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

Reider-type theorems on normal surfaces via Bridgeland stability

Anne Larsen  and  Anda Tenie
Abstract.

Using Langer’s construction of Bridgeland stability conditions on normal surfaces, we prove Reider-type theorems generalizing the work done by Arcara-Bertram in the smooth case. Our results still hold in positive characteristic or when ωXL\omega_{X}\otimes L is not necessarily a line bundle. They also hold when the dualizing sheaf is replaced by a variant arising from the theory of Du Bois complexes. For complex surfaces with at most rational double point singularities, we recover the optimal bounds for global generation and very ampleness as predicted by Fujita’s conjecture.

A.L. was partially supported by the National Science Foundation Graduate Research Fellowship under Grant No. 2141064.

1. Introduction

In [5] Fujita proposed a fundamental conjecture in algebraic geometry stating that if LL is an ample line bundle on a smooth projective complex variety XX of dimension nn, then ωXLn+1\omega_{X}\otimes L^{\otimes n+1} is globally generated and ωXLn+2\omega_{X}\otimes L^{\otimes n+2} is very ample. The conjecture has seen much interest over the years and the global generation part is now known in dimension 5\leq 5 ([17], [3], [9], [23]).

One strategy that has seen much success (including in giving effective bounds for higher dimensional varieties) involves vanishing theorems. Another approach, which is, however, hard to generalize in higher dimensions, is Reider’s strategy of using vector bundle techniques and Bogomolov’s inequality. We will focus on the latter approach in the context of normal surfaces. Our main source of inspiration is Arcara and Bertram’s reinterpretation of Reider’s theorem in the context of Bridgeland stability conditions [1].

Our work is concerned with giving Reider-type (and generation of higher jets of ωXLa\omega_{X}\otimes L^{\otimes a}) bounds on normal surfaces. For this generalization to singular surfaces, another key ingredient used in our work is Langer’s recent construction of stability conditions on normal surfaces [13].

Remark 1.1.

Other Reider-type bounds for normal surfaces under the assumption that KX+HK_{X}+H is Cartier can be found in [19], [4]. For generation of higher jets results when XX has in addition at most rational singularities one can consult [12].

We now discuss the main results of this article. Let XX be a projective normal surface defined over an algebraically closed field (of any characteristic), ZXZ\subset X a dimension 0 subscheme of length lZl_{Z}, and let LL be an ample line bundle on XX. The constant CXC_{X} will be defined in Section 2 and in characteristic 0, depends only on the singularities of XX; roughly speaking, CXC_{X} ensures a Bogomolov inequality holds on X.X. For concrete examples where this constant is computed see Section 2.2.

Theorem 1.2.

Let XX be a projective normal surface defined over an algebraically closed field, ZXZ\subset X a zero-dimensional subscheme of length lZl_{Z}, and LL an ample line bundle on XX. Suppose there is a pair of nonnegative integers l1+l2=lZl_{1}+l_{2}=l_{Z} such that L2max((CX+2l2+1)2,4(l1+l2+CX)+ϵ)L^{2}\geq\max((C_{X}+2l_{2}+1)^{2},4(l_{1}+l_{2}+C_{X})+\epsilon) and LCmax(2(2l1+CX),CX+2l2+1)L\cdot C\geq\max(2(2l_{1}+C_{X}),C_{X}+2l_{2}+1) for any curve CC. Then we have the vanishing

H1(X,ωX𝕃LIZ)=0.H^{1}(X,\omega_{X}\otimes^{\mathbb{L}}L\otimes I_{Z})=0.

In particular, this implies ωXL\omega_{X}\otimes L separates jets along Z,Z, i.e.

H0(ωXL)H0(ωXL𝒪Z)H^{0}(\omega_{X}\otimes L)\rightarrow H^{0}(\omega_{X}\otimes L\otimes\mathcal{O}_{Z})

is surjective.

Remark 1.3.

For a “Reider-type” statement see Theorem 4.10. This variant will also recover in Corollary 4.11 the very ampleness part of Reider’s theorem for smooth surfaces when LL is assumed to be an ample line bundle.

Definition 1.4.

Define a function m:0×0m:\mathbb{R}_{\geq 0}\times\mathbb{Z}_{\geq 0}\to\mathbb{R} by

(CX,l)minl1+l2=ll1,l20{max(2(2l1+CX),CX+2l2+1)}.(C_{X},l)\mapsto\min_{\begin{subarray}{c}l_{1}+l_{2}=l\\ l_{1},l_{2}\in\mathbb{Z}_{\geq 0}\end{subarray}}\{\max\left(2(2l_{1}+C_{X}),C_{X}+2l_{2}+1\right)\}.

Also define

m(CX,l):={3CX=1,l=0m(CX,l)elsem^{\prime}(C_{X},l):=\begin{cases}3&C_{X}=1,l=0\\ m(C_{X},l)&\textrm{else}\end{cases}

Note, in particular, that m(0,1)=3m^{\prime}(0,1)=3 and m(0,2)=4m^{\prime}(0,2)=4 which will recover the global generation and very ampleness bounds in Fujita’s conjecture for complex surfaces with at worst rational double point singularities.

Corollary 1.5.

Let XX be a projective normal surface defined over an algebraically closed field, ZXZ\subset X a subscheme of length lZl_{Z}, and LL an ample line bundle on XX. Then

H1(X,ωX𝕃LaIZ)=0H^{1}(X,\omega_{X}\otimes^{\mathbb{L}}L^{\otimes a}\otimes I_{Z})=0

for any integer am(CX,lZ).a\geq m^{\prime}(C_{X},l_{Z}). In particular, this implies ωXLa\omega_{X}\otimes L^{\otimes a} separates jets along Z,Z, i.e.

H0(ωXLa)H0(ωXLa𝒪Z)H^{0}(\omega_{X}\otimes L^{\otimes a})\rightarrow H^{0}(\omega_{X}\otimes L^{\otimes a}\otimes\mathcal{O}_{Z})

is surjective.

Remark 1.6.

Unlike previous work, which gives similar results for a divisor HH under the assumption that KX+HK_{X}+H is Cartier, our theorem shows separation of jets for the tensor product of an ample bundle with the canonical sheaf, even when this is not a line bundle.

We also recover the classical Fujita conjecture for surfaces:

Corollary 1.7.

If XX is a projective normal surface with CX=0C_{X}=0 (e.g. a complex surface with at most rational double point singularities) and LL an ample line bundle, then ωXLa\omega_{X}\otimes L^{\otimes a} is globally generated for a3a\geq 3 and very ample for a4a\geq 4.

Remark 1.8.

The proof of the optimal bound for very ampleness in the smooth case is contained in [1], but our argument appears to be the first proof of the correct bound for global generation via Bridgeland stability.

In positive characteristic, we can no longer take CX=0C_{X}=0 for every smooth surface XX, but there is still an explicit bound:

Corollary 1.9.

Let XX be a smooth projective surface defined over an algebraically closed field of positive characteristic and ZXZ\subset X a dimension 0 subscheme of length lZ.l_{Z}. Following Koseki [10], CXC_{X} depends only on the birational equivalence class of XX and is defined as:

  1. (1)

    If XX is a minimal surface of general type, then CX=2+5KX2χ(𝒪X).C_{X}=2+5K_{X}^{2}-\chi(\mathcal{O}_{X}).

  2. (2)

    If κ(X)=1\kappa(X)=1 and XX is quasi-elliptic, then CX=2χ(𝒪X).C_{X}=2-\chi(\mathcal{O}_{X}).

  3. (3)

    Otherwise, CX=0.C_{X}=0.

Then

H1(X,ωX𝕃LaIZ)=0H^{1}(X,\omega_{X}\otimes^{\mathbb{L}}L^{\otimes a}\otimes I_{Z})=0

for any integer am(CX,lZ).a\geq m^{\prime}(C_{X},l_{Z}). In particular, this implies ωXLa\omega_{X}\otimes L^{\otimes a} separates jets along Z,Z, i.e.

H0(ωXLa)H0(ωXLa𝒪Z)H^{0}(\omega_{X}\otimes L^{\otimes a})\rightarrow H^{0}(\omega_{X}\otimes L^{\otimes a}\otimes\mathcal{O}_{Z})

is surjective.

Remark 1.10.

It is known that Fujita’s conjecture does not hold for smooth surfaces in positive characteristic. In fact, for any power aa one can find a smooth surface XX such that ωXLa\omega_{X}\otimes L^{\otimes a} is not globally generated [7]. Our results account for this, as the bound on aa depends on the constant CXC_{X} corresponding to the surface X.X.

Remark 1.11.

Note that for lZ=0l_{Z}=0 we get a Kodaira vanishing type result in positive characteristic, again depending on CX.C_{X}.

Now specializing to the case when XX is complex, we also obtain similar results when the dualizing sheaf ωX\omega_{X} is replaced by the dual of Ω¯X00,\operatorname{\underline{\Omega}_{X}^{0}}0, where Ω¯X00\operatorname{\underline{\Omega}_{X}^{0}}0 is the 0-th Du Bois complex of XX. This line of inquiry was based on the observation that Ω¯X00[1]\operatorname{\underline{\Omega}_{X}^{0}}0[1] has the shape of a (potentially) Bridgeland stable object (namely, slope stable sheaf in degree 1-1 and dimension 0 sheaf in degree 0), and follows the philosophy of viewing the Du Bois complexes as better-behaved versions of the sheaves of Kähler differentials on XX (or in this case, of Ω¯X00\operatorname{\underline{\Omega}_{X}^{0}}0 as a better-behaved version of 𝒪X\mathscr{O}_{X}). To be more precise, consider

ωXDB:=𝐑om(Ω¯X00,ωX),\omega_{X}^{DB}:={\bf R}\mathcal{H}om(\operatorname{\underline{\Omega}_{X}^{0}}0,\omega_{X}),

which is in fact a sheaf (see Section 2.3 for details). Then we obtain similar Reider-type results for ωXDBLa.\omega_{X}^{DB}\otimes L^{\otimes a}. (It may be helpful to keep in mind that, at least in the Gorenstein case, ωXDB=ωXI\omega_{X}^{DB}=\omega_{X}\otimes I, where II is an ideal sheaf supported exactly on the non-Du Bois locus. Thus, the problem of ωXDBLa\omega_{X}^{DB}\otimes L^{\otimes a} separating jets along some zero-dimensional subscheme ZZ is more or less the problem of ωXLa\omega_{X}\otimes L^{\otimes a} separating jets along a larger subscheme containing both ZZ and the non-Du Bois locus, and so the bounds are correspondingly worse.)

Theorem 1.12.

Let XX be a projective complex normal surface, ZXZ\subset X a subscheme of length lZl_{Z}, and LL an ample line bundle on XX. Let lT:=length(1(Ω¯X00))l_{T}:=\operatorname{length}(\mathcal{H}^{1}(\operatorname{\underline{\Omega}_{X}^{0}}0)). Let l1,l2l_{1},l_{2} be a pair of nonnegative integers such that lZ+lT=l1+l2l_{Z}+l_{T}=l_{1}+l_{2}. Suppose L2max((CX+2l2+1)2,4(l1+l2+CX)+ϵ)L^{2}\geq\max((C_{X}+2l_{2}+1)^{2},4(l_{1}+l_{2}+C_{X})+\epsilon) and LCmax(2(2l1+CX),CX+2l2+1)L\cdot C\geq\max(2(2l_{1}+C_{X}),C_{X}+2l_{2}+1) for every curve CC. Then we have the vanishing

H1(X,ωXDB𝕃LIZ)=0.H^{1}(X,\omega_{X}^{DB}\otimes^{\mathbb{L}}L\otimes I_{Z})=0.

In particular, this implies ωXDBL\omega_{X}^{DB}\otimes L separates jets along Z,Z, i.e.

H0(ωXDBL)H0(ωXDBL𝒪Z)H^{0}(\omega_{X}^{DB}\otimes L)\rightarrow H^{0}(\omega_{X}^{DB}\otimes L\otimes\mathcal{O}_{Z})

is surjective.

Corollary 1.13.

Let XX be a projective normal complex surface, ZXZ\subset X a subscheme of length lZl_{Z}, and LL an ample line bundle on XX. Let lX:=length(1(Ω¯X00))l_{X}:=\operatorname{length}(\mathcal{H}^{1}(\operatorname{\underline{\Omega}_{X}^{0}}0)). Then

H1(X,ωXDB𝕃LaIZ)=0H^{1}(X,\omega_{X}^{DB}\otimes^{\mathbb{L}}L^{\otimes a}\otimes I_{Z})=0

for any integer am(CX,lZ+lX)a\geq m^{\prime}(C_{X},l_{Z}+l_{X}), where mm^{\prime} is defined in Definition 1.4. In particular, this implies ωXDBLa\omega_{X}^{DB}\otimes L^{\otimes a} separates jets along Z,Z, i.e.

H0(ωXDBLa)H0(ωXDBLa𝒪Z)H^{0}(\omega_{X}^{DB}\otimes L^{\otimes a})\rightarrow H^{0}(\omega_{X}^{DB}\otimes L^{\otimes a}\otimes\mathcal{O}_{Z})

is surjective.

We now briefly explain the strategy of the proof. First, we note the separation of jets along ZZ of ωXL\omega_{X}\otimes L is implied by the vanishing H1(X,ωX𝕃LIZ)=0,H^{1}(X,\omega_{X}\otimes^{\mathbb{L}}L\otimes I_{Z})=0, which is equivalent to the vanishing of Hom(LIZ,𝒪X[1])\operatorname{Hom}(L\otimes I_{Z},\mathscr{O}_{X}[1]). In the case of ωXDB\omega_{X}^{DB}, the object 𝒪X[1]\mathscr{O}_{X}[1] is replaced by Ω¯X00[1]\operatorname{\underline{\Omega}_{X}^{0}}0[1], and so we consider more generally the spaces of morphisms from objects of the form LIZL\otimes I_{Z} to a class of objects “of type OO” (see Definition 3.1 for details), which includes both 𝒪X[1]\mathscr{O}_{X}[1] and Ω¯X00[1]\operatorname{\underline{\Omega}_{X}^{0}}0[1]. As in Arcara-Bertram [1], the goal is to find Bridgeland stability conditions with respect to which objects of type OO (respectively, objects of the form LIZL\otimes I_{Z}) are stable, and then use Schur’s lemma to conclude. Refining the argument slightly, we see that in fact the image of a nonzero morphism ff of this form (appropriately defined) is a torsion sheaf whose support gives an effective divisor satisfying Reider-type conditions.

Outline of the paper. In Section 2, we fix notation and recall the necessary background on Bridgeland stability conditions on a (potentially singular) normal surface. We also briefly introduce Du Bois complexes and discuss some of their properties. In section 3 we find conditions guaranteeing that objects of the form LIZL\otimes I_{Z} and objects of type OO are Bridgeland stable, adapting the proof of Arcara and Bertram [1] to the more general setting of normal surfaces. We also explain why these results are stronger than we need for the proof of the main theorem, where it suffices to consider only destabilizing objects of a certain form. In section 4, we deduce the final form of our main technical theorem (Theorem 4.7), from which all our results follow.

Acknowledgements. We are grateful to Mihnea Popa for suggesting this problem and for insightful conversations throughout the project. We would also like to thank Davesh Maulik, Mircea Mustaţă, Sung Gi Park, and Wanchun Shen for valuable discussions.

2. Preliminaries

In this section, we review Langer’s construction of Bridgeland stability conditions on normal proper surfaces [13] (cf. [15]). We also give a brief introduction to Du Bois complexes. In particular, we focus on the zeroth Du Bois complex of a normal surface.

2.1. Bridgeland stability on normal surfaces

Let XX be a normal surface. Langer constructs Chern character homomorphisms

chiM:K0(X)A2i(X),\operatorname{ch}^{M}_{i}:K_{0}(X)\rightarrow A_{2-i}(X)_{\mathbb{Q}},

where ch0M\operatorname{ch}^{M}_{0} and ch1M\operatorname{ch}^{M}_{1} are defined as usual on the smooth locus and ch2M\operatorname{ch}^{M}_{2} is defined so that a Riemann–Roch formula holds (see [14] for details). We will simply denote the homomorphism by chi\operatorname{ch}_{i} in what follows. (This agrees with the standard Chern character on vector bundles, justifying the use of the notation.)

In addition to the Chern character, Langer’s definition uses the Mumford intersection product on A1(X)A_{1}(X) defined as follows [6, Ex. 7.1.16]: Let π:X~X\pi:\tilde{X}\rightarrow X be a minimal resolution of singularities. Then we define a homomorphism A1(X)A1(X~)A_{1}(X)\rightarrow A_{1}(\tilde{X})_{\mathbb{Q}} by

α=[C]α:=[C~]+Δ,\alpha=[C]\mapsto\alpha^{\prime}:=[\tilde{C}]+\Delta,

where CC is assumed to be an irreducible curve on XX, C~\tilde{C} is the proper transform of CC, and Δ\Delta is the unique \mathbb{Q}-divisor supported on the exceptional locus of π\pi such that for any component EE of the exceptional locus, we have αE=0\alpha^{\prime}\cdot E=0. For future use, we record the following observations:

Observations 2.1.
  1. (1)

    Writing sing(X)={p1,,pa}\operatorname{sing}(X)=\{p_{1},\ldots,p_{a}\} (as a normal surface has isolated singularities) and π1(pi)=Eij\pi^{-1}(p_{i})=\cup E_{ij} the set of irreducible components, then for Ni:=det((EijEik)j,k)N_{i}:=\det((E_{ij}\cdot E_{ik})_{j,k}) and N:=lcmiNiN:=\operatorname{lcm}_{i}N_{i}, we have Im(A1(X))1NA1(X~)\operatorname{Im}(A_{1}(X))\subset\frac{1}{N}\mathbb{Z}\otimes A_{1}(\tilde{X}).

  2. (2)

    If α\alpha is the class of a Cartier divisor DD, then α=[πD]A1(X~)\alpha^{\prime}=[\pi^{*}D]\in A_{1}(\tilde{X})_{\mathbb{Z}}.

  3. (3)

    Assuming from now on that XX is in addition proper, then the composition of the map A1(X)A1(X~)A_{1}(X)\rightarrow A_{1}(\tilde{X})_{\mathbb{Q}} with the standard intersection product A1(X~)A1(X~)A_{1}(\tilde{X})_{\mathbb{Q}}\otimes A_{1}(\tilde{X})_{\mathbb{Q}}\rightarrow\mathbb{Q} yields an intersection product on A1(X)A_{1}(X) satisfying the Hodge index theorem.

  4. (4)

    Given α,βA1(X)\alpha,\beta\in A_{1}(X) with α1MA1(X~)\alpha^{\prime}\in\frac{1}{M}\mathbb{Z}\otimes A_{1}(\tilde{X}), then αβ=αβ~1M\alpha\cdot\beta=\alpha^{\prime}\cdot\tilde{\beta}\in\frac{1}{M}\mathbb{Z}, as β=β~+Δ\beta^{\prime}=\tilde{\beta}+\Delta with αΔ=0\alpha^{\prime}\cdot\Delta=0. In particular, if α\alpha is the class of a Cartier divisor, then αβ\alpha\cdot\beta\in\mathbb{Z}. (More precisely, αβ\alpha\cdot\beta is given by composition of the degree map with the first Chern class intersection map of [6, Section 2.5]).

Now, let XX be a normal proper surface over an algebraically closed field, and choose \mathbb{R}-divisors ω,BA1(X)\omega,B\in A_{1}(X)_{\mathbb{R}} such that ω\omega is numerically ample. Langer defines a Bridgeland stability condition (Zω,B,Cohω,B)(Z_{\omega,B},\operatorname{Coh}_{\omega,B}) as follows: First, he shows that one can choose a constant CX0C_{X}\in\mathbb{R}_{\geq 0} satisfying the Bogomolov-type inequality

Xch1(E)22ch0(E)ch2(E)+CXch0(E)20\int_{X}\operatorname{ch}_{1}(E)^{2}-2\operatorname{ch}_{0}(E)\cdot\operatorname{ch}_{2}(E)+C_{X}\operatorname{ch}_{0}(E)^{2}\geq 0

for any torsion-free, ω\omega-slope semistable coherent sheaf EE on XX. (For example, if XX is smooth and defined over an algebraically closed field of characteristic 0, then by the ordinary Bogomolov inequality we may choose CX=0C_{X}=0. For more discussion, see 2.2)

Given the Bogomolov inequality on XX, one defines the central charge

Zω,B:K0(X),EXe(B+iω)ch(E)+CX2ch0(E).Z_{\omega,B}:K_{0}(X)\rightarrow\mathbb{C},E\mapsto-\int_{X}e^{-(B+i\omega)}\operatorname{ch}(E)+\frac{C_{X}}{2}\operatorname{ch}_{0}(E).

We note that this matches the standard definition of the central charge in the smooth case, assuming one takes CX=0C_{X}=0. The heart Cohω,B\operatorname{Coh}_{\omega,B} is defined exactly as in the smooth case, i.e., we define Cohω,B:=ω,B,𝒯ω,B\operatorname{Coh}_{\omega,B}:=\langle\mathcal{F}_{\omega,B},\mathcal{T}_{\omega,B}\rangle to be the tilt of Coh(X)\operatorname{Coh}(X) by the torsion pair

𝒯ω,B:={ECoh(X):all ω-slope HN factors are of slope μω>Bω}\mathcal{T}_{\omega,B}:=\{E\in\operatorname{Coh}(X):\textrm{all $\omega$-slope HN factors are of slope $\mu_{\omega}>B\cdot\omega$}\}
ω,B:={ECoh(X):all ω-slope HN factors are of slope μωBω},\mathcal{F}_{\omega,B}:=\{E\in\operatorname{Coh}(X):\textrm{all $\omega$-slope HN factors are of slope $\mu_{\omega}\leq B\cdot\omega$}\},

where the ω\omega-slope of a coherent sheaf EE is defined by

μω(E):=ch1(E)ωch0(E)(,]\mu_{\omega}(E):=\frac{\operatorname{ch}_{1}(E)\cdot\omega}{\operatorname{ch}_{0}(E)}\in(-\infty,\infty]

(setting μω(E)=\mu_{\omega}(E)=\infty when ch0(E)=0\operatorname{ch}_{0}(E)=0).

Recall that the central charge encodes the Bridgeland rank rω,B(E):=Zω,B(E)r_{\omega,B}(E):=\Im Z_{\omega,B}(E) and degree dω,B(E):=Zω,B(E)d_{\omega,B}(E):=-\Re Z_{\omega,B}(E). For E0Cohω,BE\neq 0\in\operatorname{Coh}_{\omega,B} we have rω,B(E)0r_{\omega,B}(E)\geq 0, and in the case of equality dω,B(E)>0d_{\omega,B}(E)>0. We then define the Bridgeland slope

νω,B(E):=dω,B(E)rω,B(E)(,].\nu_{\omega,B}(E):=\frac{d_{\omega,B}(E)}{r_{\omega,B}(E)}\in(-\infty,\infty].

An object of Cohω,B\operatorname{Coh}_{\omega,B} is said to be Bridgeland stable if it has Bridgeland slope greater than that of any proper subobject. Given Bridgeland stable objects FGCohω,BF\neq G\in\operatorname{Coh}_{\omega,B} with νω,B(F)νω,B(G)\nu_{\omega,B}(F)\geq\nu_{\omega,B}(G) we have the important property that Hom(F,G)=0\operatorname{Hom}(F,G)=0.

In this paper, we only consider the half-plane of stability conditions (ω,B)=(tH,sH)(\omega,B)=(tH,sH) for a fixed ample HA1(X)H\in A_{1}(X) and varying (t,s)>0×(t,s)\in\mathbb{R}^{>0}\times\mathbb{R}. We will use Zs,tZ_{s,t} to denote ZtH,sHZ_{tH,sH}, and similarly for rs,t,μs,t,r_{s,t},\mu_{s,t}, etc. We note that Cohs,t=:Cohs\operatorname{Coh}_{s,t}=:\operatorname{Coh}_{s} is in fact independent of the choice of t>0t\in\mathbb{R}^{>0}. For convenience, we record the following:

rs,t(E)=tH(ch1(E)sch0(E)H)r_{s,t}(E)=tH\cdot(\operatorname{ch}_{1}(E)-s\operatorname{ch}_{0}(E)H)
ds,t(E)=ch2(E)sch1(E)H+ch0(E)(s2t2)H2CX2d_{s,t}(E)=\operatorname{ch}_{2}(E)-s\operatorname{ch}_{1}(E)\cdot H+\operatorname{ch}_{0}(E)\frac{(s^{2}-t^{2})H^{2}-C_{X}}{2}

2.2. More about the constant CXC_{X}

Since the constant CXC_{X} appears in our Reider-type bounds, we will briefly recall its definition and compute it in an example.

First, it is not hard to reduce to showing the inequality for an arbitrary reflexive sheaf E.E. Let f:X~Xf:\tilde{X}\rightarrow X be a minimal resolution of XX, and set F:=(fE)F:=(f^{*}E)^{**}. Then one uses the fact that a Bogomolov inequality is satisfied for FF, i.e.,

X~Δ(F)CX~r2\int_{\tilde{X}}\Delta(F)\geq{-C_{\tilde{X}}r^{2}}

which is classical in the characteristic 0 (with CX~=0C_{\tilde{X}}=0) and a theorem of Koseki in positive characteristic [10].

Let c1(f,F)c_{1}(f,F) be the divisor supported on the exceptional locus of ff uniquely defined by the property that c1(f,F)Ei=FEic_{1}(f,F)\cdot E_{i}=F\cdot E_{i} for each irreducible exceptional divisor EiE_{i}. Langer calculates in [13] that

X~Δ(F)XΔ(E)=2rh0(X,R1fF)2r2h0(X,R1f𝒪X~)+c1(f,F)2rc1(f,F)KX~\int_{\tilde{X}}\Delta(F)-\int_{X}\Delta(E)=2rh^{0}(X,R^{1}f_{*}F)-2r^{2}h^{0}(X,R^{1}f_{*}\mathscr{O}_{\tilde{X}})+c_{1}(f,F)^{2}-rc_{1}(f,F)\cdot K_{\tilde{X}}

and so we get

XΔ(E)2rh0(X,R1fF)+2r2h0(X,R1f𝒪X~)c1(f,F)2+rc1(f,F)KX~CX~r2.\int_{X}\Delta(E)\geq-2rh^{0}(X,R^{1}f_{*}F)+2r^{2}h^{0}(X,R^{1}f_{*}\mathscr{O}_{\tilde{X}})-c_{1}(f,F)^{2}+rc_{1}(f,F)\cdot K_{\tilde{X}}{-C_{\tilde{X}}r^{2}}.

In [14], Langer shows that

h0(X,R1fF)(r+2)h0(X,R1f𝒪X~)h^{0}(X,R^{1}f_{*}F)\leq(r+2)h^{0}(X,R^{1}f_{*}\mathscr{O}_{\tilde{X}})

and so

2rh0(X,R1fF)+2r2h0(X,R1f𝒪S)4rh0(X,R1f𝒪X).-2rh^{0}(X,R^{1}f_{*}F)+2r^{2}h^{0}(X,R^{1}f_{*}\mathscr{O}_{S})\geq-4rh^{0}(X,R^{1}f_{*}\mathscr{O}_{X}).

Moreover, he calculates that for EjE_{j} a component of the exceptional divisor we have

0c1(f,F)Ej(r+2)h0(X,R1f𝒪X~)rχ(𝒪Ej)rEj2.0\leq c_{1}(f,F)\cdot E_{j}\leq(r+2)h^{0}(X,R^{1}f_{*}\mathscr{O}_{\tilde{X}})-r\chi(\mathscr{O}_{E_{j}})-rE_{j}^{2}.

Putting these together, one sees that we can bound XΔ(E)CXr2\int_{X}\Delta(E)\geq-C_{X}r^{2} for some CXC_{X} independent of EE (but depending on CX~,h0(X,R1f𝒪X~))C_{\tilde{X}},h^{0}(X,R^{1}f_{*}\mathscr{O}_{\tilde{X}})), and the genera and intersection numbers of the EjE_{j}’s).

Example 2.2.

Let CC be a degree 3 complex projective plane curve with line bundle L=𝒪2(1)|C.L=\mathcal{O}_{\mathbb{P}^{2}}(1)|_{C}. Then consider XX the projectivization of the cone over CC with conormal bundle LL, i.e. of the affine cone

Spec(m0H0(C,Lm)).\text{Spec}\left(\bigoplus_{m\geq 0}H^{0}(C,L^{m})\right).

Note that by [21, Appendix A] the cone has a Du Bois but not rational singularity. We would like to determine CXC_{X} in this case.

First, note that h0(R1f𝒪X~)=h1(𝒪C)=g(C)=1.h^{0}(R^{1}f_{*}\mathcal{O}_{\tilde{X}})=h^{1}(\mathcal{O}_{C})=g(C)=1. Moreover, EKX~=3E\cdot K_{\tilde{X}}=3 by adjunction. Now, suppose c1(f,F)=αE.c_{1}(f,F)=\alpha E. Then, by the above discussion, and since CX~=0C_{\tilde{X}}=0 by classical Bogomolov, one can choose CXC_{X} to be the smallest positive number such that

4r+3α2+3rαCXr2.-4r+3\alpha^{2}+3r\alpha\geq-C_{X}r^{2}.

Then 4r+3α2+3rα-4r+3\alpha^{2}+3r\alpha is minimized at α=r2\alpha=-\frac{r}{2} so we would like to find CXC_{X} such that 4r34r2+CXr20-4r-\frac{3}{4}r^{2}+C_{X}r^{2}\geq 0 for any integer r1.r\geq 1. This means CX=194.C_{X}=\frac{19}{4}.

Example 2.3.

One can similarly compute a bound for CXC_{X} for XX the cone over a smooth projective degree dd complex plane curve with conormal bundle L=O2(1)|C.L=O_{\mathbb{P}^{2}}(1)|_{C}. This bound grows asymptotically on the order of d3.d^{3}.

2.3. Du Bois complexes

Let XX be a complex variety. One would like to consider an analogue of the standard de Rham complex on smooth varieties for singular varieties. Fix a hyperresolution ε:XX.\varepsilon_{\bullet}:X_{\bullet}\rightarrow X. In [2], following ideas of Deligne, Du Bois introduced Ω¯X:=𝐑εΩXk,\underline{\Omega}_{X}^{\bullet}:=\mathbf{R}\varepsilon_{\bullet*}\Omega^{k}_{X_{\bullet}}, which is an object in the derived category of filtered differential complexes on XX, and showed that this is independent of the choice of hyperresolution. One can associate a filtration FpΩ¯X:=𝐑εΩXpF^{p}\underline{\Omega}_{X}^{\bullet}:=\mathbf{R}\varepsilon_{\bullet*}\Omega_{X_{\bullet}}^{\geq p} by recalling that ΩXi\Omega_{X_{i}}^{\bullet}is filtered by ΩXip.\Omega_{X_{i}}^{\geq p}. Consider then the pp-th Du Bois complex of XX

Ω¯Xp:=grFpΩ¯X[p].\underline{\Omega}_{X}^{p}:=\text{gr}_{F}^{p}\underline{\Omega}_{X}^{\bullet}[p].

For more details on Du Bois complexes, we refer the reader to [2], [8, Chapter V], or [16, Chapter 7.3].

In this paper, we will be concerned with understanding the stability of Ω¯X0,\underline{\Omega}_{X}^{0}, the 0-th Du Bois complex of X.X. A result of Saito and Schwede [18, Proposition 5.2], [20, Lemma 5.6] says that for any complex variety X,X,

0(Ω¯X00)=𝒪Xsn,\mathcal{H}^{0}(\operatorname{\underline{\Omega}_{X}^{0}}0)=\mathscr{O}_{X^{\text{sn}}},

where 𝒪Xsn\mathscr{O}_{X^{\text{sn}}} is the structure sheaf of the seminormalization of X.X.

In our case, since XX is normal, note that we simply have 0(Ω¯X00)=𝒪X.\mathcal{H}^{0}(\operatorname{\underline{\Omega}_{X}^{0}}0)=\mathscr{O}_{X}.

Moreover, for each p0p\geq 0, there is a canonical morphism ΩpkΩ¯Xp\Omega^{k}_{p}\rightarrow\underline{\Omega}_{X}^{p}, which is an isomorphism if XX is smooth; see [16, Page 175]. In particular, the i(Ω¯Xp)\mathcal{H}^{i}(\underline{\Omega}_{X}^{p}) are supported on the singular locus of X,X, for all i>0.i>0. Hence, in our case, 1(Ω¯X00)\mathcal{H}^{1}(\operatorname{\underline{\Omega}_{X}^{0}}0) is supported in codimension 2. General vanishing results for the cohomologies of Du Bois complexes show that i(Ω¯X0)=0\mathcal{H}^{i}(\underline{\Omega}_{X}^{0})=0 for i0,1.i\neq 0,1.

Note that any normal surface XX is Cohen-Macaulay (since it is S2S_{2}). Let ωX\omega_{X} be the dualizing sheaf of XX, and consider the duality functor 𝔻:=𝐑om(,ωX)\mathbb{D}:={\bf R}\mathcal{H}om(-,\omega_{X}) on Db(X).D^{b}(X). We define

ωXDB:=𝐑om(Ω¯X00,ωX).\omega_{X}^{DB}:={\bf R}\mathcal{H}om(\operatorname{\underline{\Omega}_{X}^{0}}0,\omega_{X}).

Now, by the injectivity theorem of Kovács-Schwede [11, Theorem 3.3], we have that the map

𝐑om(Ω¯X00,ωX)𝐑om(𝒪X,ωX),{\bf R}\mathcal{H}om(\operatorname{\underline{\Omega}_{X}^{0}}0,\omega_{X})\rightarrow{\bf R}\mathcal{H}om(\mathcal{O}_{X},\omega_{X}),

induced by dualizing the canonical morphism 𝒪XΩ¯X00,\mathcal{O}_{X}\rightarrow\operatorname{\underline{\Omega}_{X}^{0}}0, is injective on cohomology, i.e., xti(Ω¯X00,ωX)xti(𝒪X,ωX)\mathcal{E}xt^{i}(\operatorname{\underline{\Omega}_{X}^{0}}0,\omega_{X})\hookrightarrow\mathcal{E}xt^{i}(\mathscr{O}_{X},\omega_{X}) for all ii. In other words, ωXDB\omega_{X}^{DB} is a subsheaf of ωX\omega_{X}.

We conclude this section with a result that will be necessary in order to apply our main theorem.

Proposition 2.4.

For any closed point pXp\in X we have HomDb(X)(𝒪p,Ω¯X00[1])=0\operatorname{Hom}_{D^{b}(X)}(\mathscr{O}_{p},\operatorname{\underline{\Omega}_{X}^{0}}0[1])=0.

Proof.

We note first that 𝐑om(𝒪p,ωX)=𝒪p[2]{\bf R}\mathcal{H}om(\mathscr{O}_{p},\omega_{X})=\mathscr{O}_{p}[-2], as 𝐑om(𝒪p,ωX){\bf R}\mathcal{H}om(\mathscr{O}_{p},\omega_{X}) is supported on pp and so by the local-to-global spectral sequence it suffices to compute Exti(𝒪p,ωX)=H2i(X,𝒪p)\operatorname{Ext}^{i}(\mathscr{O}_{p},\omega_{X})=H^{2-i}(X,\mathscr{O}_{p})^{*}. Then

Hom(𝒪p,Ω¯X00[1])=Hom(𝒪p,𝐑om(𝔻(Ω¯X00),ωX)[1])=Hom(𝒪p𝕃𝔻(Ω¯X00),ωX[1])\operatorname{Hom}(\mathscr{O}_{p},\operatorname{\underline{\Omega}_{X}^{0}}0[1])=\operatorname{Hom}(\mathscr{O}_{p},{\bf R}\mathcal{H}om(\mathbb{D}(\operatorname{\underline{\Omega}_{X}^{0}}0),\omega_{X})[1])=\operatorname{Hom}(\mathscr{O}_{p}\otimes^{\mathbb{L}}\mathbb{D}(\operatorname{\underline{\Omega}_{X}^{0}}0),\omega_{X}[1])
=Hom(𝔻(Ω¯X00),𝐑om(𝒪p,ωX)[1])=Hom(𝔻(Ω¯X00),𝒪p[1])=0=\operatorname{Hom}(\mathbb{D}(\operatorname{\underline{\Omega}_{X}^{0}}0),{\bf R}\mathcal{H}om(\mathscr{O}_{p},\omega_{X})[1])=\operatorname{Hom}(\mathbb{D}(\operatorname{\underline{\Omega}_{X}^{0}}0),\mathscr{O}_{p}[-1])=0

by vanishing of negative Exts since 𝔻(Ω¯X00)\mathbb{D}(\operatorname{\underline{\Omega}_{X}^{0}}0) is a sheaf, as discussed above. ∎

We note that this proposition cannot be deduced purely from the cohomology sheaves i(Ω¯X00)\mathcal{H}^{i}(\operatorname{\underline{\Omega}_{X}^{0}}0) but requires the additional input that 𝔻(Ω¯X00)\mathbb{D}(\operatorname{\underline{\Omega}_{X}^{0}}0) is a sheaf; indeed, the statement would be false in the non-Du Bois case if Ω¯X00\operatorname{\underline{\Omega}_{X}^{0}}0 were replaced by 0(Ω¯X00)1(Ω¯X00)[1]\mathcal{H}^{0}(\operatorname{\underline{\Omega}_{X}^{0}}0)\oplus\mathcal{H}^{1}(\operatorname{\underline{\Omega}_{X}^{0}}0)[-1].

3. Bridgeland stability

In this section we prove several lemmas about Bridgeland stability of elements of Db(X)D^{b}(X) whose cohomologies match those of the objects we will consider. Namely, we study the stability of a class of objects with properties matching those of 𝒪X[1]\mathscr{O}_{X}[1] and Ω¯X00[1]\operatorname{\underline{\Omega}_{X}^{0}}0[1] (“type OO”) and of LZL\otimes\mathcal{I}_{Z} for LL an ample line bundle and ZXZ\subset X a subscheme of dimension 0. We also analyze the image of a morphism from an object LIZL\otimes I_{Z} to an object of type OO.

3.1. Conditions for stability

In this section, we derive inequalities that would be satisfied if an object of one of the two forms of interest were not Bridgeland stable, following the proof of [1] with the necessary modifications for the non-smooth case.

3.1.1. Objects of type OO

The following definition captures the essential properties of 𝒪X[1]\mathscr{O}_{X}[1] and Ω¯X00[1]\operatorname{\underline{\Omega}_{X}^{0}}0[1].

Definition 3.1.

We say an object FDb(X)F\in D^{b}(X) is of type OO if Hom(𝒪p,F)=0\operatorname{Hom}(\mathscr{O}_{p},F)=0 for every closed point pXp\in X and

i(F)={𝒪Xi=1𝒪T, for TX a zero-dimensional subschemei=00else.\mathcal{H}^{i}(F)=\begin{cases}\mathscr{O}_{X}&i=-1\\ \mathscr{O}_{T},\text{ for }T\subset X\textrm{ a zero-dimensional subscheme}&i=0\\ 0&\textrm{else}.\end{cases}
Example 3.2.

As explained in section 2.3, the object Ω¯X00[1]\operatorname{\underline{\Omega}_{X}^{0}}0[1] is of type OO. Because 𝐑om(𝒪X,ωX)=ωX{\bf R}\mathcal{H}om(\mathscr{O}_{X},\omega_{X})=\omega_{X} is also a sheaf, the argument of Proposition 2.4 shows that Hom(𝒪p,𝒪X[1])=0\operatorname{Hom}(\mathscr{O}_{p},\mathscr{O}_{X}[1])=0 for any pXp\in X, so 𝒪X[1]\mathscr{O}_{X}[1] is also of type OO. Another example is the shifted derived dual Z[1]\mathscr{I}_{Z}^{\vee}[1], where ZXZ\subset X is a 0-dimensional subscheme contained in the Gorenstein locus of XX; this is the type of object considered (in the smooth case) in [1].

Fix some numerically ample HA1(X)H\in A_{1}(X)_{\mathbb{R}}, and let FDb(X)F\in D^{b}(X) be an element of type OO. Our goal will be to find a choice of (t,s)(t,s) such that FF is a Bridgeland stable element of Cohs,t\operatorname{Coh}_{s,t}.

Proposition 3.3.

An object FF of type OO belongs to the heart Cohs,t\operatorname{Coh}_{s,t} if and only if s0s\geq 0.

Proof.

We note that 0(F)=𝒪T\mathcal{H}^{0}(F)=\mathscr{O}_{T} is torsion, therefore is contained in 𝒯s,t\mathcal{T}_{s,t} for any choice of ss and tt. So it suffices to see that 1(F)=𝒪Xs,t\mathcal{H}^{-1}(F)=\mathscr{O}_{X}\in\mathcal{F}_{s,t}. Since 𝒪X\mathscr{O}_{X} is of rank 1, and thus is automatically HH-stable, the necessary condition is that

tHsH=stH2μH(𝒪X)=ch1(𝒪X)tHch0(𝒪X)=0,tH\cdot sH=stH^{2}\geq\mu_{H}(\mathscr{O}_{X})=\frac{\operatorname{ch}_{1}(\mathscr{O}_{X})\cdot tH}{\operatorname{ch}_{0}(\mathscr{O}_{X})}=0,

and since necessarily t>0t>0, this amounts to the condition s0s\geq 0. ∎

We now explore the implications of FF being Bridgeland unstable.

Proposition 3.4.

Fix (t,s)>0×0(t,s)\in\mathbb{R}^{>0}\times\mathbb{R}^{\geq 0}. If FCohs,tF\in\operatorname{Coh}_{s,t} is destabilized by some quotient QQ^{\bullet}, then for any HH-slope HN factor EE of 1(Q)\mathcal{H}^{-1}(Q^{\bullet}) with Bridgeland rank rs,t(E)rs,t(F)r_{s,t}(E)\neq-r_{s,t}(F), we have that

(r1)sH2<AHrsH2,(r-1)sH^{2}<A\cdot H\leq rsH^{2},

where we write r:=c0(E)r:=c_{0}(E) and A:=ch1(E)A:=\operatorname{ch}_{1}(E).

Proof.

Suppose we have a destabilizing short exact sequence

0KFQ00\rightarrow K^{\bullet}\rightarrow F\rightarrow Q^{\bullet}\rightarrow 0

in Cohs,t\operatorname{Coh}_{s,t}. Then since Bridgeland rank is additive in short exact sequences and nonnegative on Cohs,t\operatorname{Coh}_{s,t}, we have that

0rs,t(E[1])rs,t(Q)rs,t(F)0\leq r_{s,t}(E[1])\leq r_{s,t}(Q^{\bullet})\leq r_{s,t}(F)

for EE as described above, as QQ^{\bullet} is a subobject of FF in Cohs,t\operatorname{Coh}_{s,t} and can be built up from a series of extensions of 0(Q)\mathcal{H}^{0}(Q^{\bullet}) and {Ei[1]}\{E_{i}[1]\}, where the EiE_{i}’s run over the HH-slope HN factors of 1(Q)\mathcal{H}^{-1}(Q^{\bullet}). Thus, using our assumption, we get

0rs,t(E[1])=tH(AsrH)<rs,t(F)=stH2srH2HA>s(r1)H2.0\leq r_{s,t}(E[1])=-tH\cdot(A-srH)<r_{s,t}(F)=stH^{2}\implies srH^{2}\geq H\cdot A>s(r-1)H^{2}.

Define lT:=length(0(F))l_{T}:=\operatorname{length}(\mathcal{H}^{0}(F)). We start by noting that ch(F)=ch(𝒪T)ch(𝒪X)\operatorname{ch}(F)=\operatorname{ch}(\mathscr{O}_{T})-\operatorname{ch}(\mathscr{O}_{X}), where 𝒪X\mathscr{O}_{X} is a line bundle and therefore Langer’s Chern character is the standard one. On the other hand, we have ch0(𝒪T)=ch1(𝒪T)=0\operatorname{ch}_{0}(\mathscr{O}_{T})=\operatorname{ch}_{1}(\mathscr{O}_{T})=0 and so Xch2(𝒪T)=χ(X,𝒪T)=lT\int_{X}\operatorname{ch}_{2}(\mathscr{O}_{T})=\chi(X,\mathscr{O}_{T})=l_{T} by [14].

In the following proposition, we will restrict attention to one particular stability condition, chosen so that s=12s=\frac{1}{2} and νs,t(F)=0\nu_{s,t}(F)=0 (taking l=lTl=l_{T}).

Proposition 3.5.

Suppose H2>4(2l+CX)H^{2}>4(2l+C_{X}), and fix s:=12,t:=142l+CXH2s:=\frac{1}{2},t:=\sqrt{\frac{1}{4}-\frac{2l+C_{X}}{H^{2}}}. Suppose EE is an HH-slope semistable sheaf with rs,t(E)0r_{s,t}(E)\leq 0 such that νs,t(E)0\nu_{s,t}(E)\leq 0. Then

AHA2r+r(CX+2l),A\cdot H\leq\frac{A^{2}}{r}+r(C_{X}+2l),

defining A:=ch1(E)A:=\operatorname{ch}_{1}(E) and r:=ch0(E)r:=\operatorname{ch}_{0}(E) as before.

Proof.

The slope inequality νs,t(E)0\nu_{s,t}(E)\leq 0 with denominator rs,t(E)0r_{s,t}(E)\leq 0 implies that the numerator

ds,t(E)=ch2(E)sAH+r(s2t2)H2CX20d_{s,t}(E)=\operatorname{ch}_{2}(E)-sA\cdot H+r\frac{(s^{2}-t^{2})H^{2}-C_{X}}{2}\geq 0
AH2ch2(E)+2rl.\implies A\cdot H\leq 2\operatorname{ch}_{2}(E)+2rl.

Now, by Langer’s Bogomolov-type inequality [13, Theorem 0.1], we have

2rch2(E)A2+CXr2AHA2r+CXr+2rl.2r\operatorname{ch}_{2}(E)\leq A^{2}+C_{X}r^{2}\implies A\cdot H\leq\frac{A^{2}}{r}+C_{X}r+2rl.

We now combine the previous two propositions to deduce a contradiction if HH is sufficiently positive.

Lemma 3.6.

Let H=c1(L),H=c_{1}(L), where LL is an ample line bundle such that H2>(CX+2l+1)2H^{2}>(C_{X}+2l+1)^{2}. Given AA1(X),r1A\in A_{1}(X),r\in\mathbb{Z}_{\geq 1} satisfying

r12H2<AHr2H2,AHA2r+r(CX+2l),\frac{r-1}{2}H^{2}<A\cdot H\leq\frac{r}{2}H^{2},\,\,A\cdot H\leq\frac{A^{2}}{r}+r(C_{X}+2l),

we must have either r=1r=1 and 0<AH<CX+2l+10<A\cdot H<C_{X}+2l+1, or r3r\geq 3 and H2<3(CX+2l+1)H^{2}<3(C_{X}+2l+1) (so in particular l=0l=0 and CX<2C_{X}<2). In the case r=1r=1, the same holds when HH is any numerically ample \mathbb{R}-divisor.

Proof.

Define D:=A/rD:=A/r. We start by showing that in the case r>1r>1 we must have D21D^{2}\geq 1. Indeed, suppose first that r3.r\geq 3. Then by our assumptions we have

D2+CX+2lDH>r12rH2H23>CX+2l+1D^{2}+C_{X}+2l\geq D\cdot H>\frac{r-1}{2r}H^{2}\geq\frac{H^{2}}{3}>C_{X}+2l+1

unless H23(CX+2l+1)H^{2}\leq 3(C_{X}+2l+1). When r=2r=2, we also use the assumption that HH is (integral) Cartier: recalling from 2.1(4) that H2H^{2} and AHA\cdot H are integers, we have

DH>H24>4(CX+2l)+14D\cdot H>\frac{H^{2}}{4}>\frac{4(C_{X}+2l)+1}{4}

which implies H24CX+2l+12\frac{H^{2}}{4}\geq C_{X}+2l+\frac{1}{2} and DHCX+2l+1,D\cdot H\geq C_{X}+2l+1, and so as above D21D^{2}\geq 1.

We now show that our assumption on H2H^{2} implies that D2<1D^{2}<1: recalling from 2.1(3) that the intersection product on XX satisfies the Hodge index theorem, we have

D2H2(DH)2(D2+CX+2l)H22D2CX+2lD^{2}H^{2}\leq(D\cdot H)^{2}\leq(D^{2}+C_{X}+2l)\frac{H^{2}}{2}\implies D^{2}\leq C_{X}+2l

and

D2H2(DH)2(D2+CX+2l)2H2(D2+CX+2lD2)2.D^{2}H^{2}\leq(D\cdot H)^{2}\leq(D^{2}+C_{X}+2l)^{2}\implies H^{2}\leq\left(\sqrt{D^{2}}+\frac{C_{X}+2l}{\sqrt{D^{2}}}\right)^{2}.

If we had D21D^{2}\geq 1, using that D2CX+2lD^{2}\leq C_{X}+2l we would conclude that

H2(CX+2l+1)2.H^{2}\leq(C_{X}+2l+1)^{2}.

However, this contradicts our assumption on HH, so we conclude that in fact D2<1D^{2}<1, which implies by the above that r=1.r=1.

It follows that

AHA2+CX+2l<CX+2l+1.A\cdot H\leq A^{2}+C_{X}+2l<C_{X}+2l+1.

We now describe conditions under which an object FF is Bridgeland stable.

Proposition 3.7.

Let FOX[1]Db(X)F\neq O_{X}[1]\in D^{b}(X) be of type OO, and suppose H=c1(L),H=c_{1}(L), where LL is an ample line bundle such that H2>(CX+2lT+1)2H^{2}>(C_{X}+2l_{T}+1)^{2} and HA1(X)MH\cdot A_{1}(X)\subset M\mathbb{Z} for some MCX+2lT+1M\geq C_{X}+2l_{T}+1. Then FF is Bridgeland stable for the stability condition (Zs,t,Cohs,t)(Z_{s,t},\operatorname{Coh}_{s,t}) with s=12,t=142lT+CXH2s=\frac{1}{2},t=\sqrt{\frac{1}{4}-\frac{2l_{T}+C_{X}}{H^{2}}}. (If F=𝒪X[1]F=\mathscr{O}_{X}[1], the same holds with the additional bound H23(CX+1)H^{2}\geq 3(C_{X}+1).)

Proof.

Suppose FCohs,tF\in\operatorname{Coh}_{s,t} is destabilized by a quotient QQ^{\bullet}, i.e., 0=νs,t(F)νs,t(Q)0=\nu_{s,t}(F)\geq\nu_{s,t}(Q^{\bullet}). Our strategy will be to find an HH-slope semistable factor EE of 1(Q)\mathcal{H}^{-1}(Q^{\bullet}) with rs,t(E)rs,t(F)r_{s,t}(E)\neq-r_{s,t}(F) and νs,t(E)νs,t(F)\nu_{s,t}(E)\leq\nu_{s,t}(F) and then to apply Lemma 3.6.

We start by showing that our assumption that Hom(𝒪p,F)\operatorname{Hom}(\mathscr{O}_{p},F) for every closed point pp implies rs,t(Q)<rs,t(F)r_{s,t}(Q^{\bullet})<r_{s,t}(F) strictly: Suppose rs,t(Q)=rs,t(F)r_{s,t}(Q^{\bullet})=r_{s,t}(F), and consider the kernel object KK^{\bullet}. We have the long exact sequence of cohomology

01(K)𝒪X1(Q)0(K)𝒪T0(Q)0,0\rightarrow\mathcal{H}^{-1}(K^{\bullet})\rightarrow\mathscr{O}_{X}\rightarrow\mathcal{H}^{-1}(Q^{\bullet})\rightarrow\mathcal{H}^{0}(K^{\bullet})\rightarrow\mathscr{O}_{T}\rightarrow\mathcal{H}^{0}(Q^{\bullet})\rightarrow 0,

where all the cohomologies 1\mathcal{H}^{-1} are by definition torsion free. This implies immediately that 1(K)0\mathcal{H}^{-1}(K^{\bullet})\cong 0 or 𝒪X\mathscr{O}_{X}, and the latter is impossible as in this case

rs,t(K)=0=rs,t(0(K))rs,t(1(K))r_{s,t}(K^{\bullet})=0=r_{s,t}(\mathcal{H}^{0}(K^{\bullet}))-r_{s,t}(\mathcal{H}^{-1}(K^{\bullet}))

where rs,t(0(K))0r_{s,t}(\mathcal{H}^{0}(K^{\bullet}))\geq 0 and rs,t(𝒪X)=stH2<0r_{s,t}(\mathscr{O}_{X})=-stH^{2}<0. So we conclude 1(K)=0\mathcal{H}^{-1}(K^{\bullet})=0, and then rs,t(0(K))=0r_{s,t}(\mathcal{H}^{0}(K^{\bullet}))=0 implies, by definition of Cohs,t\operatorname{Coh}_{s,t}, that 0(K)\mathcal{H}^{0}(K^{\bullet}) is supported in dimension 0 (as this can be decomposed into a torsion sheaf and torsion-free sheaves with rs,t>0r_{s,t}>0, and any torsion components supported in dimension 1 would pair positively with the ample divisor HH, again giving rs,t>0r_{s,t}>0). Now, restricting the inclusion KFK^{\bullet}\hookrightarrow F to any length one subsheaf 𝒪p0(K)\mathscr{O}_{p}\subset\mathcal{H}^{0}(K^{\bullet}) gives a nonzero element of Hom(𝒪p,F)\operatorname{Hom}(\mathscr{O}_{p},F), a contradiction. Thus we have

rs,t(E)=rs,t(E[1])rs,t(1(Q))=rs,t(Q)<rs,t(F)-r_{s,t}(E)=r_{s,t}(E[1])\leq r_{s,t}(\mathcal{H}^{-1}(Q^{\bullet}))=r_{s,t}(Q^{\bullet})<r_{s,t}(F)

for any semistable factor EE of 1(Q)\mathcal{H}^{-1}(Q^{\bullet}).

Now, recall that by definition of Cohs,t\operatorname{Coh}_{s,t} and the long exact sequence of cohomology, we have that 0(Q)\mathcal{H}^{0}(Q^{\bullet}) is a quotient of 0(F)=𝒪T\mathcal{H}^{0}(F)=\mathscr{O}_{T}, and 1(Q)\mathcal{H}^{-1}(Q^{\bullet}) has a filtration whose quotients are torsion-free HH-slope semistable sheaves EiE_{i} with rs,t(Ei)0r_{s,t}(E_{i})\leq 0. Then 0(Q)\mathcal{H}^{0}(Q^{\bullet}), being supported in dimension 0, has Bridgeland rank 0 and nonnegative degree, and so νs,t(1(Q))νs,t(Q)\nu_{s,t}(\mathcal{H}^{-1}(Q^{\bullet}))\leq\nu_{s,t}(Q^{\bullet}). Now, νs,t(1(Q))\nu_{s,t}(\mathcal{H}^{-1}(Q^{\bullet})) is a weighted average of the νs,t(Ei)\nu_{s,t}(E_{i}), and so at least one of these, say, EE, must have Bridgeland slope νs,t(E)νs,t(1(Q))νs,t(F)=0\nu_{s,t}(E)\leq\nu_{s,t}(\mathcal{H}^{-1}(Q^{\bullet}))\leq\nu_{s,t}(F)=0, and by the previous paragraph we have rs,t(E)rs,t(F)r_{s,t}(E)\neq-r_{s,t}(F).

Finally, we can apply Propositions 3.4 and 3.5 to conclude that r:=c0(E)r:=c_{0}(E) and A:=ch1(E)A:=\operatorname{ch}_{1}(E) satisfy

r12H2<AHr2H2,AHA2r+r(CX+2lT).\frac{r-1}{2}H^{2}<A\cdot H\leq\frac{r}{2}H^{2},\,\,A\cdot H\leq\frac{A^{2}}{r}+r(C_{X}+2l_{T}).

Then by Lemma 3.6 we get AHCX+2lT+1A\cdot H\leq C_{X}+2l_{T}+1, but this is impossible by our assumption that HA1(X)MH\cdot A_{1}(X)\subset M\mathbb{Z} with M>CX+2lT+1M>C_{X}+2l_{T}+1. ∎

In fact, to obtain the optimal bounds in our main theorem, we will use the following variant of the above proposition with a slightly weaker assumption on HH (which can now be an \mathbb{R}-divisor, and the inequality in the bounds is no longer strict). This small change turns out to be necessary to obtain the optimal Fujita-type bounds, at least in the smooth case.

Proposition 3.8.

Let FDb(X)F\in D^{b}(X) be of type OO and let HH be an numerically ample \mathbb{R}-divisor such that H2(CX+2lT+1)2H^{2}\geq(C_{X}+2l_{T}+1)^{2} and HCCX+2lT+1H\cdot C\geq C_{X}+2l_{T}+1 for any nonzero effective divisor CC. Then FF is not strictly destabilized by an injection from a torsion sheaf in cohomological degree 0 with respect to the Bridgeland stability condition (Zs,t,Cohs,t)(Z_{s,t},\operatorname{Coh}_{s,t}) for s=12,t=142lT+CXH2s=\frac{1}{2},t=\sqrt{\frac{1}{4}-\frac{2l_{T}+C_{X}}{H^{2}}} if t>0t>0 (i.e., in any case except lT=0,CX=1,H2=4l_{T}=0,C_{X}=1,H^{2}=4).

Proof.

Suppose KFK^{\bullet}\hookrightarrow F is a strictly destabilizing injection in Coh1/2\operatorname{Coh}_{1/2}, where K=0(K)=:KK^{\bullet}=\mathcal{H}^{0}(K^{\bullet})=:K is torsion. Let QQ^{\bullet} be the quotient object. Then by the long exact sequence of cohomology we conclude that E:=1(Q)E:=\mathcal{H}^{-1}(Q^{\bullet}) is of rank 1, so in particular this is HH-slope semistable, and since QQ^{\bullet} is a destabilizing quotient, we must have rs,t(E)0r_{s,t}(E)\neq 0 (or else, since 0(Q)\mathcal{H}^{0}(Q^{\bullet}) is supported in dimension 0, we would conclude that νs,t(Q)=>νs,t(F)\nu_{s,t}(Q^{\bullet})=\infty>\nu_{s,t}(F)).

For any ϵ>0\epsilon>0 define Hϵ:=(1+ϵ)HH_{\epsilon}:=(1+\epsilon)H, sϵ:=12s_{\epsilon}:=\frac{1}{2}, and tϵ:=142lT+CXHϵ2t_{\epsilon}:=\sqrt{\frac{1}{4}-\frac{2l_{T}+C_{X}}{H_{\epsilon}^{2}}}. We note that KCohsϵHϵK^{\bullet}\in\operatorname{Coh}_{s_{\epsilon}H_{\epsilon}} for all ϵ\epsilon, and QCohsϵHϵQ^{\bullet}\in\operatorname{Coh}_{s_{\epsilon}H_{\epsilon}} for ϵ\epsilon sufficiently small (where here we use that μtH(E)<stH2\mu_{tH}(E)<stH^{2} with strict inequality because rs,t(1(Q))0r_{s,t}(\mathcal{H}^{-1}(Q^{\bullet}))\neq 0 and thus the same holds for ϵ\epsilon sufficiently small). Thus the exact triangle KFQ+1K^{\bullet}\to F\to Q^{\bullet}\xrightarrow{+1} is contained in CohsϵHϵ\operatorname{Coh}_{s_{\epsilon}H_{\epsilon}} and so FQF\to Q^{\bullet} continues to be a surjection in Cohsϵ\operatorname{Coh}_{s_{\epsilon}} for ϵ\epsilon sufficiently small. Furthermore, since Bridgeland slope is a continuous function of ϵ\epsilon (since the denominators rs,t(F),rs,t(Q)0r_{s,t}(F),r_{s,t}(Q^{\bullet})\neq 0), we conclude that QQ^{\bullet} must be a strictly destabilizing quotient of FF for ϵ\epsilon sufficiently small.

Now, we apply the argument of the proof of Proposition 3.7: first, by definition of FF being of type OO we know that KK must be supported in dimension 1 (as this is torsion but not supported in dimension 0), and so in particular has positive Bridgeland rank. Also 0(Q)\mathcal{H}^{0}(Q^{\bullet}) is supported in dimension 0, so we have

0rsϵ,tϵ(Q)=rsϵ,tϵ(E)<rsϵ,tϵ(F)0\leq r_{s_{\epsilon},t_{\epsilon}}(Q^{\bullet})=-r_{s_{\epsilon},t_{\epsilon}}(E)<r_{s_{\epsilon},t_{\epsilon}}(F)

and also

νsϵ,tϵ(E)νsϵ,tϵ(Q)νsϵ,tϵ(F)=0.\nu_{s_{\epsilon},t_{\epsilon}}(E)\leq\nu_{s_{\epsilon},t_{\epsilon}}(Q^{\bullet})\leq\nu_{s_{\epsilon},t_{\epsilon}}(F)=0.

Thus we can apply Propositions 3.4 and 3.5 to end up in the situation of Lemma 3.6 with r=1r=1. Then we conclude ch1(E)H<CX+2l+1\operatorname{ch}_{1}(E)\cdot H<C_{X}+2l+1, a contradiction since ch1(E)\operatorname{ch}_{1}(E) is the effective divisor given by the support of KK. ∎

3.1.2. Objects of the form LZL\otimes\mathcal{I}_{Z}

Let LL be an ample line bundle and ZXZ\subset X be a subscheme of dimension 0. We write lZ:=length(Z)l_{Z}:=\operatorname{length}(Z) and fix H:=ch1(L)A1(X)H:=\operatorname{ch}_{1}(L)\in A_{1}(X). Our goal is to show that LZL\otimes\mathcal{I}_{Z} is Bridgeland stable for the stability condition s=12,t=142lZ+CXH2s=\frac{1}{2},t=\sqrt{\frac{1}{4}-\frac{2l_{Z}+C_{X}}{H^{2}}} considered in the last section (here with lZl_{Z} replacing lTl_{T}).

We note first of all that LZL\otimes\mathcal{I}_{Z} is HH-slope semistable as this is of rank 1, and so LZCohs,tL\otimes\mathcal{I}_{Z}\in\operatorname{Coh}_{s,t} exactly when

rs,t(LZ)=rs,t(L)=tH(HsH)>0,r_{s,t}(L\otimes\mathcal{I}_{Z})=r_{s,t}(L)=tH\cdot(H-sH)>0,

i.e., when s<1s<1. So in particular, LZCohs,tL\otimes\mathcal{I}_{Z}\in\operatorname{Coh}_{s,t} for our choice of s=12s=\frac{1}{2}.

Proposition 3.9.

Let LL, ZZ, ss, and tt be as above, and suppose H:=ch1(L)H:=\operatorname{ch}_{1}(L) satisfies H2>4(2lZ+CX)H^{2}>4(2l_{Z}+C_{X}) and HC2(2lZ+CX)H\cdot C\geq 2(2l_{Z}+C_{X}) for any nonzero effective divisor CC. Then LIZCohs,tL\otimes I_{Z}\in\operatorname{Coh}_{s,t} is not destabilized by a subsheaf LIY,ZYXL\otimes I_{Y},Z\subsetneq Y\subsetneq X.

Proof.

Let Y1A1(X)Y_{1}\in A_{1}(X) be dimension 1 component of [Y]A(X)[Y]\in A_{*}(X). We note immediately that if Y1=0Y_{1}=0, then the sheaf LIYL\otimes I_{Y} is not destabilizing, as rs,t(LIY)=rs,t(LIZ)0r_{s,t}(L\otimes I_{Y})=r_{s,t}(L\otimes I_{Z})\neq 0 and ds,t(LIY)=ds,t(LIZ)length(IZ/IY)d_{s,t}(L\otimes I_{Y})=d_{s,t}(L\otimes I_{Z})-\operatorname{length}(I_{Z}/I_{Y}).

Then since LIYL\otimes I_{Y} is of rank 1 and thus HH-slope semistable, we see that LIYCohs,tL\otimes I_{Y}\in\operatorname{Coh}_{s,t} if and only if

rs,t(LIY)>0H2>2HY1>0r_{s,t}(L\otimes I_{Y})>0\iff H^{2}>2H\cdot Y_{1}>0

where the second inequality follows from the fact that Y10Y_{1}\neq 0 is effective.

Now, for LIYCohs,tL\otimes I_{Y}\in\operatorname{Coh}_{s,t} to be a destabilizing subsheaf we must have

νs,t(LIY)νs,t(LIZ)=0,\nu_{s,t}(L\otimes I_{Y})\geq\nu_{s,t}(L\otimes I_{Z})=0,

and since rs,t(LIY)>0r_{s,t}(L\otimes I_{Y})>0 this means

ds,t(LIY)=ch2(LIY)12(HY1)H+lZ0.d_{s,t}(L\otimes I_{Y})=\operatorname{ch}_{2}(L\otimes I_{Y})-\frac{1}{2}(H-Y_{1})\cdot H+l_{Z}\geq 0.

By Langer’s Bogomolov inequality [13, Theorem 0.1] we have

2ch2(LIY)(HY1)2+CX2\operatorname{ch}_{2}(L\otimes I_{Y})\leq(H-Y_{1})^{2}+C_{X}

and so this would imply

(HY1)2(HY1)H+CX+2lZ0.(H-Y_{1})^{2}-(H-Y_{1})\cdot H+C_{X}+2l_{Z}\geq 0.

In other words, to prove our result, it suffices to show

CX+2lZ+Y12<Y1H.C_{X}+2l_{Z}+Y_{1}^{2}<Y_{1}\cdot H.

It follows from the Hodge index theorem (see 2.1(3)) and the condition that LIYCohs,tL\otimes I_{Y}\in\operatorname{Coh}_{s,t} above that

Y12(Y1H)2H2<Y1H2.Y_{1}^{2}\leq\frac{(Y_{1}\cdot H)^{2}}{H^{2}}<\frac{Y_{1}\cdot H}{2}.

Now since Y1Y_{1} is effective, by assumption we have

CX+2lZY1H2.C_{X}+2l_{Z}\leq\frac{Y_{1}\cdot H}{2}.

Putting the two together

CX+2lZ+Y12<Y1H,C_{X}+2l_{Z}+Y_{1}^{2}<Y_{1}\cdot H,

completing the proof. ∎

Proposition 3.10.

Let LL, ZZ, ss, tt be as above, and suppose H:=ch1(L)H:=\operatorname{ch}_{1}(L) satisfies H2>(CX+2lZ+1)2H^{2}>(C_{X}+2l_{Z}+1)^{2} and HA1(X)MH\cdot A_{1}(X)\subset M\mathbb{Z} for some Mmax(2(CX+2lZ),CX+1)M\geq\max(2(C_{X}+2l_{Z}),C_{X}+1). Then LIZCohs,tL\otimes I_{Z}\in\operatorname{Coh}_{s,t} is stable with respect to the Bridgeland stability condition (Zs,t,Cohs,t)(Z_{s,t},\operatorname{Coh}_{s,t}).

Proof.

Suppose that LIZL\otimes I_{Z} is destabilized by some quotient QQ^{\bullet}, i.e., that 0=νs,t(LIZ)νs,t(Q)0=\nu_{s,t}(L\otimes I_{Z})\geq\nu_{s,t}(Q^{\bullet}). Using the short exact sequence in Cohs,t\operatorname{Coh}_{s,t} given by the exact triangle

1(Q)[1]Q0(Q)+1\mathcal{H}^{-1}(Q^{\bullet})[1]\rightarrow Q^{\bullet}\rightarrow\mathcal{H}^{0}(Q^{\bullet})\xrightarrow{+1}

we conclude that at least one of νs,t(1(Q)[1])=νs,t(1(Q))\nu_{s,t}(\mathcal{H}^{-1}(Q^{\bullet})[1])=\nu_{s,t}(\mathcal{H}^{-1}(Q^{\bullet})) and νs,t(0(Q))\nu_{s,t}(\mathcal{H}^{0}(Q^{\bullet})) satisfies this inequality as well.

If νs,t(0(Q))0\nu_{s,t}(\mathcal{H}^{0}(Q^{\bullet}))\leq 0, then LIZQ0(Q)L\otimes I_{Z}\twoheadrightarrow Q^{\bullet}\twoheadrightarrow\mathcal{H}^{0}(Q^{\bullet}) is a destabilizing quotient whose kernel is a subsheaf of LIZL\otimes I_{Z}, but this is impossible by Proposition 3.9.

On the other hand, if this is not the case, then νs,t(1(Q))0\nu_{s,t}(\mathcal{H}^{-1}(Q^{\bullet}))\leq 0 and 1(Q)0\mathcal{H}^{-1}(Q^{\bullet})\neq 0, and we proceed as in the proof of Proposition 3.7: we may choose an HH-slope stable factor EE of 1(Q)\mathcal{H}^{-1}(Q^{\bullet}) with νs,t(E)νs,t(1(Q))0\nu_{s,t}(E)\leq\nu_{s,t}(\mathcal{H}^{-1}(Q^{\bullet}))\leq 0 and rs,t(E)0r_{s,t}(E)\leq 0, so that Proposition 3.5 applies. Furthermore, since 1(Q)0\mathcal{H}^{-1}(Q^{\bullet})\neq 0 and thus ch0(1(Q))>0\operatorname{ch}_{0}(\mathcal{H}^{-1}(Q^{\bullet}))>0, the same is true of 0(K)\mathcal{H}^{0}(K^{\bullet}), letting K=0(K)K^{\bullet}=\mathcal{H}^{0}(K^{\bullet}) be the kernel object. Then we have rs,t(K)>0r_{s,t}(K^{\bullet})>0 by definition of 𝒯s,t\mathcal{T}_{s,t}, and thus the inequality rs,t(Q)<rs,t(LIZ)r_{s,t}(Q^{\bullet})<r_{s,t}(L\otimes I_{Z}) is strict. In other words, we get

0rs,t(E[1])=rs,t(E)<rs,t(LIZ)r12H2<AHr2H2,0\leq r_{s,t}(E[1])=-r_{s,t}(E)<r_{s,t}(L\otimes I_{Z})\implies\frac{r-1}{2}H^{2}<A\cdot H\leq\frac{r}{2}H^{2},

writing r:=ch0(E)r:=\operatorname{ch}_{0}(E) and A:=ch1(E)A:=\operatorname{ch}_{1}(E) as before. Then by Lemma 3.6 we conclude that 0<AH<CX+2lZ+10<A\cdot H<C_{X}+2l_{Z}+1, but by our assumption on HA1(X)H\cdot A_{1}(X) this is impossible. (We have CX+2lZ+12(CX+2lZ)C_{X}+2l_{Z}+1\leq 2(C_{X}+2l_{Z}) unless lZ=0,CX<1l_{Z}=0,C_{X}<1, in which case the correct bound is CX+1C_{X}+1.)

3.2. Reduction to the torsion sheaf case

As usual, we fix a projective normal surface XX, an ample line bundle LL on XX, a finite length subscheme ZZ of length lZl_{Z}, and an object FF of type OO with length(0(F))=lT\operatorname{length}(\mathcal{H}^{0}(F))=l_{T}. Setting H:=c1(L)H:=c_{1}(L), recall that LIZ,FL\otimes I_{Z},F belong to the heart Coh1/2\operatorname{Coh}_{1/2} defined in Section 2. In particular, as this is a full abelian subcategory of Db(X)D^{b}(X), it makes sense to talk about the image of a homomorphism fHom(LIZ,F)f\in\operatorname{Hom}(L\otimes I_{Z},F).

Proposition 3.11.

Let X,L,Z,F,lZ,lTX,L,Z,F,l_{Z},l_{T} be as above, and suppose H2>4(lZ+lT+CX)H^{2}>4(l_{Z}+l_{T}+C_{X}). Then given f0Hom(LIZ,F)f\neq 0\in\operatorname{Hom}(L\otimes I_{Z},F) we have that Imf=0(Imf)\operatorname{Im}f=\mathcal{H}^{0}(\operatorname{Im}f) is a torsion sheaf.

Proof.

Let G:=ImfG:=\operatorname{Im}f. We use the fact that LIZGFL\otimes I_{Z}\twoheadrightarrow G\hookrightarrow F, combined with the definition of Coh1/2\operatorname{Coh}_{1/2}.

First, since LIZGL\otimes I_{Z}\twoheadrightarrow G, the corresponding short exact sequence in Coh1/2\operatorname{Coh}_{1/2} gives a long exact sequence of cohomology

01(G)0(kerf)LIZ0(G)00\to\mathcal{H}^{-1}(G)\to\mathcal{H}^{0}(\ker f)\to L\otimes I_{Z}\to\mathcal{H}^{0}(G)\to 0

which implies that c0(0(G))1c_{0}(\mathcal{H}^{0}(G))\leq 1 with equality if and only if 0(G)=LIZ\mathcal{H}^{0}(G)=L\otimes I_{Z} (as LIZL\otimes I_{Z} is torsion-free of rank 1). Furthermore, in this case 1(G)=0(kerf)=0\mathcal{H}^{-1}(G)=\mathcal{H}^{0}(\ker f)=0 by definition of Coh1/2\operatorname{Coh}_{1/2}, so we conclude that either c0(0(G))=0c_{0}(\mathcal{H}^{0}(G))=0 or G=LIZG=L\otimes I_{Z}. We claim the latter is impossible: indeed, since

r1/2,t(LIZ)=r1/2,t(F)=t2H2r_{1/2,t}(L\otimes I_{Z})=r_{1/2,t}(F)=\frac{t}{2}H^{2}

we conclude that the quotient cokerf0\operatorname{coker}f\neq 0 must satisfy

d1/2,t(cokerf)=d1/2,t(F)d1/2,t(LIZ)>0d_{1/2,t}(\operatorname{coker}f)=d_{1/2,t}(F)-d_{1/2,t}(L\otimes I_{Z})>0

for all t>0t>0. However, our assumption on H2H^{2} implies that this is false for tt sufficiently small. Thus c0(0(G))=0c_{0}(\mathcal{H}^{0}(G))=0.

Second, since GFG\hookrightarrow F, we have the long exact sequence

01(G)𝒪X1(cokerf)0(G)0(F)0(cokerf)0.0\to\mathcal{H}^{-1}(G)\to\mathscr{O}_{X}\to\mathcal{H}^{-1}(\operatorname{coker}f)\to\mathcal{H}^{0}(G)\to\mathcal{H}^{0}(F)\to\mathcal{H}^{0}(\operatorname{coker}f)\to 0.

Since 𝒪X\mathscr{O}_{X} and 𝒪X/1(G)1(cokerf)\mathscr{O}_{X}/\mathcal{H}^{-1}(G)\hookrightarrow\mathcal{H}^{-1}(\operatorname{coker}f) are torsion-free, we conclude that either 1(G)=0\mathcal{H}^{-1}(G)=0 or 1(G)=𝒪X\mathcal{H}^{-1}(G)=\mathscr{O}_{X}. In the latter case, the fact that 1(cokerf)\mathcal{H}^{-1}(\operatorname{coker}f) is torsion-free and injects into 0(G)\mathcal{H}^{0}(G), which is of rank 0 by the above, implies that 1(cokerf)=0\mathcal{H}^{-1}(\operatorname{coker}f)=0, so in particular 0(G)\mathcal{H}^{0}(G) is a subsheaf of 0(F)\mathcal{H}^{0}(F), therefore of dimension 0. But then we see that

μtH(0(kerf))=tHch1(𝒪X)+ch1(LIZ)2=tH22=H2tH\mu_{tH}(\mathcal{H}^{0}(\ker f))=tH\cdot\frac{\operatorname{ch}_{1}(\mathscr{O}_{X})+\operatorname{ch}_{1}(L\otimes I_{Z})}{2}=\frac{tH^{2}}{2}=\frac{H}{2}\cdot tH

which is impossible by definition of Coh1/2\operatorname{Coh}_{1/2}. So in fact 1(G)=0\mathcal{H}^{-1}(G)=0. ∎

4. Proof of the main results

We start by translating our separation of jets statement into a form that can be more effectively approached using Bridgeland stability.

Proposition 4.1.

Let XX be a projective normal surface, and GDb(X)G\in D^{b}(X). We have

H1(X,ωXDB𝕃G)=Hom(G,Ω¯X00[1]).H^{1}(X,\omega_{X}^{DB}\otimes^{\mathbb{L}}G)=\operatorname{Hom}(G,\operatorname{\underline{\Omega}_{X}^{0}}0[1])^{*}.
Proof.

We have

H1(X,ωXDB𝕃G)=H1(X,𝐑om(Ω¯X00,ωX)𝕃G)=H1(X,𝐑om(𝐑om(G,Ω¯X00),ωX))H^{1}(X,\omega_{X}^{DB}\otimes^{\mathbb{L}}G)=H^{1}(X,{\bf R}\mathcal{H}om(\operatorname{\underline{\Omega}_{X}^{0}}0,\omega_{X})\otimes^{\mathbb{L}}G)=H^{1}(X,{\bf R}\mathcal{H}om({\bf R}\mathcal{H}om(G,\operatorname{\underline{\Omega}_{X}^{0}}0),\omega_{X}))

by [22, Tag 0G4I], and since 𝔻:=𝐑om(,ωX)\mathbb{D}:={\bf R}\mathcal{H}om(-,\omega_{X}) gives an involution of Db(X)D^{b}(X), we get

=Hom(𝒪X,𝐑om(𝐑om(G,Ω¯X00),ωX)[1])=Hom(𝐑om(G,Ω¯X00),ωX[1])=\operatorname{Hom}(\mathscr{O}_{X},{\bf R}\mathcal{H}om({\bf R}\mathcal{H}om(G,\operatorname{\underline{\Omega}_{X}^{0}}0),\omega_{X})[1])=\operatorname{Hom}({\bf R}\mathcal{H}om(G,\operatorname{\underline{\Omega}_{X}^{0}}0),\omega_{X}[1])
=Hom(𝒪X,𝐑om(G,Ω¯X00)[1])=Hom(G,Ω¯X00[1]).=\operatorname{Hom}(\mathscr{O}_{X},{\bf R}\mathcal{H}om(G,\operatorname{\underline{\Omega}_{X}^{0}}0)[1])^{*}=\operatorname{Hom}(G,\operatorname{\underline{\Omega}_{X}^{0}}0[1])^{*}.

Recall we are interested in a bound for aa such that H1(X,ωXDB𝕃LaIZ)=0.H^{1}(X,\omega_{X}^{DB}\otimes^{\mathbb{L}}L^{\otimes a}\otimes I_{Z})=0. In light of Proposition 4.1, our strategy will be to use Bridgeland stability conditions in order to show Hom(LaIZ,Ω¯X00[1])=0.\operatorname{Hom}(L^{\otimes a}\otimes I_{Z},\operatorname{\underline{\Omega}_{X}^{0}}0[1])=0. More generally, we consider the vanishing of Hom(LIZ,F)\operatorname{Hom}(L\otimes I_{Z},F) where LL is an ample line bundle, ZXZ\subset X a subscheme of dimension 0, and FDb(X)F\in D^{b}(X) is of type OO. One advantage of this more general setup is that we can obtain better bounds by relating Hom(LIZ,F)\operatorname{Hom}(L\otimes I_{Z},F) to other spaces Hom(LIZ,F)\operatorname{Hom}(L\otimes I_{Z^{\prime}},F^{\prime}) given by changing the relative lengths of ZZ and 0(F)\mathcal{H}^{0}(F). In what follows, we use the notation lZ:=length(Z)l_{Z}:=\operatorname{length}(Z) and lF:=length(0(F))l_{F}:=\operatorname{length}(\mathcal{H}^{0}(F)), and similarly for ZZ^{\prime} and FF^{\prime}.

Proposition 4.2.

Let L,Z,FL,Z,F be as above, with lF>0l_{F}>0. Then if Hom(LIZ,F)0\operatorname{Hom}(L\otimes I_{Z},F)\neq 0 we can find a dimension 0 subscheme ZZ^{\prime} and FF^{\prime} of type OO with Hom(LIZ,F)0\operatorname{Hom}(L\otimes I_{Z^{\prime}},F^{\prime})\neq 0, lZlZ+1l_{Z^{\prime}}\leq l_{Z}+1, and lF=lF1l_{F^{\prime}}=l_{F}-1.

Proof.

Choose a surjection 0(F)𝒪p\mathcal{H}^{0}(F)\twoheadrightarrow\mathscr{O}_{p} and some f0Hom(LIZ,F)f\neq 0\in\operatorname{Hom}(L\otimes I_{Z},F). Then consider the map

g:LIZ𝑓F0(F)𝒪p.g:L\otimes I_{Z}\xrightarrow{f}F\twoheadrightarrow\mathcal{H}^{0}(F)\twoheadrightarrow\mathscr{O}_{p}.

Since LIZ,𝒪pL\otimes I_{Z},\mathscr{O}_{p} belong to the heart Coh1/2\operatorname{Coh}_{1/2}, we can define ker(g)Coh1/2\ker(g)\in\operatorname{Coh}_{1/2}; similarly, we define F:=ker(F0(F)𝒪p)Coh1/2F^{\prime}:=\ker(F\twoheadrightarrow\mathcal{H}^{0}(F)\twoheadrightarrow\mathscr{O}_{p})\in\operatorname{Coh}_{1/2}. It is immediate that FF^{\prime} is of type OO with lF=lF1l_{F^{\prime}}=l_{F}-1, and as for ker(g)\ker(g), we note that either g=0g=0, in which case ker(g)=LIZ\ker(g)=L\otimes I_{Z} and we take Z:=ZZ^{\prime}:=Z, or g0g\neq 0 and is thus surjective, in which case ker(g)=LIZ\ker(g)=L\otimes I_{Z^{\prime}} for some ZZZ^{\prime}\supset Z with lZ=lZ+1l_{Z}^{\prime}=l_{Z}+1 by the long exact sequence of cohomology.

We note that the restriction of ff to LIZL\otimes I_{Z^{\prime}} is nontrivial (because Hom(𝒪p,F)=0\operatorname{Hom}(\mathscr{O}_{p},F)=0 by assumption) and factors through FFF^{\prime}\hookrightarrow F, and so we find an element f0Hom(LIZ,F)f^{\prime}\neq 0\in\operatorname{Hom}(L\otimes I_{Z^{\prime}},F^{\prime}). ∎

Proposition 4.3.

Let L,Z,FL,Z,F be as above, with lZ>0l_{Z}>0. Then if Hom(LIZ,F)0\operatorname{Hom}(L\otimes I_{Z},F)\neq 0 we can find a dimension 0 subscheme ZZ^{\prime} and FF^{\prime} of type OO with Hom(LIZ,F)0\operatorname{Hom}(L\otimes I_{Z^{\prime}},F^{\prime})\neq 0, lZ=lZ1l_{Z^{\prime}}=l_{Z}-1, and lFlF+1l_{F^{\prime}}\leq l_{F}+1.

Proof.

Choose a subscheme ZZZ^{\prime}\subset Z with lZ=lZ1l_{Z^{\prime}}=l_{Z}-1, and let 𝒪p:=IZ/IZ\mathscr{O}_{p}:=I_{Z^{\prime}}/I_{Z}. Then

LIZLIZ𝒪p+1L\otimes I_{Z}\rightarrow L\otimes I_{Z^{\prime}}\rightarrow\mathscr{O}_{p}\xrightarrow{+1}

gives a map Hom(LIZ,F)Hom(𝒪p[1],F)\operatorname{Hom}(L\otimes I_{Z},F)\rightarrow\operatorname{Hom}(\mathscr{O}_{p}[-1],F). Choose f0Hom(LIZ,F)f\neq 0\in\operatorname{Hom}(L\otimes I_{Z},F), and let gHom(𝒪p[1],F)g\in\operatorname{Hom}(\mathscr{O}_{p}[-1],F) be its image. If g=0g=0, then ff is the image of some f0Hom(LIZ,F)f^{\prime}\neq 0\in\operatorname{Hom}(L\otimes I_{Z^{\prime}},F), and so for F:=FF^{\prime}:=F we get Hom(LIZ,F)0\operatorname{Hom}(L\otimes I_{Z^{\prime}},F^{\prime})\neq 0.

Otherwise, let F:=Cone(g)F^{\prime}:=\operatorname{Cone}(g). By chasing the diagram

idHom(𝒪p[1],𝒪p[1]){id\in\operatorname{Hom}(\mathscr{O}_{p}[-1],\mathscr{O}_{p}[-1])}fHom(LIZ,F){f\in\operatorname{Hom}(L\otimes I_{Z},F)}gHom(𝒪p[1],F){g\in\operatorname{Hom}(\mathscr{O}_{p}[-1],F)}fHom(LIZ,F){f^{\prime}\in\operatorname{Hom}(L\otimes I_{Z^{\prime}},F^{\prime})}f′′Hom(LIZ,F){f^{\prime\prime}\in\operatorname{Hom}(L\otimes I_{Z},F^{\prime})}0Hom(𝒪p[1],F){0\in\operatorname{Hom}(\mathscr{O}_{p}[-1],F^{\prime})}

we find a morphism f0Hom(LIZ,F)f^{\prime}\neq 0\in\operatorname{Hom}(L\otimes I_{Z^{\prime}},F^{\prime}) whose restriction f′′f^{\prime\prime} to LIZL\otimes I_{Z} is given by ff. Furthermore, we can see the cohomology groups of FF^{\prime} are of the correct form, with lF=lF+1l_{F^{\prime}}=l_{F}+1, and Hom(𝒪q,F)=0\operatorname{Hom}(\mathscr{O}_{q},F^{\prime})=0 for every closed point qq: for qpq\neq p this is clear from the fact that Exti(𝒪q,𝒪p)=0\operatorname{Ext}^{i}(\mathscr{O}_{q},\mathscr{O}_{p})=0 i\forall i, and for q=pq=p it follows from the long exact sequence

0=Hom(𝒪p,F)Hom(𝒪p,F)Hom(𝒪p,𝒪p)g[1]Hom(𝒪p,F[1])0=\operatorname{Hom}(\mathscr{O}_{p},F)\rightarrow\operatorname{Hom}(\mathscr{O}_{p},F^{\prime})\rightarrow\operatorname{Hom}(\mathscr{O}_{p},\mathscr{O}_{p})\xhookrightarrow{g[1]\circ-}\operatorname{Hom}(\mathscr{O}_{p},F[1])\rightarrow\cdots

Thus FF^{\prime} is of type OO. ∎

Remark 4.4.

For the purposes of the main theorem, it is convenient to assume that decreasing lFl_{F} by 1 will have the effect of increasing lZl_{Z} by 1 and vice versa, while the preceding propositions leave open the possibility that in fact lZl_{Z} may not need to increase if f:LIZFf:L\otimes I_{Z}\to F already factors through the chosen FFF^{\prime}\subset F, and similarly lFl_{F} may not need to increase if f:LIZFf:L\otimes I_{Z}\to F can be extended to LIZL\otimes I_{Z}^{\prime} for the chosen ZZZ^{\prime}\subsetneq Z. However, we note that it is always possible to arrange for lZ+lF=lZ+lFl_{Z}+l_{F}=l_{Z^{\prime}}+l_{F^{\prime}} to remain constant: in the first case, one takes an arbitrary ZZZ^{\prime}\supset Z with lZ=lZ+1l_{Z^{\prime}}=l_{Z}+1 and restricts LIZFL\otimes I_{Z}\to F^{\prime} to LIZL\otimes I_{Z^{\prime}}, and in the second case, one chooses an arbitrary point pXsupp(0(F))p\in X\setminus\operatorname{supp}(\mathcal{H}^{0}(F)) and takes FF^{\prime} to be the cone of an arbitrary nonzero element of Hom(𝒪p[1],F)\operatorname{Hom}(\mathscr{O}_{p}[-1],F) and extends LIZFL\otimes I_{Z^{\prime}}\to F by FFF\hookrightarrow F^{\prime}. (Note that by assumption on pp, we have Hom(𝒪p[1],F)=Hom(𝒪p[1],𝒪[1])=Hom(𝔻(𝒪[2]),𝔻(𝒪p))=Hom(ωX,𝒪p)0\operatorname{Hom}(\mathscr{O}_{p}[-1],F)=\operatorname{Hom}(\mathscr{O}_{p}[-1],\mathscr{O}[1])=\operatorname{Hom}(\mathbb{D}(\mathscr{O}[2]),\mathbb{D}(\mathscr{O}_{p}))=\operatorname{Hom}(\omega_{X},\mathscr{O}_{p})\neq 0.)

Before proving our main results, we note that a priori our methods give the vanishing of H1(X,ωX𝕃(LIZ))=0,H^{1}(X,\omega_{X}\otimes^{\mathbb{L}}(L\otimes I_{Z}))=0, where the tensor products are derived. To obtain Reider-type results one would of course like to have the subjectivity of the map H0(X,ωXL)H0(X,ωXL𝒪Z).H^{0}(X,\omega_{X}\otimes L)\rightarrow H^{0}(X,\omega_{X}\otimes L\otimes\mathcal{O}_{Z}). This is implied by the following result:

Proposition 4.5.

Let XX be a normal surface, GCoh(X)G\in\operatorname{Coh}(X) a coherent sheaf and ZXZ\subset X a 0-dimensional subscheme. Then the vanishing H1(X,G𝕃IZ)=0H^{1}(X,G\otimes^{\mathbb{L}}I_{Z})=0 implies the surjectivity of H0(X,G)H0(X,G𝒪Z).H^{0}(X,G)\rightarrow H^{0}(X,G\otimes\mathcal{O}_{Z}).

Proof.

The vanishing H1(X,G𝕃IZ)=0H^{1}(X,G\otimes^{\mathbb{L}}I_{Z})=0 implies the map H0(X,G)H0(X,G𝕃𝒪Z)H^{0}(X,G)\rightarrow H^{0}(X,G\otimes^{\mathbb{L}}\mathcal{O}_{Z}) is surjective. Consider the spectral sequence

E2p,q=Hp(X,q(G𝕃𝒪Z))Hp+q(X,G𝕃𝒪Z).E_{2}^{p,q}=H^{p}(X,\mathcal{H}^{q}(G\otimes^{\mathbb{L}}\mathcal{O}_{Z}))\Rightarrow H^{p+q}(X,G\otimes^{\mathbb{L}}\mathcal{O}_{Z}).

Note that 0(G𝕃𝒪Z)=G𝒪Z\mathcal{H}^{0}(G\otimes^{\mathbb{L}}\mathcal{O}_{Z})=G\otimes\mathcal{O}_{Z} and i(G𝕃𝒪Z)\mathcal{H}^{i}(G\otimes^{\mathbb{L}}\mathcal{O}_{Z}) is supported in codimension 2 for all ii. Therefore only the terms E20,qE_{2}^{0,q} are nontrivial, so the spectral sequence degenerates and H0(X,G𝕃OZ)=E20,0=H0(X,GOZ)H^{0}(X,G\otimes^{\mathbb{L}}O_{Z})=E_{2}^{0,0}=H^{0}(X,G\otimes O_{Z}). ∎

Remark 4.6.

One can similarly show H1(X,ωX𝕃(LIZ))=0H^{1}(X,\omega_{X}\otimes^{\mathbb{L}}(L\otimes I_{Z}))=0 implies H1(X,ωX(LIZ))=0H^{1}(X,\omega_{X}\otimes(L\otimes I_{Z}))=0 since the non-Gorenstein locus is codimension 22 but we will not make use of this fact.

We now come to the main technical theorem of the paper.

Theorem 4.7.

Let XX be a projective normal surface over an algebraically closed field, ZXZ\subset X a subscheme of length lZl_{Z}, and LL an ample line bundle on XX with c1(L)=:Hc_{1}(L)=:H. Let FF be an object of type OO with lT:=length(0(F))l_{T}:=\operatorname{length}(\mathcal{H}^{0}(F)). Choose nonnegative integers l1,l2l_{1},l_{2} with l1+l2=lZ+lTl_{1}+l_{2}=l_{Z}+l_{T}. Then

Hom(LIZ,F)=0\operatorname{Hom}(L\otimes I_{Z},F)=0

if H2max((CX+2l2+1)2,4(l1+l2+CX)+ϵ)H^{2}\geq\max((C_{X}+2l_{2}+1)^{2},4(l_{1}+l_{2}+C_{X})+\epsilon) and HCmax(2(2l1+CX),CX+2l2+1)H\cdot C\geq\max(2(2l_{1}+C_{X}),C_{X}+2l_{2}+1) for any nonzero effective divisor CC on XX.

Proof.

By repeated applications of Propositions 4.2 and 4.3, we see that to show vanishing of Hom(LIZ,F)\operatorname{Hom}(L\otimes I_{Z},F), it suffices show that Hom(LIZ,F)=0\operatorname{Hom}(L\otimes I_{Z^{\prime}},F^{\prime})=0 for all subschemes ZZ^{\prime} of length l1l_{1} and objects FF^{\prime} of type OO with length(0(F))=l2\operatorname{length}(\mathcal{H}^{0}(F^{\prime}))=l_{2}.

By Proposition 3.11, if f0Hom(LIZ,F)f\neq 0\in\operatorname{Hom}(L\otimes I_{Z^{\prime}},F^{\prime}), then Imf=:G\operatorname{Im}f=:G is a torsion sheaf. By Proposition 3.9, we see that the quotient LIZGL\otimes I_{Z^{\prime}}\twoheadrightarrow G is not destabilizing at the stability condition H:=c1(L),s:=12,t1:=142l1+CXH2H:=c_{1}(L),s:=\frac{1}{2},t_{1}:=\sqrt{\frac{1}{4}-\frac{2l_{1}+C_{X}}{H^{2}}} as long as H2>4(2l1+CX)H^{2}>4(2l_{1}+C_{X}) and Hch1(G)2(2l1+CX)H\cdot\operatorname{ch}_{1}(G)\geq 2(2l_{1}+C_{X}). So we conclude in particular that ds,t1(Imf)>0=νs,t1(LIZ)d_{s,t_{1}}(\operatorname{Im}f)>0=\nu_{s,t_{1}}(L\otimes I_{Z^{\prime}}). Then we note that ds,t(Imf)d_{s,t}(\operatorname{Im}f) is independent of tt since ch0(Imf)=0\operatorname{ch}_{0}(\operatorname{Im}f)=0, so ds,t2(Imf)>0d_{s,t_{2}}(\operatorname{Im}f)>0 for t2:=142l2+CXH2t_{2}:=\sqrt{\frac{1}{4}-\frac{2l_{2}+C_{X}}{H^{2}}}, which implies in particular that ImfF\operatorname{Im}f\hookrightarrow F^{\prime} is strictly destabilizing at (s,t2)(s,t_{2}). But this is impossible if H2(CX+2l2+1)2H^{2}\geq(C_{X}+2l_{2}+1)^{2} and Hch1(G)CX+2l2+1H\cdot\operatorname{ch}_{1}(G)\geq C_{X}+2l_{2}+1 by Proposition 3.8. ∎

In order to turn the previous theorem into a Fujita-type bound, we use the auxiliary functions m,mm,m^{\prime} defined in Definition 1.4.

Remark 4.8.

We can also describe m(CX,l)m(C_{X},l) as

min(max(2CX,CX+2l+1),CX+24l+CX+16+1,2CX+42lCX+16)\min\left(\max(2C_{X},C_{X}+2l+1),C_{X}+2\left\lceil{\frac{4l+C_{X}+1}{6}}\right\rceil+1,2C_{X}+4\left\lceil{\frac{2l-C_{X}+1}{6}}\right\rceil\right)

which shows, for example, that m(CX,l)43lm(C_{X},l)\approx\frac{4}{3}l for l0l\gg 0. In the following corollary, we also prove the lower bound m(CX,l)CX+lm(C_{X},l)\geq C_{X}+l.

Corollary 4.9.

Let XX be a projective normal surface over an algebraically closed field, ZXZ\subset X a subscheme of length lZl_{Z}, LL an ample line bundle on XX, and FF an object of type OO with length(0(F))=lT\operatorname{length}(\mathcal{H}^{0}(F))=l_{T}. Let l:=lZ+lTl:=l_{Z}+l_{T}. Then

Hom(LaIZ,F)=0\operatorname{Hom}(L^{\otimes a}\otimes I_{Z},F)=0

whenever am(CX,l)a\geq m^{\prime}(C_{X},l).

Proof.

Using that (ach1(L))2a2(a\operatorname{ch}_{1}(L))^{2}\geq a^{2}, setting am(CX,l)a\geq m(C_{X},l) clearly accounts for all the bounds in Theorem 4.7 except for H2>4(l1+l2+CX)H^{2}>4(l_{1}+l_{2}+C_{X}). We claim that the only case in which this bound is relevant is l=0,CX=1l=0,C_{X}=1: indeed, for any choice of l1,l20l_{1},l_{2}\in\mathbb{Z}_{\geq 0} with l1+l2=ll_{1}+l_{2}=l, we clearly have

max(2(2l1+CX),CX+2l2+1)max(CX+2l1,CX+2l2)CX+l\max(2(2l_{1}+C_{X}),C_{X}+2l_{2}+1)\geq\max(C_{X}+2l_{1},C_{X}+2l_{2})\geq C_{X}+l

and so

m(CX,l)CX+l>2CX+lm(C_{X},l)\geq C_{X}+l>2\sqrt{C_{X}+l}

for CX+l>2C_{X}+l>2. It therefore suffices to consider the case CX+l2C_{X}+l\leq 2.

When l=0l=0, we have

m(CX,0)=max(2CX,CX+1)2CXm(C_{X},0)=\max(2C_{X},C_{X}+1)\geq 2\sqrt{C_{X}}

with equality exactly when CX=1C_{X}=1, in which case a>2CXa>2\sqrt{C_{X}} implies we need a3a\geq 3. When l=1l=1 and CX1C_{X}\leq 1 we compute

m(CX,1)\displaystyle m(C_{X},1) =min(max(2CX,CX+3),max(2CX+4,CX+1))\displaystyle=\min(\max(2C_{X},C_{X}+3),\max(2C_{X}+4,C_{X}+1))
=min(CX+3,2CX+4)>2CX+1\displaystyle=\min(C_{X}+3,2C_{X}+4)>2\sqrt{C_{X}+1}

and when l=2l=2 and CX=0C_{X}=0 we again have

m(0,2)=4>22.m(0,2)=4>2\sqrt{2}.

Finally, we also give a more explicitly “Reider-type” statement relating the existence of a nontrivial homomorphism to the existence of an effective divisor passing through the points of interest with certain intersection numbers.

Theorem 4.10.

Let XX be a projective normal surface over an algebraically closed field, ZXZ\subset X a subscheme of length lZl_{Z}, LL an ample line bundle on XX, and FF an object of type OO with length(0(F))=lT\operatorname{length}(\mathcal{H}^{0}(F))=l_{T}. Define l:=2lZ+lT+12l^{\prime}:=2\left\lfloor{\frac{l_{Z}+l_{T}+1}{2}}\right\rfloor, i.e., l=lZ+lTl^{\prime}=l_{Z}+l_{T} if this is even and lZ+lT+1l_{Z}+l_{T}+1 otherwise. Assume that H:=c1(L)H:=c_{1}(L) satisfies H2>(CX+l+1)2H^{2}>(C_{X}+l^{\prime}+1)^{2}. Then, if

Hom(LIZ,F)0\operatorname{Hom}(L\otimes I_{Z},F)\neq 0

there is an effective divisor DD such that D2<1D^{2}<1 and

0<DHD2+CX+l.0<D\cdot H\leq D^{2}+C_{X}+l^{\prime}.

Moreover, if one assumes

Hom(LIZ,F)=0\operatorname{Hom}(L\otimes I_{Z^{\prime}},F^{\prime})=0

for every pair of subscheme ZZZ^{\prime}\subset Z and subobject FFF^{\prime}\subset F of type OO (i.e., with 0(F)0(F)\mathcal{H}^{0}(F^{\prime})\subset\mathcal{H}^{0}(F^{\prime})) other than (Z,F)(Z,F), then DD passes through all the points of suppZsupp0(F)\operatorname{supp}Z\cup\operatorname{supp}\mathcal{H}^{0}(F)

Proof.

Choose f0Hom(LIZ,F)f\neq 0\in\operatorname{Hom}(L\otimes I_{Z},F). Then by Proposition 3.11 (whose bound is implied by our bound on H2H^{2}) we have that ImF\operatorname{Im}F is a torsion sheaf, so in particular c1(Imf)=:Dc_{1}(\operatorname{Im}f)=:D is an effective divisor. (Note that Imf\operatorname{Im}f does not contain any points outside of DD by the assumption that there is no nontrivial map from a skyscraper sheaf to FF; in particular, by assumption f0f\neq 0, we have that DD is nontrivial.) Note as well that if DD did not pass through some point psuppZp\in\operatorname{supp}Z, then letting ZZZ^{\prime}\subsetneq Z be the subscheme supported on suppZp\operatorname{supp}Z\setminus p, we would have a map

LIZLIZImfL\otimes I_{Z}\hookrightarrow L\otimes I_{Z^{\prime}}\twoheadrightarrow\operatorname{Im}f

given by LIZ=LIZImFL\otimes I_{Z}=L\otimes I_{Z^{\prime}}\twoheadrightarrow\operatorname{Im}F on XpX\setminus p and 0 on XsuppDX\setminus\operatorname{supp}D, or in other words, we would get that ff is the restriction of some nonzero element of Hom(LIZ,F)\operatorname{Hom}(L\otimes I_{Z^{\prime}},F). Similarly, if DD did not pass through some point psupp0(F)p\in\operatorname{supp}\mathcal{H}^{0}(F), letting

F:=ker(F0(F)0(F)|p)F^{\prime}:=\ker(F\twoheadrightarrow\mathcal{H}^{0}(F)\twoheadrightarrow\mathcal{H}^{0}(F)|_{p})

we see that the composition

ImfF0(F)|p\operatorname{Im}f\hookrightarrow F\twoheadrightarrow\mathcal{H}^{0}(F)|_{p}

must be trivial, and thus ImfF\operatorname{Im}f\hookrightarrow F factors through FF^{\prime}, showing that ff gives a nonzero element of Hom(LIZ,F).\operatorname{Hom}(L\otimes I_{Z},F^{\prime}).

Now, as before, we note that the Bridgeland degree d1/2,t(Imf)d_{1/2,t}(\operatorname{Im}f) is independent of tt, as ch0(Imf)=0\operatorname{ch}_{0}(\operatorname{Im}f)=0. So in particular, either d1/2,t(Imf)0d_{1/2,t}(\operatorname{Im}f)\leq 0, in which case LIZImfL\otimes I_{Z}\twoheadrightarrow\operatorname{Im}f is a destabilizing quotient at t=142lZ+CXH2t=\sqrt{\frac{1}{4}-\frac{2l_{Z}+C_{X}}{H^{2}}}, or d1/2,t(Imf)>0d_{1/2,t}(\operatorname{Im}f)>0, in which case ImfF\operatorname{Im}f\hookrightarrow F is a strictly destabilizing subsheaf at t=142lT+CXH2t=\sqrt{\frac{1}{4}-\frac{2l_{T}+C_{X}}{H^{2}}}. In the first case, we see as in the proof of Proposition 3.9 that we must have

DHD2+CX+2lZD\cdot H\leq D^{2}+C_{X}+2l_{Z}

which, combined with the fact that

0<r1/2,t(Imf)r1/2,t(LIZ)=tH220<r_{1/2,t}(\operatorname{Im}f)\leq r_{1/2,t}(L\otimes I_{Z})=t\frac{H^{2}}{2}

(the rank being nonzero because ch0(Imf)=0\operatorname{ch}_{0}(\operatorname{Im}f)=0 and HD>0H\cdot D>0), means that we are in the situation of Lemma 3.6, assuming H2>(CX+2lZ+1)2H^{2}>(C_{X}+2l_{Z}+1)^{2}, and thus D2<1D^{2}<1. Similarly, in the second case, as in the proof of Proposition 3.7 (where here E=1(cokerf)E=\mathcal{H}^{-1}(\operatorname{coker}f) satisfies ch0(E)=1,ch1(E)=D\operatorname{ch}_{0}(E)=1,\operatorname{ch}_{1}(E)=D) we get

DHD2+CX+2lTD\cdot H\leq D^{2}+C_{X}+2l_{T}

and thus again are in the situation of Lemma 3.6, assuming H2>(CX+2lT+1)2H^{2}>(C_{X}+2l_{T}+1)^{2}, and can deduce D2<1D^{2}<1. So we conclude that we must have

DHD2+CX+2max(lZ,lT)D\cdot H\leq D^{2}+C_{X}+2\max(l_{Z},l_{T})

with D2<1D^{2}<1.

Finally, it remains to note that we can optimize by using Propositions 4.2 and 4.3 to redistribute points between ZZ and 0(F)\mathcal{H}^{0}(F) while preserving suppZsupp0(F)\operatorname{supp}Z\cup\operatorname{supp}\mathcal{H}^{0}(F), lZ+lTl_{Z}+l_{T}, and DD (this last point being because the different homomorphisms constructed all agree away from suppZsupp0(F)\operatorname{supp}Z\cup\operatorname{supp}\mathcal{H}^{0}(F), which is of codimension 2 and so has no effect on the divisor corresponding to Imf\operatorname{Im}f). In particular, we see that the lower bound for H2H^{2} is maximized and the upper bound for DHD\cdot H is minimized when lZ,lTl_{Z^{\prime}},l_{T^{\prime}} are as close to equal as possible, i.e., either both lZ+lT2\frac{l_{Z}+l_{T}}{2} or {lZ+lT+12,lZ+lT12}\{\frac{l_{Z}+l_{T}+1}{2},\frac{l_{Z}+l_{T}-1}{2}\}. ∎

Our desired applications immediately follow by applying the above to F=Ω¯X00[1]F=\operatorname{\underline{\Omega}_{X}^{0}}0[1] and 𝒪X[1]\mathcal{O}_{X}[1].

Proof of Theorem 1.12.

First of all, by Proposition 4.5 it suffices to show H1(X,ωXDB𝕃LIZ)=0.H^{1}(X,\omega_{X}^{DB}\otimes^{\mathbb{L}}L\otimes I_{Z})=0. By Proposition 4.1, vanishing of H1(X,ωXDB𝕃LIZ)H^{1}(X,\omega_{X}^{DB}\otimes^{\mathbb{L}}L\otimes I_{Z}) is equivalent to vanishing of Hom(LIZ,Ω¯X00[1])\operatorname{Hom}(L\otimes I_{Z},\operatorname{\underline{\Omega}_{X}^{0}}0[1]). The conclusion then follows by Theorem 4.7 where F=Ω¯X00[1]F=\operatorname{\underline{\Omega}_{X}^{0}}0[1]. ∎

By letting F=𝒪X[1]F=\mathcal{O}_{X}[1] in Theorem 4.7 we can similarly deduce Theorem 1.2. In particular, the classical Fujita’s conjecture for surfaces follows:

Proof of Corollary 1.7.

This follows by applying Corollary 1.5 with lZ=1l_{Z}=1 and 2, respectively. In the first case, we see that for l1=0,l2=1l_{1}=0,l_{2}=1 we have max(2(2l1+CX),CX+2l2+1)=3\max(2(2l_{1}+C_{X}),C_{X}+2l_{2}+1)=3, and in the second case, for l1=1=l2l_{1}=1=l_{2}, we have max(2(2l1+CX),CX+2l2+1)=4\max(2(2l_{1}+C_{X}),C_{X}+2l_{2}+1)=4. ∎

The Reider-type theorem also recovers a version of the classical Reider theorem for very ampleness (differing from [17] only in that LL is assumed to be ample, not nef):

Corollary 4.11.

Let XX be a smooth projective normal surface with CX=0C_{X}=0 (e.g. a smooth complex surface) and LL an ample line bundle such that H:=c1(L)H:=c_{1}(L) satisfies H2>9H^{2}>9. Let ZZ be a subscheme of length 2. Then if

H0(X,ωXL)H0(X,ωXLOZ)H^{0}(X,\omega_{X}\otimes L)\to H^{0}(X,\omega_{X}\otimes L\otimes O_{Z})

is not surjective there exists an effective divisor DD passing through ZZ (i.e., passing through pp if suppZ={p}\operatorname{supp}Z=\{p\}, and if suppZ={p,q}\operatorname{supp}Z=\{p,q\}, then DD passes through at least one of p,qp,q, and both if neither is a base point of ωXL\omega_{X}\otimes L) such that

either DH=1 and D2=0,1 or DH=2 and D2=0.\textrm{either }D\cdot H=1\textrm{ and }D^{2}=0,-1\textrm{ or }D\cdot H=2\textrm{ and }D^{2}=0.
Proof.

If this is not a surjection, then we must have H1(X,ωXLIZ)0H^{1}(X,\omega_{X}\otimes L\otimes I_{Z})\neq 0, which is equivalent to Hom(LIZ,𝒪X[1])0\operatorname{Hom}(L\otimes I_{Z},\mathscr{O}_{X}[1])\neq 0 by Proposition 4.1. Thus, we can apply Theorem 4.10 with lZ=2l_{Z}=2 and lT=CX=0l_{T}=C_{X}=0 (thus l=2l^{\prime}=2) to conclude that there is an effective divisor DD such that D2<1D^{2}<1 and 0<DHD2+20<D\cdot H\leq D^{2}+2. (For the statement about support, note that H0(X,ωXL)=0H^{0}(X,\omega_{X}\otimes L)=0 by Theorem 4.7, so DD must at least pass through some point of ZZ.) Since XX was assumed to be smooth, we have D2D^{2}\in\mathbb{Z} (and the same, of course, is true for DHD\cdot H), so the statement follows. ∎

References

  • [1] Daniele Arcara and Aaron Bertram “Reider’s theorem and Thaddeus pairs revisited” In Grassmannians, moduli spaces and vector bundles 14, Clay Math. Proc. Amer. Math. Soc., 2011, pp. 51–68
  • [2] Philippe Du Bois “Complexe de de Rham filtré d’une variété singulière” In Bulletin de la Société mathématique de France 109, 1981, pp. 41–81
  • [3] Lawrence Ein and Robert Lazarsfeld “Global generation of pluricanonical and adjoint linear series on smooth projective threefolds” In Journal of the American Mathematical Society 6.4 JSTOR, 1993, pp. 875–903
  • [4] Lawrence Ein, Robert Lazarsfeld and Vladimir Maşek “Global Generation Of Linear Series On Terminal Threefolds” In International Journal of Mathematics 6.01 World Scientific, 1995, pp. 1–18
  • [5] Takao Fujita “Contribution to birational geometry of algebraic varieties: open problems” In the 23rd International Symposium, Division of Mathematics, the Taniguchi Foundation; August 22-27, 1988, Katata, 1988
  • [6] William Fulton “Intersection theory” Springer Science & Business Media, 2013
  • [7] Yi Gu, Lei Zhang and Yongming Zhang “Counterexamples to Fujita’s conjecture on surfaces in positive characteristic” In Advances in Mathematics 400 Elsevier, 2022, pp. 108271
  • [8] Francisco Guillen, Vincente Navarro Aznar, Pedro Pascual-Gainza and Fernando Puerta “Hyperrésolutions cubiques et descente cohomologique” Springer, 2006
  • [9] Yujiro Kawamata “On Fujita’s freeness conjecture for 3-folds and 4-folds” In arXiv preprint alg-geom/9510004, 1995
  • [10] Naoki Koseki “On the Bogomolov–Gieseker inequality in positive characteristic” In International Mathematics Research Notices 2023.24 Oxford University Press, 2023, pp. 20784–20811
  • [11] Sándor Kovács and Karl Schwede “Du Bois singularities deform” In Minimal Models and Extremal Rays (Kyoto, 2011), 2016, pp. 49–65 Math. Soc. Japan
  • [12] Adrian Langer “A note on k-jet ampleness on surfaces” In arXiv preprint math/9802098, 1998
  • [13] Adrian Langer “Bridgeland stability conditions on normal surfaces” In Annali di Matematica Pura ed Applicata (1923-) Springer, 2024, pp. 1–12
  • [14] Adrian Langer “Intersection theory and Chern classes on normal varieties” In arXiv preprint arXiv:2210.08766, 2022
  • [15] Howard Nuer and Alan Sorani “The Bogomolov-Gieseker-Koseki inequality on surfaces with canonical singularities in arbitrary characteristic” In arXiv preprint arXiv:2308.03307, 2023
  • [16] Chris AM Peters and Joseph HM Steenbrink “Mixed hodge structures” Springer Science & Business Media, 2008
  • [17] Igor Reider “Vector bundles of rank 2 and linear systems on algebraic surfaces” In Annals of Mathematics 127.2 JSTOR, 1988, pp. 309–316
  • [18] Morihiko Saito “Mixed Hodge complexes on algebraic varieties” In Mathematische Annalen 316 Springer, 2000, pp. 283–331
  • [19] Fumio Sakai “Reider-Serrano’s method on normal surfaces” In Algebraic Geometry: Proceedings of the International Conference held in L’Aquila, Italy, May 30–June 4, 1988, 2006, pp. 301–319 Springer
  • [20] Karl Schwede “F-injective singularities are Du Bois” In American journal of mathematics 131.2 Johns Hopkins University Press, 2009, pp. 445–473
  • [21] Wanchun Shen, Sridhar Venkatesh and Anh Duc Vo “On kk-Du Bois and kk-rational singularities” In arXiv preprint arXiv:2306.03977, 2023
  • [22] The Stacks Project Authors Stacks Project”, https://stacks.math.columbia.edu, 2018
  • [23] Fei Ye and Zhixian Zhu “On Fujita’s freeness conjecture in dimension 5” In Advances in Mathematics 371 Elsevier, 2020, pp. 107210

Department of Mathematics, MIT, 77 Massachusetts Avenue, Cambridge, MA 02139, USA

E-mail address: [email protected]

Department of Mathematics, Harvard University, 1 Oxford Street, Cambridge, MA 02138, USA

E-mail address: [email protected]