2length(LC)\chi(\operatorname{\mathcal{O}}_{L\cup C})=2-\mathrm{length}(L\cap C), then the result follows from (1).

Finally, we prove (3). Since dimΓ(X)=1\dim\Gamma(X)=1, all lines on XX sweep out a surface SS in XX. By Pic(X)=𝒪X(H)\operatorname{Pic}(X)=\mathbb{Z}\operatorname{\mathcal{O}}_{X}(H), we see S|mH|S\in|mH| for m>0m>0. Thus C.SC.H>0C.S\geq C.H>0 for any conic CC. In other words, CSC\cap S\neq\varnothing, hence any conic on XX intersects with a line. Thus 𝒞(X)\mathcal{I}\to\mathcal{C}(X) is surjective. Similarly, since XX is covered by conics, any line intersects with a conic. Then Γ(X)\mathcal{I}\to\Gamma(X) is surjective. ∎

Lemma A.6.

Let XX be a GM threefold. Then there exists a line LL and twisted cubics CC and DD on XX such that

  1. (1)

    [L]Γ(X)[L]\in\Gamma(X) is a smooth point,

  2. (2)

    IC𝒦u(X)I_{C}\notin\mathcal{K}u(X) and LC=Z(s)L\cup C=Z(s) for a section sH0()s\in H^{0}(\mathcal{E}^{\vee}),

  3. (3)

    LDL\subset D, ID𝒦u(X)I_{D}\in\mathcal{K}u(X) and ext1(ID,ID)=3\operatorname{ext}^{1}(I_{D},I_{D})=3.

Proof.

Let Γ(X)×𝒞(X)\mathcal{I}\subset\Gamma(X)\times\mathcal{C}(X) be the incidence variety. We denote by 𝒞1\mathcal{C}_{1} the sublocus of 𝒞(X)\mathcal{C}(X) parametrizing smooth conics ZZ such that their involutive conics are also smooth and hom(,IZ)=1\hom(\mathcal{E},I_{Z})=1. By Remark 7.17, 𝒞1\mathcal{C}_{1} is an open subscheme of 𝒞(X)\mathcal{C}(X). Let 1:=|Γ(X)×𝒞1\mathcal{I}_{1}:=\mathcal{I}|_{\Gamma(X)\times\mathcal{C}_{1}}. From [25, Theorem 3.4 (iii)] and [23, Section 3.1], Γ(X)\Gamma(X) is generic smooth. This implies that the image of p:1Γ(X)p\colon\mathcal{I}_{1}\to\Gamma(X) contains a smooth point.

Let LXL\subset X be a line such that [L]Γ(X)[L]\in\Gamma(X) is smooth and contained in the image of pp. Then p1([L])p^{-1}([L]) is non-empty and there is a conic [Z]𝒞1[Z]\in\mathcal{C}_{1} such that LZL\cap Z\neq\varnothing. We set D:=LZD:=L\cup Z. And since Hom(,IL)0\operatorname{Hom}(\mathcal{E},I_{L})\neq 0, there is a section sH0()s\in H^{0}(\mathcal{E}) such that LZ(s)L\subset Z(s). We define CC to be the residue curve of LL in Z(s)Z(s). It is clear that CC and DD are twisted cubics by Lemma A.5. Moreover, LL and ZZ intersect transversely at a single point. Then it remains to check IC𝒦u(X)I_{C}\notin\mathcal{K}u(X), ext1(ID,ID)=3\operatorname{ext}^{1}(I_{D},I_{D})=3 and ID𝒦u(X)I_{D}\in\mathcal{K}u(X).

Since CZ(s)C\subset Z(s), it is clear that Hom(,IC)0\operatorname{Hom}(\mathcal{E},I_{C})\neq 0, i.e. IC𝒦u(X)I_{C}\notin\mathcal{K}u(X). Moreover, by the construction, we have an exact sequence

0IDIZ𝒪L(1)0.0\to I_{D}\to I_{Z}\to\operatorname{\mathcal{O}}_{L}(-1)\to 0. (9)

Note that all conics are connected, hence the smoothness implies irreducibility. Since Hom(,IZ)=k\operatorname{Hom}(\mathcal{E},I_{Z})=k and the involutive conic ZZ^{\prime} is smooth, we know that ZZZ\cup Z^{\prime} only has two irreducible components which are both of degree 22, hence does not contain DD. This means the unique non-zero map in Hom(,IZ)=k\operatorname{Hom}(\mathcal{E},I_{Z})=k does not factor through IDI_{D}. Hence the induced map Hom(,IZ)Hom(,𝒪L(1))\operatorname{Hom}(\mathcal{E},I_{Z})\to\operatorname{Hom}(\mathcal{E},\operatorname{\mathcal{O}}_{L}(-1)) is injective. By RHom(,𝒪L(1))=k\mathrm{RHom}^{\bullet}(\mathcal{E},\operatorname{\mathcal{O}}_{L}(-1))=k, this map is actually an isomorphism. Therefore, applying Hom(,)\operatorname{Hom}(\mathcal{E},-) to (9), we obtain RHom(,ID)=0\mathrm{RHom}^{\bullet}(\mathcal{E},I_{D})=0, which implies ID𝒦u(X)I_{D}\in\mathcal{K}u(X).

To show ext1(ID,ID)=3\operatorname{ext}^{1}(I_{D},I_{D})=3, since hom(ID,ID)=1\hom(I_{D},I_{D})=1 and ext3(ID,ID)=0\operatorname{ext}^{3}(I_{D},I_{D})=0, by χ(ID,ID)=2\chi(I_{D},I_{D})=-2 we only need to prove ext2(ID,ID)=0\operatorname{ext}^{2}(I_{D},I_{D})=0. From the construction above, we see ext2(IZ,IZ)=0\operatorname{ext}^{2}(I_{Z},I_{Z})=0. Moreover, ext2(𝒪L(1),𝒪L(1))=0\operatorname{ext}^{2}(\operatorname{\mathcal{O}}_{L}(-1),\operatorname{\mathcal{O}}_{L}(-1))=0 since [L]Γ(X)[L]\in\Gamma(X) is a smooth point. And by the transversality of the intersection of LL and ZZ, we see the derived restriction 𝒪Z|L𝒪LZDb(L)\operatorname{\mathcal{O}}_{Z}|_{L}\cong\operatorname{\mathcal{O}}_{L\cap Z}\in\mathrm{D}^{b}(L). Hence ext2(𝒪Z,𝒪L(1))=ext1(IZ,𝒪L(1))=0\operatorname{ext}^{2}(\operatorname{\mathcal{O}}_{Z},\operatorname{\mathcal{O}}_{L}(-1))=\operatorname{ext}^{1}(I_{Z},\operatorname{\mathcal{O}}_{L}(-1))=0. Finally, by Lemma A.5 we have hom(IZ,𝒪L(2))=ext3(𝒪L(1),IZ)=0\hom(I_{Z},\operatorname{\mathcal{O}}_{L}(-2))=\operatorname{ext}^{3}(\operatorname{\mathcal{O}}_{L}(-1),I_{Z})=0. Then ext2(ID,ID)=0\operatorname{ext}^{2}(I_{D},I_{D})=0 follows from applying [47, Lemma 2.27] to (9). ∎

Lemma A.7.

Let XX be a GM threefold and L,C,DL,C,D as in Lemma A.6. We define F1:=pr(IC),F1:=IDF_{1}:=\mathrm{pr}^{\prime}(I_{C}),F_{1}^{\prime}:=I_{D} and F2:=pr(IL)F_{2}:=\mathrm{pr}^{\prime}(I_{L}). Then the objects F1,F1,F2F_{1},F_{1}^{\prime},F_{2} are stable with respect to any Serre-invariant stability condition on 𝒦u(X)\mathcal{K}u(X). Moreover, F1F_{1} and F1F_{1}^{\prime} have the same phase.

Proof.

From the construction, we see ext1(F1,F1)=3\operatorname{ext}^{1}(F_{1}^{\prime},F_{1}^{\prime})=3. By the same argument as in [54, Corollary 5.4], we have ext1(F1,F1)=3\operatorname{ext}^{1}(F_{1},F_{1})=3. Finally, applying [47, Lemma 2.27] to 2ILF2\mathcal{E}^{\oplus 2}\to I_{L}\to F_{2} and using RHom(𝒪L,𝒪L)=kk[1]\mathrm{RHom}^{\bullet}(\operatorname{\mathcal{O}}_{L},\operatorname{\mathcal{O}}_{L})=k\oplus k[-1] implies ext1(F2,F2)=3\operatorname{ext}^{1}(F_{2},F_{2})=3. Then the stability of F1,F1F_{1},F_{1}^{\prime} and F2F_{2} follows from Proposition 4.12.

As [F1]=[F1]𝒩(𝒦u(X))[F_{1}]=[F_{1}^{\prime}]\in\mathcal{N}(\mathcal{K}u(X)), we have χ(F1,F1)<0\chi(F_{1},F_{1}^{\prime})<0. Since exti(F1,F1)=exti(IC,ID)=0\operatorname{ext}^{i}(F_{1},F_{1}^{\prime})=\operatorname{ext}^{i}(I_{C},I_{D})=0 for i{1,2}i\notin\{1,2\}, we get Hom(F1,F1[1])=Hom(F1[1],S𝒦u(X)(F1))0\operatorname{Hom}(F_{1},F_{1}^{\prime}[1])=\operatorname{Hom}(F_{1}^{\prime}[1],S_{\mathcal{K}u(X)}(F_{1}))\neq 0. Using Lemma A.4, we obtain ϕσ(F1)1<ϕσ(F1)<ϕσ(F1)+1\phi_{\sigma}(F_{1}^{\prime})-1<\phi_{\sigma}(F_{1})<\phi_{\sigma}(F_{1}^{\prime})+1 for any Serre-invariant stability condition σ\sigma. Thus ϕσ(F1)=ϕσ(F1)\phi_{\sigma}(F_{1})=\phi_{\sigma}(F_{1}^{\prime}) since [F1]=[F1]𝒩(𝒦u(X))[F_{1}]=[F_{1}^{\prime}]\in\mathcal{N}(\mathcal{K}u(X)). ∎

Lemma A.8.

Let 𝒟=𝒦u(Yd)\mathcal{D}=\mathcal{K}u(Y_{d}) for 2d32\leq d\leq 3 or 𝒦u(X10)\mathcal{K}u(X_{10}). Then there exist two objects F1,F2𝒟F_{1},F_{2}\in\mathcal{D} such that for any Serre-invariant stability condition σ\sigma on 𝒟\mathcal{D}, F1F_{1} and F2F_{2} are σ\sigma-stable with

ϕσ(F2)1<ϕσ(F1)<ϕσ(F2).\phi_{\sigma}(F_{2})-1<\phi_{\sigma}(F_{1})<\phi_{\sigma}(F_{2}).

In particular, the image of the central charge is not contained in a line for any Serre-invariant stability condition on 𝒟\mathcal{D}.

Proof.

When 𝒟=𝒦u(Yd)\mathcal{D}=\mathcal{K}u(Y_{d}), we define F2:=Rom(IL,𝒪Yd(H))[1]F_{2}:=R\mathcal{H}om(I_{L},\operatorname{\mathcal{O}}_{Y_{d}}(-H))[1] and F1=ILF_{1}=I_{L}, where LYdL\subset Y_{d} is a line. Then by [52, Lemma 5.13] and [52, Remark 4.8], F1F_{1} and F2F_{2} are σ\sigma-stable for any Serre-invariant stability condition σ\sigma on 𝒟\mathcal{D} with ϕσ(F2)1<ϕσ(F1)<ϕσ(F2).\phi_{\sigma}(F_{2})-1<\phi_{\sigma}(F_{1})<\phi_{\sigma}(F_{2}).

Now assume that 𝒟=𝒦u(X10)\mathcal{D}=\mathcal{K}u(X_{10}). We take F1,F1F_{1},F_{1}^{\prime} and F2F_{2} as in Lemma A.7. By [54, Proposition 3.3, 5.3], we have pr(IC)pr(G)\mathrm{pr}^{\prime}(I_{C})\cong\mathrm{pr}^{\prime}(G), where GG fits into an exact triangle

𝒪X(H)[1]G𝒪L(2)\operatorname{\mathcal{O}}_{X}(-H)[1]\to G\to\operatorname{\mathcal{O}}_{L}(-2)

and is the twisted derived dual of the line LL.

First, we prove that Hom(F2,F1[1])0\operatorname{Hom}(F_{2},F_{1}[1])\neq 0. By adjunction, we have Hom(F2,F1[1])=Hom(IL,pr(G)[1])\operatorname{Hom}(F_{2},F_{1}[1])=\operatorname{Hom}(I_{L},\mathrm{pr}^{\prime}(G)[1]). And by [54, Proposition 5.3], pr(G)\mathrm{pr}^{\prime}(G) fits into an exact triangle

Gpr(G).G\to\mathrm{pr}^{\prime}(G)\to\mathcal{E}. (10)

Since |L𝒪L𝒪L(1)\mathcal{E}|_{L}\cong\operatorname{\mathcal{O}}_{L}\oplus\operatorname{\mathcal{O}}_{L}(-1), it is easy to see Exti(IL,)\operatorname{Ext}^{i}(I_{L},\mathcal{E}) for i2i\neq 2. So applying Hom(IL,)\operatorname{Hom}(I_{L},-) to (10), we get Hom(IL,pr(G)[1])=Hom(IL,G[1])=Hom(IL,𝒪L(2)[1])\operatorname{Hom}(I_{L},\mathrm{pr}^{\prime}(G)[1])=\operatorname{Hom}(I_{L},G[1])=\operatorname{Hom}(I_{L},\operatorname{\mathcal{O}}_{L}(-2)[1]). As the normal bundle NL/X10N_{L/X_{10}} is either 𝒪L𝒪L(1)\operatorname{\mathcal{O}}_{L}\oplus\operatorname{\mathcal{O}}_{L}(-1) or 𝒪L(1)𝒪L(2)\operatorname{\mathcal{O}}_{L}(1)\oplus\operatorname{\mathcal{O}}_{L}(-2) by [50, Lemma 4.2.1], we see the derived restriction IL|LNL/X𝒪L(1)[1]I_{L}|_{L}\cong N_{L/X}^{\vee}\oplus\operatorname{\mathcal{O}}_{L}(1)[1] from [21, Proposition 11.8]. Then Hom(IL,𝒪L(2)[1])0\operatorname{Hom}(I_{L},\operatorname{\mathcal{O}}_{L}(-2)[1])\neq 0 follows from a direct computation.

Next, we show that Hom(F1,F2)0\operatorname{Hom}(F_{1}^{\prime},F_{2})\neq 0. By the definition of pr\mathrm{pr}^{\prime}, pr(IL)\mathrm{pr}^{\prime}(I_{L}) fits into an exact triangle 2ILpr(IL)\mathcal{E}^{\oplus 2}\to I_{L}\to\mathrm{pr}^{\prime}(I_{L}). Then applying Hom(ID,)\operatorname{Hom}(I_{D},-) to this triangle, the result follows from Hom(ID,)=0\operatorname{Hom}(I_{D},\mathcal{E})=0 and Hom(ID,IL)0\operatorname{Hom}(I_{D},I_{L})\neq 0 since LDL\subset D.

By Lemma A.7, F1,F1F_{1},F_{1}^{\prime} and F2F_{2} are all stable with respect to any Serre-invariant stability condition on 𝒟\mathcal{D}. Therefore, combined with above results we get ϕσ(F1)=ϕσ(F1)<ϕσ(F2)<ϕσ(F1)+1\phi_{\sigma}(F_{1})=\phi_{\sigma}(F_{1}^{\prime})<\phi_{\sigma}(F_{2})<\phi_{\sigma}(F_{1})+1 as desired. ∎

Now we are ready to verify conditions (D) and (E) in Theorem A.1.

Lemma A.9.

Let 𝒟=𝒦u(Yd)\mathcal{D}=\mathcal{K}u(Y_{d}) for 2d32\leq d\leq 3 or 𝒦u(X10)\mathcal{K}u(X_{10}). Then there exists a Serre-invariant stability condition σ1\sigma_{1} on 𝒟\mathcal{D} with two σ1\sigma_{1}-stable objects D1,D2D_{1},D_{2} satisfying (D) and (E). Moreover,

  • we can assume that for any Serre-invariant stability condition σ\sigma on 𝒟\mathcal{D}, D1D_{1} and D2D_{2} are σ\sigma-stable with

    ϕσ(D1)1<ϕσ(D2)<ϕσ(D1),orϕσ(D1)<ϕσ(D2)<ϕσ(D1)+1,\phi_{\sigma}(D_{1})-1<\phi_{\sigma}(D_{2})<\phi_{\sigma}(D_{1}),\leavevmode\nobreak\ \text{or}\leavevmode\nobreak\ \leavevmode\nobreak\ \phi_{\sigma}(D_{1})<\phi_{\sigma}(D_{2})<\phi_{\sigma}(D_{1})+1,
  • any Serre-invariant stability condition on 𝒟\mathcal{D} with the same central charge as σ1\sigma_{1} satisfies (D) and (E).

Proof.

When 𝒟=𝒦u(Yd)\mathcal{D}=\mathcal{K}u(Y_{d}), we define D1:=Rom(IL,𝒪Yd(H))[1]D_{1}:=R\mathcal{H}om(I_{L},\operatorname{\mathcal{O}}_{Y_{d}}(-H))[1] and D2=IL[1]D_{2}=I_{L}[1], where LYdL\subset Y_{d} is a line. Then by [52, Lemma 5.13] and [52, Remark 4.8], D1D_{1} and D2D_{2} are σ\sigma-stable for any Serre-invariant stability condition σ\sigma on 𝒟\mathcal{D} with ϕσ(D1)<ϕσ(D2)<ϕσ(D1)+1.\phi_{\sigma}(D_{1})<\phi_{\sigma}(D_{2})<\phi_{\sigma}(D_{1})+1. In this case, we take σ1:=σ(α,12)\sigma_{1}:=\sigma(\alpha,-\frac{1}{2}) for α>0\alpha>0 sufficiently small. Then by [52, Section 4], D1,D2𝒜(α,12)D_{1},D_{2}\in\mathcal{A}(\alpha,-\frac{1}{2}). Now a direct computation shows that, for any object EE with [E]=av+bw[E]=av+bw, we have

  • χ(E,D2)=a+(d1)b\chi(E,D_{2})=a+(d-1)b, χ(D2,E)=a+b\chi(D_{2},E)=a+b; and μα,120(E)>μα,120(D2)b<0\mu^{0}_{\alpha,-\frac{1}{2}}(E)>\mu^{0}_{\alpha,-\frac{1}{2}}(D_{2})\iff b<0

  • χ(E,D1)=b\chi(E,D_{1})=-b, χ(D1,E)=[(d2)a+(d1)b]\chi(D_{1},E)=-[(d-2)a+(d-1)b]; and μα,120(E)>μα,120(D1)a+b<0\mu^{0}_{\alpha,-\frac{1}{2}}(E)>\mu^{0}_{\alpha,-\frac{1}{2}}(D_{1})\iff a+b<0.

Then it is straightforward to check (D) and (E) for σ1\sigma_{1}.

When 𝒟=𝒦u(X10)\mathcal{D}=\mathcal{K}u(X_{10}), we use the equivalence Ξ\Xi in Lemma 3.7 and prove every thing on 𝒜X10\mathcal{A}_{X_{10}}. Let σ1:=σ(α,β)\sigma_{1}:=\sigma(\alpha,\beta), where β<0\beta<0 and α>0\alpha>0 with β-\beta and α\alpha are sufficiently small. We set D1=IC[1]D_{1}=I_{C}[1] and D2=pr(F)[1]D_{2}=\mathrm{pr}(F)[1], where CXC\subset X is a smooth conic with IC𝒜X10I_{C}\in\mathcal{A}_{X_{10}} and FMG(2,1,5)F\in M_{G}(2,1,5) is non-locally free. It is clear that D1,D2𝒜(α,β)D_{1},D_{2}\in\mathcal{A}(\alpha,\beta) and are stable with respect to any Serre-invariant stability condition on 𝒜X10\mathcal{A}_{X_{10}} by Lemma 7.5 and Proposition 8.7. As in the previous case, it is straightforward to check (D) and (E) for σ1\sigma_{1}. Now we show that for any Serre-invariant stability condition σ\sigma on 𝒜X10\mathcal{A}_{X_{10}}, we have ϕσ(D1)1<ϕσ(D2)<ϕσ(D1).\phi_{\sigma}(D_{1})-1<\phi_{\sigma}(D_{2})<\phi_{\sigma}(D_{1}). Indeed, if σ=σ1\sigma=\sigma_{1}, then this follows from a direct computation of the slope function of σ1\sigma_{1}. When σσ1\sigma\neq\sigma_{1}, by Lemma A.8, up to GL~+(2,)\widetilde{\mathrm{GL}}^{+}(2,\mathbb{R})-action, we can assume that σ\sigma and σ1\sigma_{1} has the same central charge and ϕσ(D1)=ϕσ1(D1)\phi_{\sigma}(D_{1})=\phi_{\sigma_{1}}(D_{1}). Thus ϕσ(D2)ϕσ1(D2)2\phi_{\sigma}(D_{2})-\phi_{\sigma_{1}}(D_{2})\in 2\mathbb{Z}. We claim that ϕσ(D1)2<ϕσ(D2)<ϕσ(D1)\phi_{\sigma}(D_{1})-2<\phi_{\sigma}(D_{2})<\phi_{\sigma}(D_{1}), which implies ϕσ(D2)=ϕσ1(D2)\phi_{\sigma}(D_{2})=\phi_{\sigma_{1}}(D_{2}) and the result follows. Indeed, by Proposition 8.1 we have an exact sequence 0F𝒪L(1)00\to F\to\mathcal{E}\to\operatorname{\mathcal{O}}_{L}(-1)\to 0 for a line LX10L\subset X_{10}. Hence applying Hom(,D1)\operatorname{Hom}(-,D_{1}) to this exact sequence and use Hom(,D1[1])0\operatorname{Hom}(\mathcal{E},D_{1}[-1])\neq 0 (Lemma 6.3) and adjunction of pr\mathrm{pr}, we have Hom(F,IC)=Hom(D2,D1)=Hom(D1,S𝒜X10(D2))0\operatorname{Hom}(F,I_{C})=\operatorname{Hom}(D_{2},D_{1})=\operatorname{Hom}(D_{1},S_{\mathcal{A}_{X_{10}}}(D_{2}))\neq 0. Then by Lemma A.4 we obtain ϕσ(D1)2<ϕσ(D2)<ϕσ(D1)\phi_{\sigma}(D_{1})-2<\phi_{\sigma}(D_{2})<\phi_{\sigma}(D_{1}) as desired.

The final statement follows from the fact that (D) and (E) in this case only depend on the central charge and numerical classes [D1][D_{1}] and [D2][D_{2}], as we have seen above. ∎

Applying Theorem A.1, we obtain the uniqueness of Serre-invariant stability conditions.

Theorem A.10.

Let 𝒟=𝒦u(Yd)\mathcal{D}=\mathcal{K}u(Y_{d}) for 2d32\leq d\leq 3 or 𝒦u(X10)\mathcal{K}u(X_{10}). Then all Serre-invariant stability conditions on 𝒟\mathcal{D} are in the same GL~+(2,)\widetilde{\mathrm{GL}}^{+}(2,\mathbb{R})-orbit.

Proof.

Let σ1\sigma_{1}, D1D_{1} and D2D_{2} as in Lemma A.9. Let σ2\sigma_{2} be another Serre-invariant stability condition on 𝒟\mathcal{D}. By Lemma A.8, up to GL~+(2,)\widetilde{\mathrm{GL}}^{+}(2,\mathbb{R})-action, we can assume that σ1\sigma_{1} and σ2\sigma_{2} have the same central charge. Moreover, up to shift we can assume that ϕσ1(D1)=ϕσ2(D1)\phi_{\sigma_{1}}(D_{1})=\phi_{\sigma_{2}}(D_{1}). Thus ϕσ1(D2)ϕσ2(D2)2\phi_{\sigma_{1}}(D_{2})-\phi_{\sigma_{2}}(D_{2})\in 2\mathbb{Z}. And by Lemma A.9, σ2\sigma_{2} also satisfies (D) and (E).

Now from Lemma A.9, we have ϕσk(D1)1<ϕσk(D2)<ϕσk(D1)\phi_{\sigma_{k}}(D_{1})-1<\phi_{\sigma_{k}}(D_{2})<\phi_{\sigma_{k}}(D_{1}) or ϕσk(D1)<ϕσk(D2)<ϕσk(D1)+1\phi_{\sigma_{k}}(D_{1})<\phi_{\sigma_{k}}(D_{2})<\phi_{\sigma_{k}}(D_{1})+1 for any k{1,2}k\in\{1,2\}. This implies ϕσ1(D2)=ϕσ2(D2)\phi_{\sigma_{1}}(D_{2})=\phi_{\sigma_{2}}(D_{2}). Therefore, by Lemma A.4 and Lemma A.9, we can apply Theorem A.1 and get σ1=σ2\sigma_{1}=\sigma_{2}. ∎

Remark A.11.

The idea of the proof of Theorem A.10 was first explained to us by Arend Bayer. In [54, Proposition 4.21], one of the authors made an attempt to prove this statement but the argument is incomplete. Here, we fill the gaps and give a more general argument. Later, in [16, Theorem 3.1], the authors also prove the uniqueness of Serre-invariant stability conditions for a general triangulated category satisfying a list of assumptions and include Kuznetsov components of cubic threefolds and very general cubic fourfolds. The assumptions used in [16, Theorem 3.1] are (A), (B), and the Serre functor of 𝒟\mathcal{D} satisfies S𝒟r=[k]S^{r}_{\mathcal{D}}=[k] with 0<k/r<20<k/r<2 or r=2r=2 and k=4k=4, while our Theorem A.1 also works for general triangulated categories which are not fractional Calabi–Yau but with extra assumptions (C), (D) and (E). Indeed, if we take D1=DD_{1}=D and D2=S𝒟(D)[2]D_{2}=S_{\mathcal{D}}(D)[-2] in (D) and (E) where DD is an object in (B), then one can show that when k/r<2k/r<2, Theorem A.1 implies [16, Theorem 3.1]. Moreover, Theorem A.1 can be applied to the derived category of a smooth projective curve or a generalized Kronecker quiver as well.

References

  • APR [22] Matteo Altavilla, Marin Petković, and Franco Rota. Moduli spaces on the Kuznetsov component of Fano threefolds of index 2. Épijournal Géom. Algébrique, 6:Art. 13, 31, 2022.
  • BB [17] Arend Bayer and Tom Bridgeland. Derived automorphism groups of K3 surfaces of Picard rank 1. Duke Math. J., 166(1):75–124, 2017.
  • BBF+ [20] Arend Bayer, Sjoerd Beentjes, Soheyla Feyzbakhsh, Georg Hein, Diletta Martinelli, Fatemeh Rezaee, and Benjamin Schmidt. The desingularization of the theta divisor of a cubic threefold as a moduli space. arXiv preprint, arXiv:2011.12240, 2020.
  • BF [14] Maria Chiara Brambilla and Daniele Faenzi. Vector bundles on Fano threefolds of genus 7 and Brill–Noether loci. Internat. J. Math., 25(3):1450023, 59, 2014.
  • BLMS [23] Arend Bayer, Martí Lahoz, Emanuele Macrì, and Paolo Stellari. Stability conditions on Kuznetsov components. Ann. Sci. Éc. Norm. Supér. (4), 56(2):517–570, 2023. With an appendix by Bayer, Lahoz, Macrì, Stellari and X. Zhao.
  • BMK [23] Arend Bayer, Emanuele Macri, and Alexander Kuznetsov. Vector bundles Fano threefolds. preprint, 2023.
  • BMMS [12] Marcello Bernardara, Emanuele Macrì, Sukhendu Mehrotra, and Paolo Stellari. A categorical invariant for cubic threefolds. Adv. Math., 229(2):770–803, 2012.
  • BMS [16] Arend Bayer, Emanuele Macrì, and Paolo Stellari. The space of stability conditions on abelian threefolds, and on some Calabi–Yau threefolds. Invent. Math., 206(3):869–933, 2016.
  • BMT [14] Arend Bayer, Emanuele Macrì, and Yukinobu Toda. Bridgeland stability conditions on threefolds I: Bogomolov–Gieseker type inequalities. J. Algebraic Geom., 23(1):117–163, 2014.
  • BO [01] Alexei Bondal and Dmitri Orlov. Reconstruction of a variety from the derived category and groups of autoequivalences. Compositio Math., 125(3):327–344, 2001.
  • Bri [07] Tom Bridgeland. Stability conditions on triangulated categories. Ann. of Math. (2), 166(2):317–345, 2007.
  • BT [16] Marcello Bernardara and Gonçalo Tabuada. From semi-orthogonal decompositions to polarized intermediate Jacobians via Jacobians of noncommutative motives. Mosc. Math. J., 16(2):205–235, 2016.
  • DIM [12] Olivier Debarre, Atanas Iliev, and Laurent Manivel. On the period map for prime Fano threefolds of degree 10. J. Algebraic Geom., 21(1):21–59, 2012.
  • DK [18] Olivier Debarre and Alexander Kuznetsov. Gushel–Mukai varieties: classification and birationalities. Algebr. Geom., 5(1):15–76, 2018.
  • DK [19] George Dimitrov and Ludmil Katzarkov. Bridgeland stability conditions on wild Kronecker quivers. Adv. Math., 352:27–55, 2019.
  • FP [23] Soheyla Feyzbakhsh and Laura Pertusi. Serre-invariant stability conditions and Ulrich bundles on cubic threefolds. Épijournal Géom. Algébrique, 7:Art. 1, 32, 2023.
  • FV [21] Daniele Faenzi and Alessandro Verra. Fano threefolds of genus 10 and Coble cubics. In preparation, 2021.
  • GLZ [22] Hanfei Guo, Zhiyu Liu, and Shizhuo Zhang. Conics on Gushel–Mukai fourfolds, EPW sextics and Bridgeland moduli spaces. arXiv preprint, arXiv:2203.05442, 2022.
  • HLOY [03] Shinobu Hosono, Bong H. Lian, Keiji Oguiso, and Shing-Tung Yau. Fourier–Mukai partners of a K3K3 surface of Picard number one. In Vector bundles and representation theory (Columbia, MO, 2002), volume 322 of Contemp. Math., pages 43–55. Amer. Math. Soc., Providence, RI, 2003.
  • HRS [96] Dieter Happel, Idun Reiten, and Sverre O. Smalø. Tilting in abelian categories and quasitilted algebras. Mem. Amer. Math. Soc., 120(575):viii+ 88, 1996.
  • Huy [06] Daniel Huybrechts. Fourier–Mukai transforms in algebraic geometry. Oxford Mathematical Monographs. The Clarendon Press, Oxford University Press, Oxford, 2006.
  • [22] Atanas Iliev. The Fano surface of the Gushel threefold. Compositio Math., 94(1):81–107, 1994.
  • [23] Atanas Iliev. Lines on the Gushel’ threefold. Indag. Math. (N.S.), 5(3):307–320, 1994.
  • IM [11] Atanas Iliev and Laurent Manivel. Fano manifolds of degree ten and EPW sextics. Ann. Sci. Éc. Norm. Supér. (4), 44(3):393–426, 2011.
  • Isk [78] V. A. Iskovskih. Fano threefolds. II. Izv. Akad. Nauk SSSR Ser. Mat., 42(3):506–549, 1978.
  • JLZ [22] Augustinas Jacovskis, Zhiyu Liu, and Shizhuo Zhang. Brill–Noether theory and categorical Torelli theorems for Kuznetsov components of index 1 Fano threefolds. arXiv preprint, arXiv:2207.01021, 2022.
  • Kos [20] Naoki Koseki. On the Bogomolov–Gieseker inequality for hypersurfaces in the projective spaces. arXiv preprint, arXiv:2008.09799, 2020.
  • KP [17] Alexander Kuznetsov and Alexander Perry. Derived categories of cyclic covers and their branch divisors. Selecta Math. (N.S.), 23(1):389–423, 2017.
  • KP [18] Alexander Kuznetsov and Alexander Perry. Derived categories of Gushel–Mukai varieties. Compos. Math., 154(7):1362–1406, 2018.
  • KP [23] Alexander Kuznetsov and Alexander Perry. Categorical cones and quadratic homological projective duality. Ann. Sci. Éc. Norm. Supér. (4), 56(1):1–57, 2023.
  • KPS [18] Alexander G. Kuznetsov, Yuri G. Prokhorov, and Constantin A. Shramov. Hilbert schemes of lines and conics and automorphism groups of Fano threefolds. Jpn. J. Math., 13(1):109–185, 2018.
  • Kuz [04] Alexander Kuznetsov. Derived category of a cubic threefold and the variety V14V_{14}. Tr. Mat. Inst. Steklova, 246:183–207, 2004.
  • Kuz [09] Alexander Kuznetsov. Derived categories of Fano threefolds. Tr. Mat. Inst. Steklova, 264:116–128, 2009.
  • Kuz [10] Alexander Kuznetsov. Derived categories of cubic fourfolds. In Cohomological and geometric approaches to rationality problems, volume 282 of Progr. Math., pages 219–243. Birkhäuser Boston, Boston, MA, 2010.
  • Kuz [19] Alexander Kuznetsov. Calabi–Yau and fractional Calabi–Yau categories. J. Reine Angew. Math., 753:239–267, 2019.
  • Li [19] Chunyi Li. Stability conditions on Fano threefolds of Picard number 1. J. Eur. Math. Soc. (JEMS), 21(3):709–726, 2019.
  • Li [21] Chunyi Li. Stronger Bogomolov–Gieseker inequality for Fano varieties of Picard number one. In preparation, 2021.
  • LNSZ [21] Chunyi Li, Howard Nuer, Paolo Stellari, and Xiaolei Zhao. A refined derived Torelli theorem for Enriques surfaces. Math. Ann., 379(3-4):1475–1505, 2021.
  • Log [12] Dmitry Logachev. Fano threefolds of genus 6. Asian J. Math., 16(3):515–559, 2012.
  • LSZ [22] Chunyi Li, Paolo Stellari, and Xiaolei Zhao. A refined derived Torelli theorem for Enriques surfaces, II: the non-generic case. Math. Z., 300(4):3527–3550, 2022.
  • LZ [22] Zhiyu Liu and Shizhuo Zhang. A note on Bridgeland moduli spaces and moduli spaces of sheaves on X14X_{14} and Y3Y_{3}. Math. Z., 302(2):803–837, 2022.
  • Mac [07] Emanuele Macrì. Stability conditions on curves. Math. Res. Lett., 14(4):657–672, 2007.
  • Muk [92] Shigeru Mukai. Fano 33-folds. In Complex projective geometry (Trieste, 1989/Bergen, 1989), volume 179 of London Math. Soc. Lecture Note Ser., pages 255–263. Cambridge Univ. Press, Cambridge, 1992.
  • Ogu [02] Keiji Oguiso. K3 surfaces via almost-primes. Math. Res. Lett., 9(1):47–63, 2002.
  • Orl [97] D. O. Orlov. Equivalences of derived categories and K3K3 surfaces. J. Math. Sci. (New York), 84(5):1361–1381, 1997. Algebraic geometry, 7.
  • Per [22] Alexander Perry. The integral Hodge conjecture for two-dimensional Calabi–Yau categories. Compos. Math., 158(2):287–333, 2022.
  • Pir [23] Dmitrii Pirozhkov. Admissible subcategories of del Pezzo surfaces. Adv. Math., 424:Paper No. 109046, 62, 2023.
  • [48] Laura Pertusi and Ethan Robinett. Stability conditions on Kuznetsov components of Gushel–Mukai threefolds and Serre functor. Math. Nachr., 296(7):2975–3002, 2023.
  • [49] Marin Petković and Franco Rota. A note on the Kuznetsov component of the Veronese double cone. J. Pure Appl. Algebra, 227(3):Paper No. 107214, 23, 2023.
  • PS [99] A. N. Parshin and I. R. Shafarevich, editors. Algebraic geometry. V, volume 47 of Encyclopaedia of Mathematical Sciences. Springer-Verlag, Berlin, 1999. Fano varieties, A translation of Algebraic geometry. 5 (Russian), Ross. Akad. Nauk, Vseross. Inst. Nauchn. i Tekhn. Inform., Moscow.
  • PS [23] Laura Pertusi and Paolo Stellari. Categorical Torelli theorems: results and open problems. Rend. Circ. Mat. Palermo (2), 72(5):2949–3011, 2023.
  • PY [22] Laura Pertusi and Song Yang. Some remarks on Fano threefolds of index two and stability conditions. Int. Math. Res. Not. IMRN, (17):12940–12983, 2022.
  • San [14] Giangiacomo Sanna. Rational curves and instantons on the Fano threefold Y5Y_{5}. arXiv preprint, arXiv:1411.7994, 2014.
  • Zha [20] Shizhuo Zhang. Bridgeland moduli spaces and Kuznetsov’s Fano threefold conjecture. arXiv preprint, arXiv:2012.12193, 2020.