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

Gaeta resolutions and strange duality
over rational surfaces

Thomas Goller The College of New Jersey, NJ, USA [email protected]  and  Yinbang Lin Tongji University, Shanghai, China [email protected]
Abstract.

Over the projective plane and at most two-step blowups of Hirzebruch surfaces, where there are strong full exceptional sequences of line bundles, we obtain foundational results about Gaeta resolutions of coherent sheaves by these line bundles. Under appropriate conditions, we show the locus of semistable sheaves not admitting Gaeta resolutions has codimension at least 2. We then study Le Potier’s strange duality conjecture. Over these surfaces, for two orthogonal numerical classes where one has rank one and the other has sufficiently positive first Chern class, we show that the strange morphism is injective. The main step in the proof is to use Gaeta resolutions to show that certain relevant Quot schemes are finite and reduced, allowing them to be enumerated using the authors’ previous paper.

Key words and phrases:
Exceptional sequence; Gaeta resolution; Moduli of sheaves; Strange duality; Quot scheme
2020 Mathematics Subject Classification:
Primary: 14D20, 14F06; Secondary: 14F08, 14J26

1. Introduction

In the moduli theory of sheaves over complex algebraic surfaces, there is a famous conjecture by Le Potier [LP05], called the strange duality conjecture, which relates the global sections of two determinant line bundles on certain pairs of moduli spaces of sheaves. Known results over rational surfaces are mostly in the cases where one of the moduli spaces parametrizes pure dimension 1 sheaves, see e.g. [Dan02, Abe10, Yua21]. On other surfaces, the conjecture requires different formulations, see e.g. [BMOY17]. In an attempt to provide a unified treatment of the conjecture over rational surfaces, the first author, Bertram, and Johnson [BGJ16] proposed to use Grothendieck’s Quot schemes [Gro61], following Marian and Oprea’s ideas [MO07a, MO07b] over curves. A key tool in the study of Quot schemes over 2\mathbb{P}^{2} in [BGJ16] is Gaeta resolutions of coherent sheaves in terms of the strong full exceptional sequence of line bundles (𝒪(2),𝒪(1),𝒪)(\mathcal{O}(-2),\mathcal{O}(-1),\mathcal{O}).

This leads us to the study of exceptional sequences and Gaeta resolutions over rational surfaces. A strong full exceptional sequence, if it exists, completely captures the derived category in an explicit way, by theorems of Baer [Bae88] and Bondal [Bon89]. While coherent sheaves can always be resolved by locally free sheaves by the Hilbert Syzygy Theorem, we can bring the resolution under better control if the locally free sheaves are taken from an exceptional sequence. Resolutions built from such exceptional sequences, called Gaeta resolutions, have been applied toward a variety of problems in the study of sheaves on rational surfaces [Dre86, LP98, CH18, CH20, CH21]. Though the existence of strong full exceptional sequences of line bundles in general is an open question, the answer is affirmative over a rational surface that can be obtained from a Hirzebruch surface 𝔽e\mathbb{F}_{e} by blowing up at most two sets of points [HP11], which we call a two-step blowup of 𝔽e\mathbb{F}_{e}.

Over 2\mathbb{P}^{2} or a two-step blowup of 𝔽e\mathbb{F}_{e}, we choose a particular strong full exceptional sequence, determine when a sheaf admits Gaeta resolutions, and study general properties of such sheaves. We then apply Gaeta resolutions to the study of strange duality, proving the injectivity of the strange morphism in some cases. One of the key points in the proof is to show that relevant Quot schemes are finite and reduced, which we accomplish using Gaeta resolutions. A parallel statement over 2\mathbb{P}^{2} was proved in [BGJ16]. Another crucial point is to enumerate the length of the finite Quot scheme, which was settled in our previous paper [GL22] via the study of the moduli space of limit stable pairs [Lin18].

We now set up the study. Let SS be a smooth projective algebraic surface over an algebraically closed field kk of characteristic 0, with a strong full exceptional sequence of line bundles

(1,2,,n).(\mathcal{E}_{1},\mathcal{E}_{2},\dots,\mathcal{E}_{n}).

Given a coherent sheaf FF on SS, we would like to find a resolution of FF of the form111To avoid clumsy notation, we drop the direct sum symbol from the exponents. If we want to denote tensor products, we will use \otimes.

01a1dadd+1ad+1nanF0,0\to\mathcal{E}_{1}^{a_{1}}\oplus\cdots\oplus\mathcal{E}_{d}^{a_{d}}\to\mathcal{E}_{d+1}^{a_{d+1}}\oplus\cdots\oplus\mathcal{E}_{n}^{a_{n}}\to F\to 0, (1.1)

which is called a Gaeta resolution. If a Gaeta resolution exists, the exponents a1,,ana_{1},\dots,a_{n} are uniquely determined by the numerical class of FF.

One of our main technical results is a criterion to determine when a Gaeta resolution exists. We state here the criterion for a two-step blowup of 𝔽e\mathbb{F}_{e}, where we know there are strong full exceptional sequences of line bundles, see § 3.3. Let S0S_{0} denote the set of blown-up points in the first step, with corresponding exceptional divisors EiE_{i} for iS0i\in S_{0}. The second set of blown-up points, which lie on these exceptional divisors, is denoted S1S_{1}, and the corresponding exceptional divisors are EjE_{j} for jS1j\in S_{1}. Using the exceptional sequence § 3.3(c), we get the following result.

Proposition 1.1 (Proposition 4.4(b)).

On a two-step blowup of 𝔽e\mathbb{F}_{e}, a torsion-free sheaf FF has a Gaeta resolution if and only if

  1. (i)

    Hp(F)H^{p}(F) vanishes for p0p\neq 0;

  2. (ii)

    Hp(F(D))H^{p}(F(D)) vanishes for p1p\neq 1 and DD the divisors A+iS0Ei-A+\sum_{i\in S_{0}}E_{i}, B+iS0Ei-B+\sum_{i\in S_{0}}E_{i}, and AB+iS0Ei+jS0Ej-A-B+\sum_{i\in S_{0}}E_{i}+\sum_{j\in S_{0}}E_{j}; and

  3. (iii)

    H1(F(Ei))=0H^{1}(F(E_{i}))=0 and H1(F(Ej))=0H^{1}(F(E_{j}))=0 for all iS0i\in S_{0} and jS1j\in S_{1}.

Section 4 contains a similar statement for 2\mathbb{P}^{2}, which is known.

We will apply Gaeta resolutions towards moduli problems, using the connections to prioritary sheaves. We impose a mild technical condition on XX a two-step blowup of 𝔽e\mathbb{F}_{e}, which we call admissibility (Definition 2.7). We use AA to denote both the class of the fibers of the ruling 𝔽e1\mathbb{F}_{e}\to\mathbb{P}^{1} and the pullback of this class to XX. We denote a numerical class ff in the Grothendieck group K(X)\operatorname{K}(X) by

f=(r,L,χ),f=(r,L,\chi),

where rr is the rank, LL is the first Chern class (equivalent to the determinant line bundle), and χ\chi is the Euler characteristic. We have

Proposition 1.2.

Let XX be an admissible blowup of 𝔽e\mathbb{F}_{e} and HH be a polarization such that H(KX+A)<0H\cdot(K_{X}+A)<0. For a numerical class fK(X)f\in\operatorname{K}(X) with fixed rank r>0r>0 and fixed χ0\chi\geqslant 0, suppose the first Chern class LL is sufficiently positive. Then a general HH-semistable sheaf of class ff admits Gaeta resolutions.

We mostly use Gieseker stability and use either “HH-semistable” or “semistable”. If we want to use slope stability, we will specify. For the precise meaning of being sufficiently positive in Proposition 1.2, see the conditions in Proposition 4.5(a) as well as Proposition 5.1(b.ii). There is another statement, Proposition 6.2, in which the rank and first Chern class are fixed, which asserts that if the discriminant is sufficiently large then general semistable sheaves admit Gaeta resolution up to a twist by a line bundle. In each case, we can immediately deduce that the moduli space M(f)M(f) is unirational, which is known [Bal87].

By imposing stronger conditions on the numerical class ff and the polarization HH, we prove a refinement of Proposition 1.2:

Theorem 1.3.

Let XX be an admissible blowup. Assume the class ff is of rank r2r\geqslant 2, admits Gaeta resolutions in which the exponents aia_{i} are strictly positive and satisfy (6.2, 6.3), that the polarization HH is general and satifies (6.4, 6.5), and that the discriminant of ff is sufficiently large in the sense of (6.9). Then the closed subset ZM(f)Z\subset M(f) of S-equivalence classes of semistable sheaves whose Jordan-Hölder gradings do not admit Gaeta resolutions has codimension 2\geqslant 2 in M(f)M(f).

If X=𝔽eX=\mathbb{F}_{e}, the conditions (6.2, 6.4, 6.5) are vacuous. Using Proposition 1.2 and Theorem 1.3, we deduce

Corollary 1.4.

Assuming the conditions from Proposition 1.2 and appropriate conditions from Proposition 5.1(b), a general sheaf in M(f)M(f) is torsion-free, is locally-free if r2r\geqslant 2, satisfies the cohomological vanishing conditions in Proposition 4.4 (b.i-iii), and is globally generated if χr+2\chi\geqslant r+2. Assuming the stronger conditions from Theorem 1.3, the same properties hold away from a locus of codimension 2\geqslant 2 in M(f)M(f) (with locally free replaced by torsion-free).

For the proof of Theorem 1.3, we will need the following statement which is of general interest.

Theorem 1.5.

Suppose SS is a rational surface other than 2\mathbb{P}^{2} and S1S\to\mathbb{P}^{1} is a morphism where a general fiber DD is isomorphic to 1\mathbb{P}^{1}. Let HH be a general ample divisor such that H(KS+2D)<0H\cdot(K_{S}+2D)<0. Assume there is a slope stable vector bundle with rank r2r\geqslant 2, first Chern class c1c_{1}, and second Chern class c21c_{2}-1 over SS. Then

PicM(r,c1,c2)PicS.\operatorname{Pic}M(r,c_{1},c_{2})\cong\mathbb{Z}\oplus\operatorname{Pic}S.

We will apply Gaeta resolutions towards the study of the strange duality conjecture. Let SS be 2\mathbb{P}^{2} or XX an admissible blowup of 𝔽e\mathbb{F}_{e}. In the Grothendieck group K(S)\operatorname{K}(S), let

σ=(r,L,r) for r2andρ=(1,0,1) for 1\sigma=(r,L,r\ell)\mbox{ for }r\geqslant 2\qquad\mbox{and}\qquad\rho=(1,0,1-\ell)\mbox{ for }\ell\geqslant 1

be two numerical classes. Notice that they are orthogonal: χ(σρ)=0\chi(\sigma\cdot\rho)=0. Let HH denote the hyperplane class on 2\mathbb{P}^{2} or a polarization satisfying (6.4, 6.5) on XX. Let M(σ)M(\sigma) and M(ρ)M(\rho) denote the moduli spaces of HH-semistable sheaves, where M(ρ)M(\rho) is isomorphic to the Hilbert scheme S[]S^{[\ell]} of points. Let 𝒵S[]×S\mathscr{Z}\subset S^{[\ell]}\times S be the universal subscheme and I𝒵I_{\mathscr{Z}} its ideal sheaf. For a coherent sheaf WW of class σ\sigma, consider the determinant line bundle

Θσ:=det(p!(I𝒵LqW))\Theta_{\sigma}:=\det\left(p_{!}\left(I_{\mathscr{Z}}\stackrel{{\scriptstyle L}}{{\otimes}}{q}^{*}W\right)\right)^{*}

on S[]S^{[\ell]}, where p{p} and q{q} are the projections from S[]×SS^{[\ell]}\times S to the first and second factors, respectively. There is also a similar line bundle Θρ\Theta_{\rho} on M(σ)M(\sigma). On M(σ)×M(ρ)M(\sigma)\times M(\rho), the line bundle Θσ,ρΘρΘσ\Theta_{\sigma,\rho}\cong\Theta_{\rho}\boxtimes\Theta_{\sigma} has a canonical section, which induces the strange morphism

SDσ,ρ:H0(S[],Θσ)H0(M(σ),Θρ).\operatorname{SD}_{\sigma,\rho}\colon H^{0}(S^{[\ell]},\Theta_{\sigma})^{*}\to H^{0}(M(\sigma),\Theta_{\rho}).

The strange duality conjecture says that SDσ,ρ\operatorname{SD}_{\sigma,\rho} is an isomorphism. For a more general setup, see § 7.1.

In this context, we prove the following result in support of the strange duality conjecture. In the statement, for VV a vector bundle on SS, we write

V[]=p(𝒪𝒵qV).V^{[\ell]}={p}_{*}(\mathcal{O}_{\mathscr{Z}}\otimes{q}^{*}V). (1.2)
Theorem 1.6.

Let SS be 2\mathbb{P}^{2} or XX an admissible blowup of 𝔽e\mathbb{F}_{e}, and σ\sigma, ρ\rho, and HH as above. If LL is sufficiently positive, then:

  1. (a)

    The rank of the strange morphism is bounded below by S[]c2(V[])\int_{S^{[\ell]}}c_{2\ell}({V}^{[\ell]}), for VV a vector bundle with numerical class σ+ρ\sigma+\rho;

  2. (b)

    The strange morphism SDσ,ρ\mathrm{SD}_{\sigma,\rho} is injective.

We sketch the proof. Let VV be a vector bundle of class σ+ρ\sigma+\rho that admits a general Gaeta resolution and consider quotients of VV^{*} of class ρ\rho. Then the Quot scheme has expected dimension 0. If it is finite and reduced, a simple argument shows that its length provides a lower bound for SDσ,ρ\operatorname{SD}_{\sigma,\rho}. According to [GL22], in this case its length is S[]c2(V[])\int_{S^{[\ell]}}c_{2\ell}(V^{[\ell]}). On the other hand, Theorem 7.4 relates this top Chern class to χ(S[],Θσ)\chi(S^{[\ell]},\Theta_{\sigma}), and the determinant line bundle Θσ\Theta_{\sigma} has no higher cohomology when LL is sufficiently positive, which finishes the proof. Thus, the crucial point is to establish that the Quot scheme is finite and reduced, which we prove by considering the relative Quot scheme over the space of Gaeta resolutions and calculating the dimension of the relative Quot scheme. The following theorem summarizes the key results related to the Quot scheme.

Theorem 1.7.

Let SS, σ\sigma, ρ\rho, and HH be as in Theorem 1.6, and VV be a vector bundle of class σ+ρ\sigma+\rho that admits a general Gaeta resolution. If LL is sufficiently positive, then:

  1. (a)

    The Quot scheme Quot(V,ρ){\rm Quot}(V^{*},\rho) parametrizing quotient sheaves of VV^{*} with numerical class ρ\rho is finite and reduced;

  2. (b)

    For every point [VF][V^{*}\twoheadrightarrow F] of Quot(V,ρ){\rm Quot}(V^{*},\rho), FF is an ideal sheaf IZI_{Z} for general ZS[]Z\in S^{[\ell]} and the kernel is semistable;

  3. (c)

    The length of Quot(V,ρ){\rm Quot}(V^{*},\rho) is S[]c2(V[])\int_{S^{[\ell]}}c_{2\ell}({V}^{[\ell]}).

In [BGJ16], parts (a) and (b) were proved over 2\mathbb{P}^{2} and calculations were made that informed Johnson’s expectation that the counting formula (c) should be true for del Pezzo surfaces [Joh18]. The formula (c) was proved in [GL22] by the authors of the current paper, for a general smooth regular projective surface, assuming that the Quot scheme is finite and reduced.

The positivity conditions on LL in these theorems, which are stronger than for Proposition 1.2, are summarized in the appendix. Theorem 1.7 requires (A.2, A.3), and Theorem 1.6 requires (A.4) as well.

We organize the paper as follows. In § 2, we review basic facts about divisors and line bundles on Hirzebruch surfaces and their blowups. In § 3, we review exceptional sequences in the derived category. In § 4, we obtain criteria for the existence of Gaeta resolutions, including Proposition 1.1, and classify the numerical classes of sheaves admitting Gaeta resolutions. In § 5, we prove some general properties of such sheaves and relate them to prioritary sheaves. In § 6, we discuss connections to semistable sheaves and prove Theorem 1.3. In § 7, we set up the strange morphism and prove Theorem 1.6. In § 8, we prove Theorem 1.7. Finally, the appendix contains a summary of the positivity conditions on LL required in the proofs of Theorems 1.6 and 1.7.

Acknowledgment. YL would like to thank Alina Marian and Dragos Oprea for helpful correspondences. TG would like to thank Lothar Göttsche for providing updates on his work with Anton Mellit. YL is supported by grants from the Fundamental Research Funds for the Central Universities and Applied Basic Research Programs of Science and Technology Commission Foundation of Shanghai Municipality.

2. Hirzebruch surfaces and blowups

We review some basic results on divisors and cohomology of line bundles on Hirzebruch surfaces and their blow-ups. Along the way, we introduce two technical assumptions (2.1) and (2.2). The first is not a restriction, while the second is a mild condition.

2.1. Divisors on blowups of Hirzebruch surfaces

Let 𝔽e\mathbb{F}_{e} denote the Hirzebruch surface (𝒪1𝒪1(e))\mathbb{P}(\mathcal{O}_{\mathbb{P}^{1}}\oplus\mathcal{O}_{\mathbb{P}^{1}}(e)) with e0e\geqslant 0, which is a rational surface that is ruled over 1\mathbb{P}^{1}. Note that 𝔽01×1\mathbb{F}_{0}\cong\mathbb{P}^{1}\times\mathbb{P}^{1}. Letting AA denote the class of a fiber and BB the 0-section with self-intersection e-e, AA and BB generate the effective cone of 𝔽e\mathbb{F}_{e}, and A2=0A^{2}=0, AB=1A\cdot B=1, B2=eB^{2}=-e. The canonical divisor is K𝔽e=(e+2)A2BK_{\mathbb{F}_{e}}=-(e+2)A-2B. The divisor classes eA+BeA+B and AA generate the nef cone. To simplify notation, let

C=eA+B.C=eA+B.

The linear system |A||A| induces the morphism to 1\mathbb{P}^{1} giving the ruling, while |C||C| induces a morphism 𝔽ee+1\mathbb{F}_{e}\to\mathbb{P}^{e+1}; if e=0e=0, this is the other ruling of 1×1\mathbb{P}^{1}\times\mathbb{P}^{1}, while if e>0e>0, this contracts BB to a point and maps the fibers of the ruling to distinct lines through that point in e+1\mathbb{P}^{e+1}. In particular, suppose x,yx,y are distinct points on 𝔽e\mathbb{F}_{e}, where we allow yy to be infinitely near to xx. Then |A||A| separates x,yx,y unless x,yx,y are distinct points on the same fiber or yy corresponds to the tangent direction along the fiber at xx. Similarly, |C||C| separates x,yx,y unless x,yx,y are distinct points on BB or xBx\in B and yy is the tangent direction along BB at xx.

Remark 2.1.

The blowup of 2\mathbb{P}^{2} at any point is isomorphic to 𝔽1\mathbb{F}_{1}. The blowup of 1×1\mathbb{P}^{1}\times\mathbb{P}^{1} at any point is isomorphic to the blowup of 𝔽1\mathbb{F}_{1} at a point not on BB. For e>0e>0, the blowup of 𝔽e\mathbb{F}_{e} at a point on BB is isomorphic to the blowup of 𝔽e+1\mathbb{F}_{e+1} at a point not on BB. Thus, when considering the surfaces that arise from blowing up 2\mathbb{P}^{2} or Hirzebruch surfaces, it suffices to consider blowups of 𝔽e\mathbb{F}_{e} for e>0e>0 where the blown-up points are not on BB [GH78, p.519]. So, quite often, in the case e>0e>0, we impose the condition that

the blowup avoids BB (2.1)

in the sense that none of the blown-up points p1,,psp_{1},\dots,p_{s} is on BB.

Let XX be obtained from a sequence of blowups

X=XtXt1X1X0=𝔽e,X=X_{t}\to X_{t-1}\to\cdots\to X_{1}\to X_{0}=\mathbb{F}_{e},

where bi:XiXi1b_{i}\colon X_{i}\to X_{i-1} is the iith blowup at a point piXi1p_{i}\in X_{i-1}. Assume that the indices {1,,t}\{1,\dots,t\} can be partitioned into two sets

S0={1,,s}andS1={s+1,,t}S_{0}=\{1,\dots,s\}\quad\mbox{and}\quad S_{1}=\{s+1,\dots,t\}

such that the bib_{i} within each of the sets {b1,,bs}\{b_{1},\dots,b_{s}\} and {bs+1,,bt}\{b_{s+1},\dots,b_{t}\} commute. In other words, XX can be obtained from 𝔽e\mathbb{F}_{e} by up to two blowups, each possibly at multiple points. By Remark 2.1, it suffices to consider the case e>0e>0 and that the blowup avoids BB.

We define a partial ordering on the set {p1,,pt}\{p_{1},\dots,p_{t}\} by pjpip_{j}\succ p_{i} if pjp_{j} is on the exceptional divisor of bib_{i}, and say that the height of pip_{i} is 0 if pip_{i} is minimal with respect to \succ, while otherwise pip_{i} has height 1. We can consider points of height 0 as lying on 𝔽e\mathbb{F}_{e}, while if pjpip_{j}\succ p_{i} then pjp_{j} is infinitely near to pip_{i} and can be viewed as a tangent direction at pip_{i} on 𝔽e\mathbb{F}_{e}. For simplicity, we choose the partition S0S_{0} and S1S_{1} such that

iS0\displaystyle i\in S_{0}  iff pi has height 0, while\displaystyle\quad\mbox{ iff }p_{i}\mbox{ has height 0, while }
jS1\displaystyle j\in S_{1}  iff pj has height 1.\displaystyle\quad\mbox{ iff }p_{j}\mbox{ has height }1.

Thus, we think of XX as being obtained from 𝔽e\mathbb{F}_{e} by first blowing up a collection of points {p1,,ps}\{p_{1},\dots,p_{s}\} on 𝔽e\mathbb{F}_{e} and then blowing up a collection of points {ps+1,,pt}\{p_{s+1},\dots,p_{t}\} on the exceptional divisors of the first blowup.

Let EiE_{i} denote the total transform in XX of the exceptional divisor of bib_{i}. We abuse notation by writing A,BA,B for the pullbacks of the divisors A,BA,B on 𝔽e\mathbb{F}_{e}. The Picard group of XX is generated by A,B,E1,,EtA,B,E_{1},\dots,E_{t}, with the following intersections:

A2=0,AB=1,B2=e,AEi=0,BEi=0,EiEj=δi,j.A^{2}=0,\quad A\cdot B=1,\quad B^{2}=-e,\quad A\cdot E_{i}=0,\quad B\cdot E_{i}=0,\quad E_{i}\cdot E_{j}=-\delta_{i,j}.

Here, δ\delta is the Kronecker delta function. The canonical divisor is

KX=(e+2)A2B+i=1tEi.K_{X}=-(e+2)A-2B+\sum_{i=1}^{t}E_{i}.

Let E~i\tilde{E}_{i} denote the strict transform of EiE_{i}. If pjp_{j} has height 1, then E~j=Ej\tilde{E}_{j}=E_{j}, while if pip_{i} has height 0, then E~i=Eij:pjpiEj\tilde{E}_{i}=E_{i}-\sum_{j\colon p_{j}\succ p_{i}}E_{j}. Note that E~i2=1#{j:pjpi}\tilde{E}_{i}^{2}=-1-\#\{j\colon p_{j}\succ p_{i}\}.

The following classification of base loci of certain linear systems on XX will be useful. When describing the linear systems below, we write a divisor in parentheses, as in (D)(D), to indicate that it is a fixed part of the linear system.

Lemma 2.2.

Suppose DD on XX is the pullback of an effective divisor on 𝔽e\mathbb{F}_{e}. Then if pjpip_{j}\succ p_{i},

|DEj|=|DEi|+(EiEj).|D-E_{j}|=|D-E_{i}|+(E_{i}-E_{j}).
Proof.

Tensoring the short exact sequence

0𝒪(Ei+Ej)𝒪𝒪(EiEj)00\to\mathcal{O}(-E_{i}+E_{j})\to\mathcal{O}\to\mathcal{O}_{(E_{i}-E_{j})}\to 0

by 𝒪(DEj)\mathcal{O}(D-E_{j}) and taking cohomology, we get an exact sequence

0H0(𝒪(DEi))H0(𝒪(DEj))H0(𝒪(DEj)|(EiEj)).0\to H^{0}(\mathcal{O}(D-E_{i}))\to H^{0}(\mathcal{O}(D-E_{j}))\to H^{0}(\mathcal{O}(D-E_{j})|_{(E_{i}-E_{j})}).

As E~i(DEj)=E~iEj=1\tilde{E}_{i}\cdot(D-E_{j})=-\tilde{E}_{i}\cdot E_{j}=-1, we see that H0(𝒪(DEj)|(EiEj))=0H^{0}(\mathcal{O}(D-E_{j})|_{(E_{i}-E_{j})})=0 as (EiEj)(E_{i}-E_{j}) is a connected (possibly reducible) curve and every section of this line bundle must be 0 on the component E~i\tilde{E}_{i}. ∎

Remark 2.3.

If DD is a divisor on 𝔽e\mathbb{F}_{e} and pip_{i} is a point of height 0, then the curves in the linear series |DEi||D-E_{i}| on XX are in bijection with curves in |D||D| on 𝔽e\mathbb{F}_{e} that contain pip_{i}. Similarly, if pjpip_{j}\succ p_{i}, then curves in |DEiEj||D-E_{i}-E_{j}| on XX are in bijection with curves in |D||D| on 𝔽e\mathbb{F}_{e} that contain pip_{i} and have tangent direction pjp_{j} at pip_{i}. The curves in the linear system on XX are obtained as pullbacks of the corresponding curves on 𝔽e\mathbb{F}_{e}, with one copy of the appropriate exceptional divisors removed.

The linear systems on XX in the following examples will play an important role. We assume that e>0e>0 and that the blowup avoids BB.

Example 2.4.

If jS1j\in S_{1}, let iS0i\in S_{0} denote the index such that pjpip_{j}\succ p_{i}. Then

|AEj|=(AEi)+(EiEj).|A-E_{j}|=(A-E_{i})+(E_{i}-E_{j}).

Moreover, if pjp_{j} is the tangent direction along the fiber AA at pip_{i}, then

|AEj|=(AEiEj)+(Ei)|A-E_{j}|=(A-E_{i}-E_{j})+(E_{i})

The curve (AEi)(A-E_{i}) can be obtained by considering bib_{i} to be the first blown-up point, taking the strict transform of the unique fiber AA containing pip_{i} under bib_{i}, and then taking the pullback of that strict transform under the remaining blowups, which may be reducible if other pip_{i} lie on that fiber or are infinitely near to points on that fiber. In particular, if pjp_{j} is the tangent direction along the fiber AA at pjp_{j}, then pjp_{j} lies on (AEi)(A-E_{i}), and taking the strict transform with respect to bjb_{j} yields the curve (AEiEj)(A-E_{i}-E_{j}).

Example 2.5.

For jS1j\in S_{1}, let iS0i\in S_{0} denote the index such that pjpip_{j}\succ p_{i}. Then

|CEj|=|CEi|+(EiEj),|C-E_{j}|=|C-E_{i}|+(E_{i}-E_{j}),

where |CEi||C-E_{i}| is basepoint-free. This follows from the fact that |C||C| separates points on 𝔽e\mathbb{F}_{e} (including infinitely near points) as long as the points are not contained on BB, and by assumption the blown-up points are not on BB.

Example 2.6.

For iS0i\in S_{0} and jS1j\in S_{1}, the base locus of |C+AEiEj||C+A-E_{i}-E_{j}| can be described as follows:

  1. (a)

    If pjpip_{j}\succ p_{i}, then |C+AEiEj||C+A-E_{i}-E_{j}| is basepoint-free unless pjp_{j} is the tangent direction along the fiber AA containing pip_{i}, in which case

    |C+AEiEj|=|C|+(AEiEj).|C+A-E_{i}-E_{j}|=|C|+(A-E_{i}-E_{j}).
  2. (b)

    If pjpip_{j}\not\succ p_{i}, let iS0i^{\prime}\in S_{0} denote the index such that pjpip_{j}\succ p_{i^{\prime}}. Then

    |C+AEiEj|=|C+AEiEi|+(EiEj),|C+A-E_{i}-E_{j}|=|C+A-E_{i}-E_{i^{\prime}}|+(E_{i^{\prime}}-E_{j}),

    and |C+AEiEi||C+A-E_{i}-E_{i^{\prime}}| is basepoint-free unless pip_{i} and pip_{i^{\prime}} lie on the same fiber AA, in which case

    |C+AEiEj|=|C|+(AEiEi)+(EiEj).|C+A-E_{i}-E_{j}|=|C|+(A-E_{i}-E_{i^{\prime}})+(E_{i^{\prime}}-E_{j}).

We explain the two parts. For (b), the linear system contains the union of (AEi)(A-E_{i}) and a curve in |CEi||C-E_{i^{\prime}}|. As the latter is basepoint-free, the only possible basepoints are on (AEi)(A-E_{i}). By the same argument with the roles of ii and ii^{\prime} reversed, we see that the linear system is basepoint-free unless pip_{i} and pip_{i^{\prime}} lie on the same fiber.

For (a), we note that if the linear system has a basepoint pp, which we may assume is a point of 𝔽e\mathbb{F}_{e}, then the linear system of curves in |C+A||C+A| that contain pip_{i} and pp must have pjp_{j} as an (infinitely near) basepoint, which, by a similar argument as the one for (b), implies that pip_{i} and pp lie on the same fiber and pjp_{j} is the tangent direction along that fiber.

Some of the calculations in later sections are simplified if the linear systems |C+AEiEj||C+A-E_{i}-E_{j}| for pjpip_{j}\succ p_{i} are basepoint-free. For this purpose, we will often assume that

the blowup avoids fiber directions (2.2)

in the sense that ps+1,,ptp_{s+1},\dots,p_{t} are distinct from the point where the strict transform of the fiber AA containing pip_{i} meets the exceptional divisor of bib_{i}, for all iS0i\in S_{0}.

We summarize the assumptions on XX in the following definition:

Definition 2.7.

The rational surface XX is an admissible blowup of 𝔽e\mathbb{F}_{e} if it is an at most two-step blowup and the following conditions hold:

  • if e=0e=0, then S0=S1=S_{0}=S_{1}=\emptyset;

  • if e>0e>0, then the blowup avoids BB and avoids fiber directions.

We emphasize that S0S_{0} or S1S_{1} can be empty. In particular, the definition includes 𝔽e\mathbb{F}_{e}. Then, by the above discussion and particularly Example 2.6, we have shown that if XX is admissible, then every divisor in the set

𝒟={A,C}{CEiiS0}{C+AEiEjpjpi}\mathcal{D}=\{A,C\}\cup\{C-E_{i}\mid i\in S_{0}\}\cup\{C+A-E_{i}-E_{j}\mid p_{j}\succ p_{i}\} (2.3)

is basepoint-free. This leads to the following result.

Proposition 2.8.

Suppose XX is an admissible blowup of 𝔽e\mathbb{F}_{e}. Suppose LL is the line bundle associated to a positive integral linear combination of all divisors in the set

𝒟={A,C}{CEiiS0}{C+AEiEjpjpi}.\mathcal{D}=\{A,C\}\cup\{C-E_{i}\mid i\in S_{0}\}\cup\{C+A-E_{i}-E_{j}\mid p_{j}\succ p_{i}\}.

Then LL is very ample.

Proof.

The linear system associated to LL contains unions of divisors in the linear systems associated to the divisors in 𝒟\mathcal{D}, so since these divisors are all basepoint-free, it suffices to show that the divisors in DD collectively separate points and tangents on XX in the following sense:

  1. 1)

    For any distinct points q1,q2Xq_{1},q_{2}\in X, there is a divisor DD linearly equivalent to a divisor in 𝒟\mathcal{D} such that q1Dq_{1}\in D and q2Dq_{2}\notin D;

  2. 2)

    For any qXq\in X, there are divisors D1,D2D_{1},D_{2} that contain qq, are each linearly equivalent to a divisor in 𝒟\mathcal{D}, and whose images in 𝔪q/𝔪q2\mathfrak{m}_{q}/\mathfrak{m}_{q}^{2} are linearly independent.

For 1), |C||C| can be used to separate points on the complement of BiS0EiB\cup\bigcup_{i\in S_{0}}E_{i} since it is very ample there, |A||A| can separate two points on BB, |CEi||C-E_{i}| can separate two points on E~i\tilde{E}_{i}, and |C+AEiEj||C+A-E_{i}-E_{j}| separates any two points on EjE_{j}. Separating points on different exceptional curves or a point on the exceptional locus from a point on the complement is similarly easy.

For 2), as CC is very ample on the complement of BiS0EiB\cup\bigcup_{i\in S_{0}}E_{i}, it suffices to consider the cases qDq\in D, where DD is BB, E~i\tilde{E}_{i} for iS0i\in S_{0}, or EjE_{j} for jS1j\in S_{1}. In each case, it suffices to choose D1D_{1} transversal to DD at qq and D2D_{2} which is a union of DD with another divisor that does not contain qq. These divisors can be chosen to be general in the following linear subsystems:

D1D_{1} D2D_{2}
qBq\in B |A||A| (B)+|eA||C|(B)+|eA|\subset|C|
qEij:pjpiEjq\in E_{i}\setminus\bigcup_{j\colon p_{j}\succ p_{i}}E_{j} |CEi||C-E_{i}| (Ei)+|CEi||C|(E_{i})+|C-E_{i}|\subset|C|
qEjq\in E_{j} |C+AEiEj||C+A-E_{i}-E_{j}| (Ej)+|CEiEj||CEi|(E_{j})+|C-E_{i}-E_{j}|\subset|C-E_{i}|

This completes the proof. ∎

As very ample is equivalent to 1-very ample and mm-very ampleness is additive under tensor products [HTT05], we immediately see that if LL is the line bundle associated to a positive integral linear combination of all divisors in 𝒟\mathcal{D} in which the weight of each divisor is m\geqslant m, then LL is mm-very ample.

2.2. Cohomology of line bundles

First, we summarize how to calculate the cohomology groups of line bundles on the Hirzebruch surface 𝔽e\mathbb{F}_{e}, following [CH18]. By Hirzebruch-Riemann-Roch,

χ(𝒪(aA+bB))=(a+1)(b+1)12eb(b+1).\chi(\mathcal{O}(aA+bB))=(a+1)(b+1)-\frac{1}{2}eb(b+1).

Since the effective cone of 𝔽e\mathbb{F}_{e} is generated by AA and BB,

H0(𝒪(aA+bB))0if and only ifa,b0.H^{0}(\mathcal{O}(aA+bB))\not=0\quad\mbox{if and only if}\quad a,b\geqslant 0.

Then Serre duality implies that

H2(𝒪(aA+bB))0if and only ifa(e+2) and b2.H^{2}(\mathcal{O}(aA+bB))\not=0\quad\mbox{if and only if}\quad a\leqslant-(e+2)\mbox{ and }b\leqslant-2.

It suffices to assume that b1b\geqslant-1, as other cases can then be obtained via Serre duality. In this case, as h2h^{2} vanishes and the Euler characteristic is known, it suffices to calculate h0h^{0}, which can be done as follows:

  1. (a)

    hi(𝒪(aAB))=0h^{i}(\mathcal{O}(aA-B))=0, for all ii and aa.

  2. (b)

    h0(𝒪(aA))=a+1h^{0}(\mathcal{O}(aA))=a+1 if a1a\geqslant-1 and 0 otherwise.

  3. (c)

    Let b1b\geqslant 1. If abe1a\geqslant be-1, then

    h0(𝒪(aA+bB))=χ(𝒪(aA+bB)),h^{0}(\mathcal{O}(aA+bB))=\chi(\mathcal{O}(aA+bB)),

    while if abe2a\leqslant be-2, then the equality

    h0(𝒪(aA+bB)=h0(𝒪(aA+(b1)B))h^{0}(\mathcal{O}(aA+bB)=h^{0}(\mathcal{O}(aA+(b-1)B))

    allows h0h^{0} to be determined by induction on bb.

In particular, we deduce the following:

Lemma 2.9.

Let L=𝒪𝔽e(aA+bB)L=\mathcal{O}_{\mathbb{F}_{e}}(aA+bB). Then

  1. (a)

    H2(L)=0H^{2}(L)=0 if and only if b1b\geqslant-1 or a1ea\geqslant-1-e;

  2. (b)

    H1(L)=0H^{1}(L)=0 if b=1b=-1, if b=0b=0 and a1a\geqslant-1, or if b1b\geqslant 1 and abe1a\geqslant be-1.

In order to use these calculations on 𝔽e\mathbb{F}_{e} to obtain information about the cohomology of line bundles on XX a two-step blowup of 𝔽e\mathbb{F}_{e}, we use the following general result.

Lemma 2.10.

Let π:Y~Y\pi\colon\tilde{Y}\to Y be a blowup of a smooth projective surface at distinct (possibly infinitely near) points. For a line bundle LL on YY, Hi(Y~,πL)Hi(Y,L)H^{i}(\tilde{Y},\pi^{*}L)\cong H^{i}(Y,L).

Proof.

According to [Har77, V. Proposition 3.4], π𝒪Y~𝒪Y\pi_{*}\mathcal{O}_{\tilde{Y}}\cong\mathcal{O}_{Y} and Riπ𝒪Y~=0R^{i}\pi_{*}\mathcal{O}_{\tilde{Y}}=0 for i>0i>0. Then ππLL\pi_{*}\pi^{*}L\cong L and Riπ(πL)=0R^{i}\pi_{*}(\pi^{*}L)=0 for i>0i>0 by the projection formula. The spectral sequence Hi(Y,Rjπ(πL))Hi+j(Y~,πL)H^{i}(Y,R^{j}\pi_{*}(\pi^{*}L))\Rightarrow H^{i+j}(\tilde{Y},\pi^{*}L) gives the result. ∎

Then, letting XX denote a two-step blowup of 𝔽e\mathbb{F}_{e}, we have:

Lemma 2.11.

Let LL be a line bundle on XX such that L|Ei𝒪EiL|_{E_{i}}\cong\mathcal{O}_{E_{i}} for some 1it1\leqslant i\leqslant t. Then

  1. (a)

    Hp(L(Ei))Hp(L)H^{p}(L(E_{i}))\cong H^{p}(L) for all pp;

  2. (b)

    H2(L(Ei))H2(L)H^{2}(L(-E_{i}))\cong H^{2}(L);

  3. (c)

    If the base locus of |L||L| does not contain EiE_{i}, then H1(L(Ei))H1(L)H^{1}(L(-E_{i}))\cong H^{1}(L).

Proof.

For (a), consider the short exact sequence

0LL(Ei)L(Ei)|Ei𝒪Ei(1)0.0\to L\to L(E_{i})\to L(E_{i})|_{E_{i}}\cong\mathcal{O}_{E_{i}}(-1)\to 0.

Since 𝒪Ei(1)𝒪1(1)\mathcal{O}_{E_{i}}(-1)\cong\mathcal{O}_{\mathbb{P}^{1}}(-1) has no cohomology, we get the result. For (b) and (c), consider the short exact sequence

0L(Ei)LL|Ei𝒪Ei0.0\to L(-E_{i})\to L\to L|_{E_{i}}\cong\mathcal{O}_{E_{i}}\to 0.

Since H2(𝒪Ei)=0H^{2}(\mathcal{O}_{E_{i}})=0, we immediately obtain (b). If the base locus of |L||L| does not contain EiE_{i}, then H0(L)H0(𝒪Ei)kH^{0}(L)\to H^{0}(\mathcal{O}_{E_{i}})\cong k is surjective, which gives the result since H1(𝒪Ei)=0H^{1}(\mathcal{O}_{E_{i}})=0. ∎

3. Exceptional sequences

We review basic facts about Hom functors and exceptional sequences in the bounded derived category of a smooth projective variety YY over kk. In particular, we discuss how to replace a general complex with a complex built from a strong full exceptional sequence 𝔈\mathfrak{E}, which we call an 𝔈\mathfrak{E}-complex.

3.1. Hom functors

If AA^{\bullet} and BB^{\bullet} are bounded complexes of coherent sheaves, then Hom(A,B){\rm Hom}^{\bullet}(A^{\bullet},B^{\bullet}) is the complex of vector spaces defined by

Homi(A,B)=qHom(Aq,Bq+i)andd(f)=dBf(1)ifdA.{\rm Hom}^{i}(A^{\bullet},B^{\bullet})=\bigoplus_{q}{\rm Hom}(A^{q},B^{q+i})\quad\text{and}\quad d(f)=d_{B}\circ f-(-1)^{i}f\circ d_{A}.

The degree-0 cohomology of this complex is the vector space of chain maps ABA^{\bullet}\to B^{\bullet} modulo chain homotopy. This complex is especially useful when it computes the derived functor

RHom(A,B)=jHom(A,B[j]),R{\rm Hom}(A^{\bullet},B^{\bullet})=\bigoplus_{j\in\mathbb{Z}}{\rm Hom}(A^{\bullet},B^{\bullet}[j]),

as in the lemma below. The graded summands RjHom(A,B)=Hom(A,B[j])R^{j}{\rm Hom}(A^{\bullet},B^{\bullet})={\rm Hom}(A^{\bullet},B^{\bullet}[j]) are denoted Extj(A,B){\rm Ext}^{j}(A^{\bullet},B^{\bullet}). In the case when AA^{\bullet} and BB^{\bullet} are sheaves AA and BB in degree 0,

RHom(A,B)=j0Extj(A,B)[j]R{\rm Hom}(A,B)=\bigoplus_{j\geqslant 0}{\rm Ext}^{j}(A,B)[-j]

consists of the usual Extj\mathrm{Ext}^{j} groups for sheaves in each degree jj.

Similarly, om(A,B)\mathcal{H}om^{\bullet}(A^{\bullet},B^{\bullet}) is defined by

omi(A,B)=qom(Aq,Bq+i)andd(f)=dBf(1)ifdA.\mathcal{H}om^{i}(A^{\bullet},B^{\bullet})=\bigoplus_{q}\mathcal{H}om(A^{q},B^{q+i})\quad\text{and}\quad d(f)=d_{B}\circ f-(-1)^{i}f\circ d_{A}.

If either A{A^{\bullet}} or BB^{\bullet} is a complex of locally free sheaves, then om(A,B)\mathcal{H}om^{\bullet}(A^{\bullet},B^{\bullet}) represents Rom(A,B)R\mathcal{H}om(A^{\bullet},B^{\bullet}).

We also recall a few facts about derived functors. The cohomology groups of RΓR\Gamma, often denoted i\mathbb{H}^{i}, are called hypercohomology, and hypercohomology of a sheaf is just sheaf cohomology. The derived functor RHomR{\rm Hom} has the property that RiHom(A,B)=Hom(A,B[i])R^{i}{\rm Hom}(A^{\bullet},B^{\bullet})={\rm Hom}(A^{\bullet},B^{\bullet}[i]). Moreover, RΓRom=R(Γom)=RHomR\Gamma R\mathcal{H}om=R(\Gamma\circ\mathcal{H}om)=R{\rm Hom} and there is a spectral sequence

E1p,q=Hp(Aq)p+q(A).E_{1}^{p,q}=H^{p}(A^{q})\implies\mathbb{H}^{p+q}(A^{\bullet}).

See [Huy06] for more details.

Lemma 3.1.

If AA^{\bullet} and BB^{\bullet} are any complexes composed of locally free sheaves such that all higher Exts between AiA^{i} and BjB^{j} vanish for all i,ji,j, then Hom(A,B){\rm Hom}^{\bullet}(A^{\bullet},B^{\bullet}) computes RHom(A,B)R{\rm Hom}(A^{\bullet},B^{\bullet}). In particular, Hom(A,B){\rm Hom}(A^{\bullet},B^{\bullet}) is the space of chain maps ABA^{\bullet}\to B^{\bullet} modulo chain homotopy.

Proof.

We calculate RHom(A,B)R{\rm Hom}(A^{\bullet},B^{\bullet}) as follows. First, we represent Rom(A,B)R\mathcal{H}om(A^{\bullet},B^{\bullet}) by om(A,B)\mathcal{H}om(A^{\bullet},B^{\bullet}). Then, since AiA^{i} and BjB^{j} have no higher Exts between them, om(Ai,Bj)\mathcal{H}om(A^{i},B^{j}) has no higher cohomology, so by the above spectral sequence we can calculate RΓRom(A,B)R\Gamma R\mathcal{H}om(A^{\bullet},B^{\bullet}) simply as Γom(A,B)=Hom(A,B)\Gamma\mathcal{H}om(A^{\bullet},B^{\bullet})={\rm Hom}^{\bullet}(A^{\bullet},B^{\bullet}). Thus, the complex Hom(A,B){\rm Hom}^{\bullet}(A^{\bullet},B^{\bullet}) represents RHom(A,B)R{\rm Hom}(A^{\bullet},B^{\bullet}), which gives the result. ∎

3.2. Exceptional sequences

The material reviewed in this section can be found in [GK04].

Definition 3.2.

An object Db(Y)\mathcal{E}\in\operatorname{D^{b}}(Y) is exceptional if

Hom(,[])={k,=00,otherwise.{\rm Hom}(\mathcal{E},\mathcal{E}[\ell])=\begin{cases}{k},&\ell=0\\ 0,&\text{otherwise}.\end{cases}

An exceptional sequence is a sequence (1,,n)(\mathcal{E}_{1},\dots,\mathcal{E}_{n}) of exceptional objects such that

Hom(i,j[])=0, for i>j and all .{\rm Hom}(\mathcal{E}_{i},\mathcal{E}_{j}[\ell])=0,\mbox{ for }i>j\mbox{ and all }\ell.

It is strong if in addition

Hom(i,j[])=0, for all i,j and 0.{\rm Hom}(\mathcal{E}_{i},\mathcal{E}_{j}[\ell])=0,\mbox{ for all }i,j\mbox{ and }\ell\not=0.

It is full if {i}i=1n\{\mathcal{E}_{i}\}_{i=1}^{n} generates Db(Y)\operatorname{D^{b}}(Y) as a triangulated category.

Let 𝔈=(1,,n)\mathfrak{E}=(\mathcal{E}_{1},\dots,\mathcal{E}_{n}) be a strong full exceptional sequence of locally free sheaves on YY. The full triangulated subcategories i\langle\mathcal{E}_{i}\rangle generated by individual i\mathcal{E}_{i} yield a semi-orthogonal decomposition of Db(Y)D^{b}(Y). Thus, for each object TT in Db(Y)D^{b}(Y), there is a diagram of morphisms

Cn\textstyle{C_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Cn1\textstyle{C_{n-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\cdots}C2\textstyle{C_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}C1\textstyle{C_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}T=Tn\textstyle{T=T_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Tn1\textstyle{T_{n-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}[1]\scriptstyle{[1]}\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}[1]\scriptstyle{[1]}T1\textstyle{T_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}[1]\scriptstyle{[1]}T00\textstyle{T_{0}\cong 0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}[1]\scriptstyle{[1]} (3.1)

in which each triangle CiTiTi1C_{i}\to T_{i}\to T_{i-1} is distinguished and CiC_{i} is in i\langle\mathcal{E}_{i}\rangle. Each Ti1T_{i-1} can be constructed as the left mutation of TiT_{i} through i\mathcal{E}_{i}, namely as the cone

RHom(i,Ti)iTiTi1.R{\rm Hom}(\mathcal{E}_{i},T_{i})\otimes\mathcal{E}_{i}\to T_{i}\to T_{i-1}.

We call Ci=RHom(i,Ti)iC_{i}=R{\rm Hom}(\mathcal{E}_{i},T_{i})\otimes\mathcal{E}_{i} the factors of TT with respect to the exceptional sequence. The diagram is functorial and in particular, the factors of TT are unique up to isomorphism.

Using the diagram, the factors of TT can be assembled to produce a complex isomorphic to TT. We define an 𝔈\mathfrak{E}-complex to be a bounded complex AA^{\bullet} such that each AiA^{i} is a direct sum of sheaves in the exceptional sequence.

Lemma 3.3.

TT is isomorphic to an 𝔈\mathfrak{E}-complex whose sheaves are the same as in the complex i=1nCi\bigoplus_{i=1}^{n}C_{i} (but with different maps).

Proof.

We prove that each TiT_{i} is isomorphic to an 𝔈\mathfrak{E}-complex by induction on ii. Assume Ti1T_{i-1} is isomorphic to an 𝔈\mathfrak{E}-complex AA^{\bullet}. By the previous lemma, the morphism Ti1[1]CiT_{i-1}[-1]\to C_{i} can be represented by a chain map A[1]CiA^{\bullet}[-1]\to C_{i}, whose mapping cone is an 𝔈\mathfrak{E}-complex whose sheaves are the same as ACiA^{\bullet}\oplus C_{i} and which represents TiT_{i}. ∎

Example 3.4.

Suppose n=3n=3 and TT is a sheaf in degree 0 that has a resolution 01a12a23a3T00\to\mathcal{E}_{1}^{a_{1}}\to\mathcal{E}_{2}^{a_{2}}\oplus\mathcal{E}_{3}^{a_{3}}\to T\to 0. Then the diagram (3.1) can be realized as

3a3\textstyle{\mathcal{E}_{3}^{a_{3}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}2a2\textstyle{\mathcal{E}_{2}^{a_{2}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1a1[1]\textstyle{\mathcal{E}_{1}^{a_{1}}[1]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}T[1a12a23a3]\textstyle{T\cong[\mathcal{E}_{1}^{a_{1}}\to\mathcal{E}_{2}^{a_{2}}\oplus\mathcal{E}_{3}^{a_{3}}]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}[1a12a2]\textstyle{[\mathcal{E}_{1}^{a_{1}}\to\mathcal{E}_{2}^{a_{2}}]\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}[1]\scriptstyle{[1]}1a1[1]\textstyle{\mathcal{E}_{1}^{a_{1}}[1]\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}[1]\scriptstyle{[1]}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}[1]\scriptstyle{[1]}

The proof gives an inductive algorithm for assembling the factors of TT into an 𝔈\mathfrak{E}-complex isomorphic to TT. We say that an 𝔈\mathfrak{E}-complex AA^{\bullet} is minimal if, among all 𝔈\mathfrak{E}-complexes isomorphic to AA^{\bullet}, the total number of sheaves in the complex is as small as possible. If each CiC_{i} is represented by the minimal 𝔈\mathfrak{E}-complex described above, then the complex obtained from this algorithm is minimal as well, and we call it the minimal 𝔈\mathfrak{E}-complex of TT. By Lemma 3.1, it is unique up to quasi-isomorphism.

Using 𝔈\mathfrak{E}-complexes to represent objects in the derived category is useful because the sheaves in an exceptional sequence have no higher Exts between them, so Lemma 3.1 implies that any morphism between two objects represented by 𝔈\mathfrak{E}-complexes can be realized as a chain map between the 𝔈\mathfrak{E}-complexes.

There is a direct way to identify the factors of TT by making use of the dual of the exceptional sequence. The (left) dual of a full exceptional sequence (1,,n)(\mathcal{E}_{1},\dots,\mathcal{E}_{n}) is a full exceptional sequence (n,,1)({\vphantom{\mathcal{E}}}^{\vee}\!\mathcal{E}_{n},\dots,{\vphantom{\mathcal{E}}}^{\vee}\!\mathcal{E}_{1}) with the property that

Hom(i,j[])={k,if =0 and i=j;0,otherwise.{\rm Hom}({\vphantom{\mathcal{E}}}^{\vee}\!\mathcal{E}_{i},\mathcal{E}_{j}[\ell])=\begin{cases}k,&\text{if $\ell=0$ and $i=j$;}\\ 0,&\text{otherwise.}\end{cases} (3.2)

The dual sequence always exists, can be constructed from (1,,n)(\mathcal{E}_{1},\dots,\mathcal{E}_{n}) by mutations, and is characterized up to isomorphism by (3.2).

Lemma 3.5.

Suppose (1,,n)(\mathcal{E}_{1},\dots,\mathcal{E}_{n}) is a strong full exceptional sequence and (n,,1)({\vphantom{\mathcal{E}}}^{\vee}\!\mathcal{E}_{n},\dots,{\vphantom{\mathcal{E}}}^{\vee}\!\mathcal{E}_{1}) is its dual. Then the factors CiC_{i} of an object TT satisfy

CiRHom(i,T)i.C_{i}\cong R{\rm Hom}({\vphantom{\mathcal{E}}}^{\vee}\!\mathcal{E}_{i},T)\otimes\mathcal{E}_{i}.
Proof.

Let AA^{\bullet} be the minimal 𝔈\mathfrak{E}-complex of TT, whose sheaves are the same as iCi\bigoplus_{i}C_{i}. Then, as the higher Exts between i{\vphantom{\mathcal{E}}}^{\vee}\!\mathcal{E}_{i} and the sheaves in (1,,n)(\mathcal{E}_{1},\dots,\mathcal{E}_{n}) vanish, Hom(i,A){\rm Hom}^{\bullet}({\vphantom{\mathcal{E}}}^{\vee}\!\mathcal{E}_{i},A^{\bullet}) computes RHom(i,T)R{\rm Hom}({\vphantom{\mathcal{E}}}^{\vee}\!\mathcal{E}_{i},T). Thus, by property (3.2),

RHom(i,T)Hom(i,Ci)RHom(i,Ti),R{\rm Hom}({\vphantom{\mathcal{E}}}^{\vee}\!\mathcal{E}_{i},T)\cong{\rm Hom}^{\bullet}({\vphantom{\mathcal{E}}}^{\vee}\!\mathcal{E}_{i},C_{i})\cong R{\rm Hom}(\mathcal{E}_{i},T_{i}),

and tensoring by i\mathcal{E}_{i} gives the result. ∎

3.3. Main examples of strong full exceptional sequences

In later sections, we will focus on the following choices of strong full exceptional sequences of line bundles on 2\mathbb{P}^{2}, Hirzebruch surfaces, and, more generally, two-step blowups of Hirzebruch surfaces.

  1. (a)

    On 2\mathbb{P}^{2}, the exceptional sequence (𝒪(2),𝒪(1),𝒪)(\mathcal{O}(-2),\mathcal{O}(-1),\mathcal{O}) is strong and full (see [Huy06, Corollary 8.29, Exercise 8.32] for a more general result on n\mathbb{P}^{n}).

  2. (b)

    On 𝔽e\mathbb{F}_{e}, the exceptional sequence (𝒪(CA),𝒪(C),𝒪(A),𝒪)(\mathcal{O}(-C-A),\mathcal{O}(-C),\mathcal{O}(-A),\mathcal{O}) is strong and full ([Orl92], [HP11] Proposition 5.2). Note that the exceptional and strong properties can easily be checked using Lemma 2.9.

  3. (c)

    On XX a two-step blowup of 𝔽e\mathbb{F}_{e}, consider the exceptional sequence

    𝒪(CA),𝒪(CA+Es+1),,𝒪(CA+Et),𝒪(C),𝒪(A),𝒪(E1),,𝒪(Es),𝒪.\mathcal{O}(-C-A),\mathcal{O}(-C-A+E_{s+1}),\dots,\mathcal{O}(-C-A+E_{t}),\\ \mathcal{O}(-C),\mathcal{O}(-A),\mathcal{O}(-E_{1}),\dots,\mathcal{O}(-E_{s}),\mathcal{O}. (3.3)

    This sequence is obtained from the sequence on 𝔽e\mathbb{F}_{e} in (b) by standard augmentations, so it is strong and full ([HP11] Theorem 5.8). The exceptional and strong properties can easily be checked by using Lemma 2.11 to reduce to calculations on 𝔽e\mathbb{F}_{e}. Note that this example specializes to (b) by allowing the set of blown-up points to be empty.

The dual sequences are as follows and can be verified by checking (3.2):

  • On 2\mathbb{P}^{2}, the dual exceptional sequence is

    𝒪,T(1)[1],𝒪(1)[2],\mathcal{O},T(-1)[-1],\mathcal{O}(1)[-2],

    where TT is the tangent sheaf, which can be checked by Bott’s formula [Bot57].

  • On 𝔽e\mathbb{F}_{e}, the dual exceptional sequence is

    𝒪,𝒪(A)[1],𝒪(B)[1],𝒪(A+B)[2],\mathcal{O},\mathcal{O}(A)[-1],\mathcal{O}(B)[-1],\mathcal{O}(A+B)[-2],

    which follows from the calculations in § 2.2.

  • On the two-step blowup of 𝔽e\mathbb{F}_{e}, the dual exceptional sequence is

    𝒪,𝒪Es[1],,𝒪E1[1],𝒪(AiS0Ei)[1],𝒪(BiS0Ei)[1],\displaystyle\mathcal{O},\mathcal{O}_{E_{s}}[-1],\dots,\mathcal{O}_{E_{1}}[-1],\mathcal{O}(A-\sum_{i\in S_{0}}E_{i})[-1],\mathcal{O}(B-\sum_{i\in S_{0}}E_{i})[-1],
    𝒪Et[2],,𝒪Es+1[2],𝒪(A+BiS0EijS1Ej)[2].\displaystyle\qquad\qquad\qquad\mathcal{O}_{E_{t}}[-2],\dots,\mathcal{O}_{E_{s+1}}[-2],\mathcal{O}(A+B-\sum_{i\in S_{0}}E_{i}-\sum_{j\in S_{1}}E_{j})[-2].

    This can be seen by using the fact that 𝒪(C2AB+iS0Ei+jS1Ej)KX\mathcal{O}(-C-2A-B+\sum_{i\in S_{0}}E_{i}+\sum_{j\in S_{1}}E_{j})\cong K_{X}, the short exact sequences 0𝒪(Ei)𝒪𝒪Ei00\to\mathcal{O}(-E_{i})\to\mathcal{O}\to\mathcal{O}_{E_{i}}\to 0, and Lemma 2.11(a) to reduce to the calculations on 𝔽e\mathbb{F}_{e}.

4. Gaeta resolutions

We study two-step resolutions of coherent sheaves by the exceptional sheaves in the previous section. We call such resolutions Gaeta resolutions. We provide a general criterion and a criterion specialized for rational surfaces that detect when a sheaf admits a Gaeta resolution. We classify numerical classes over two-step blowups of 𝔽e\mathbb{F}_{e} of sheaves admitting Gaeta resolutions, including the case of allowing a twist by a line bundle on 𝔽e\mathbb{F}_{e}. These results lay the foundation for our applications of Gaeta resolutions in later sections.

4.1. Definition of Gaeta resolutions

Let 𝔈=(1,,n)\mathfrak{E}=(\mathcal{E}_{1},\dots,\mathcal{E}_{n}) be a strong full exceptional sequence on a smooth projective variety YY over kk. We are particularly interested in minimal 𝔈\mathfrak{E}-complexes of the following form.

Definition 4.1.

For a coherent sheaf FF, a resolution of FF of the form

01a1dadd+1ad+1nanF00\to\mathcal{E}_{1}^{a_{1}}\oplus\cdots\oplus\mathcal{E}_{d}^{a_{d}}\to\mathcal{E}_{d+1}^{a_{d+1}}\oplus\cdots\oplus\mathcal{E}_{n}^{a_{n}}\to F\to 0

is a Gaeta resolution. If the minimal 𝔈\mathfrak{E}-complex of FF is of this form, then we say that FF admits a Gaeta resolution. The non-negative integers a1,,ana_{1},\dots,a_{n} are called the exponents of the Gaeta resolution.

Clearly, the exponents aia_{i} determine the numerical class of FF. Conversely, the class of FF determines the exponents inductively using semi-orthogonality as

ai={χ(i,F)j=i+1najhom(i,j)for i>d;χ(F,i)j=1i1ajhom(j,i)for id.a_{i}=\begin{cases}\chi(\mathcal{E}_{i},F)-\sum_{j=i+1}^{n}a_{j}\hom(\mathcal{E}_{i},\mathcal{E}_{j})&\text{for $i>d$;}\\ -\chi(F,\mathcal{E}_{i})-\sum_{j=1}^{i-1}a_{j}\hom(\mathcal{E}_{j},\mathcal{E}_{i})&\text{for $i\leqslant d$}.\end{cases}

The dual sequence (n,,1)({\vphantom{\mathcal{E}}}^{\vee}\!\mathcal{E}_{n},\dots,{\vphantom{\mathcal{E}}}^{\vee}\!\mathcal{E}_{1}) can be used to obtain a criterion for when a sheaf admits a Gaeta resolution.

Proposition 4.2 (General criterion).

Let FF be a coherent sheaf. Then FF admits a Gaeta resolution if and only if

Hom(i[j],F){\rm Hom}({\vphantom{\mathcal{E}}}^{\vee}\!\mathcal{E}_{i}[j],F)

vanish for all ii and jj except possibly for

Hom(i,F),d+1in;\displaystyle{\rm Hom}({\vphantom{\mathcal{E}}}^{\vee}\!\mathcal{E}_{i},F),\quad d+1\leqslant i\leqslant n;
Hom(i[1],F),1id.\displaystyle{\rm Hom}({\vphantom{\mathcal{E}}}^{\vee}\!\mathcal{E}_{i}[1],F),\quad 1\leqslant i\leqslant d.

Moreover, if FF admits a Gaeta resolution, then the exponents are

ai={hom(i,F),d+1in;hom(i[1],F),1id.a_{i}=\begin{cases}\hom({\vphantom{\mathcal{E}}}^{\vee}\!\mathcal{E}_{i},F),&d+1\leqslant i\leqslant n;\\ \hom({\vphantom{\mathcal{E}}}^{\vee}\!\mathcal{E}_{i}[1],F),&1\leqslant i\leqslant d.\end{cases}
Proof.

By Lemma 3.5 and (3.2), the condition on Hom(i[j],F){\rm Hom}({\vphantom{\mathcal{E}}}^{\vee}\!\mathcal{E}_{i}[j],F) is equivalent to CiC_{i} being a direct sum of copies of i\mathcal{E}_{i} in degree 0 if d+1ind+1\leqslant i\leqslant n and in degree 1-1 if 1id1\leqslant i\leqslant d. By Lemma 3.3, this proves the first statement. The second statement also follows from the calculation of the CiC_{i}. ∎

4.2. Gaeta resolutions on rational surfaces

In the context of the strong full exceptional sequences in § 3.3, we will focus on the following Gaeta resolutions.

Example 4.3.
  1. (a)

    On 2\mathbb{P}^{2}, we consider Gaeta resolutions of the form

    0𝒪(2)α1𝒪(1)α2𝒪α3\displaystyle 0\to\mathcal{O}(-2)^{\alpha_{1}}\to\mathcal{O}(-1)^{\alpha_{2}}\oplus\mathcal{O}^{\alpha_{3}}\to F0.\displaystyle F\to 0. (4.1)
  2. (b)

    On XX a two-step blowup of 𝔽e\mathbb{F}_{e}, we consider Gaeta resolutions of the form

    0𝒪(CA)α1\displaystyle 0\to\mathcal{O}(-C-A)^{\alpha_{1}} jS1𝒪(CA+Ej)γj\displaystyle\oplus\bigoplus_{j\in S_{1}}\mathcal{O}(-C-A+E_{j})^{\gamma_{j}}
    𝒪(C)α2𝒪(A)α3iS0𝒪(Ei)γi𝒪α4F0.\displaystyle\to\mathcal{O}(-C)^{\alpha_{2}}\oplus\mathcal{O}(-A)^{\alpha_{3}}\oplus\bigoplus_{i\in S_{0}}\mathcal{O}(-E_{i})^{\gamma_{i}}\oplus\mathcal{O}^{\alpha_{4}}\to F\to 0. (4.2)

Note that in the case when the set of blown-up points is empty, (b) specializes to the Gaeta resolutions on 𝔽e\mathbb{F}_{e} that were considered in [CH18].

For these examples, by applying the general criterion for having a Gaeta resolution (Proposition 4.2) with the explicit dual sequences in § 3.3, we deduce a more explicit criterion.

Proposition 4.4.
  1. (a)

    On 2\mathbb{P}^{2}, a sheaf FF admits a Gaeta resolution of the form (4.1) if and only if

    Hp(F)=0H^{p}(F)=0 for p0p\neq 0,  Hp(F(1))=0H^{p}(F(-1))=0 for p1p\neq 1,  and  Hom(T(1),F)=0{\rm Hom}(T(-1),F)=0.
  2. (b)

    On XX a two-step blowup of 𝔽e\mathbb{F}_{e}, a torsion-free sheaf FF admits a Gaeta resolution of the form (b) if and only if

    1. (i)

      Hp(F)H^{p}(F) vanishes for p0p\neq 0;

    2. (ii)

      Hp(F(D))H^{p}(F(D)) vanishes for p1p\neq 1 and DD the divisors A+iS0Ei-A+\sum_{i\in S_{0}}E_{i}, B+iS0Ei-B+\sum_{i\in S_{0}}E_{i}, and AB+iS0Ei+jS0Ej-A-B+\sum_{i\in S_{0}}E_{i}+\sum_{j\in S_{0}}E_{j}; and

    3. (iii)

      H1(F(Ei))=0H^{1}(F(E_{i}))=0 and H1(F(Ej))=0H^{1}(F(E_{j}))=0 for all iS0i\in S_{0} and jS1j\in S_{1}.

Proof.

For (a), Proposition 4.2 includes the first two conditions as well as the condition that Extp(T(1),F){\rm Ext}^{p}(T(-1),F) vanishes for p=0,2p=0,2. Applying Hom(,F){\rm Hom}(-,F) to the short exact sequence 0T(1)𝒪(1)3𝒪(2)00\to T(-1)\to\mathcal{O}(1)^{3}\to\mathcal{O}(2)\to 0 and using the vanishing of H2(F(1))H^{2}(F(-1)) shows that the first two conditions already guarantee Ext2(T(1),F)=0{\rm Ext}^{2}(T(-1),F)=0.

For (b), Proposition 4.2 includes (i) and (ii) as well as the condition that Extp(𝒪Ei,F){\rm Ext}^{p}(\mathcal{O}_{E_{i}},F) vanishes for p=0p=0 and p=2p=2 for all iS0S1i\in S_{0}\cup S_{1}. The vanishing Hom(𝒪Ei,F)=0{\rm Hom}(\mathcal{O}_{E_{i}},F)=0 is guaranteed since FF is torsion-free, while applying Hom(,F){\rm Hom}(-,F) to the short exact sequence 0𝒪(Ei)𝒪𝒪Ei00\to\mathcal{O}(-E_{i})\to\mathcal{O}\to\mathcal{O}_{E_{i}}\to 0 and using H1(F)=H2(F)=0H^{1}(F)=H^{2}(F)=0 shows that Ext1(𝒪(Ei),F)Ext2(𝒪Ei,F){\rm Ext}^{1}(\mathcal{O}(-E_{i}),F)\cong{\rm Ext}^{2}(\mathcal{O}_{E_{i}},F), hence H1(F(Ei))=0H^{1}(F(E_{i}))=0 is an equivalent condition. ∎

4.3. Chern characters and Gaeta resolutions

Recall that the exponents in the Gaeta resolutions are determined by the numerical class of FF. The results in this subsection classify numerical classes that arise as cokernels of Gaeta resolutions. First, we review some useful numerical invariants.

On a smooth projective surface SS over kk, if FF is a coherent sheaf of positive rank rr, first Chern class c1c_{1}, and second Chern character ch2{\rm ch}_{2}, set

ν=c1randΔ=12ν2ch2r,\nu=\frac{c_{1}}{r}\qquad\text{and}\qquad\Delta=\frac{1}{2}\nu^{2}-\frac{{\rm ch}_{2}}{r},

which are called the total slope and discriminant, respectively, of FF. A simple calculation shows that the discriminant of a line bundle is 0 and the discriminant of FF is unchanged when FF is tensored by a line bundle. Using these invariants, the second Chern class of FF can be written as

c2=(r2)ν2+rΔ.c_{2}=\binom{r}{2}\nu^{2}+r\Delta.

Set P(ν)=χ(𝒪S)+12ν(νKS)P(\nu)=\chi(\mathcal{O}_{S})+\frac{1}{2}\nu(\nu-K_{S}). Then by Riemann-Roch the Euler characteristic can be written as

χ(F)=r(P(ν)Δ),\chi(F)=r(P(\nu)-\Delta),

and similarly, if F1,F2F_{1},F_{2} are sheaves and rir_{i}, νi\nu_{i}, Δi\Delta_{i} are the rank, total slope, and discriminant of FiF_{i}, then the Euler pairing is

χ(F1,F2)=r1r2(P(ν2ν1)Δ1Δ2).\chi(F_{1},F_{2})=r_{1}r_{2}(P(\nu_{2}-\nu_{1})-\Delta_{1}-\Delta_{2}). (4.3)

On XX the two-step blowup of a Hirzebruch surface, writing c1=αA+βBiS0γiEijS1γjEjc_{1}=\alpha A+\beta B-\sum_{i\in S_{0}}\gamma_{i}E_{i}-\sum_{j\in S_{1}}\gamma_{j}E_{j}, we compute

P(ν)=(αr+1eβ2r)(βr+1)12iS0S1γir(γir+1).P(\nu)=\left(\frac{\alpha}{r}+1-\frac{e\beta}{2r}\right)\left(\frac{\beta}{r}+1\right)-\frac{1}{2}\sum_{i\in S_{0}\cup S_{1}}\frac{\gamma_{i}}{r}\left(\frac{\gamma_{i}}{r}+1\right).

We write the numerical class of a sheaf as a triple (r,c1,χ)(r,c_{1},\chi) in which rr is a non-negative integer, c1c_{1} is an integral divisor class, and χ\chi is an integer. We say that a numerical class f=(r,c1,χ)f=(r,c_{1},\chi) admits Gaeta resolutions if there is a sheaf FF of rank rr, first Chern class c1c_{1}, and Euler characteristic χ\chi, such that FF admits a Gaeta resolution. For a sheaf FF of class ff and a line bundle LL, we denote the class of FLF\otimes L as f(L)f(L), and we write c2(f)c_{2}(f), ν(f)\nu(f) and Δ(f)\Delta(f) for the second Chern class, total slope, and discriminant of FF, which depend only on ff.

Proposition 4.5.

On XX a two-step blowup of 𝔽e\mathbb{F}_{e}, consider the numerical class

f=(r,αA+βBiS0S1γiEi,χ)f=\Big{(}r,\alpha A+\beta B-\sum_{i\in S_{0}\cup S_{1}}\gamma_{i}E_{i},\chi\Big{)}

of positive rank. Then

  1. (a)

    ff admits Gaeta resolutions (b) if and only if γi\gamma_{i}, γj\gamma_{j}, α4:=χ\alpha_{4}:=\chi, and the following three integers are all 0\geqslant 0:

    α1\displaystyle\alpha_{1} :=χ(f(AB+iS0Ei+jS1Ej))\displaystyle:=-\chi\Big{(}f\Big{(}-A-B+\sum_{i\in S_{0}}E_{i}+\sum_{j\in S_{1}}E_{j}\Big{)}\Big{)}
    α2\displaystyle\alpha_{2} :=χ(f(B+iS0Ei))\displaystyle:=-\chi\Big{(}f\Big{(}-B+\sum_{i\in S_{0}}E_{i}\Big{)}\Big{)}
    α3\displaystyle\alpha_{3} :=χ(f(A+iS0Ei))\displaystyle:=-\chi\Big{(}f\Big{(}-A+\sum_{i\in S_{0}}E_{i}\Big{)}\Big{)}
  2. (b)

    Assume γi0\gamma_{i}\geqslant 0 and γj0\gamma_{j}\geqslant 0. If the discriminant Δ(f)\Delta(f) is sufficiently large, then there is a line bundle LL pulled back from 𝔽e\mathbb{F}_{e} such that f(L)f(L) admits Gaeta resolutions.

Proof.

For (a), assuming ff admits Gaeta resolutions, by comparing first Chern classes, the exponent of 𝒪(CA+Ej)\mathcal{O}(-C-A+E_{j}) must be γj\gamma_{j} and the exponent of 𝒪(Ei)\mathcal{O}(-E_{i}) must be γi\gamma_{i}. The remaining exponents can be easily calculated. Conversely, the inequalities show that we can define the exponents in the same way, and a simple calculation shows that the numerical class of the cokernel must be ff.

For (b), consider the numerical class f=(r,αA+βB,χ+iS0γi)f^{\prime}=\Big{(}r,\alpha A+\beta B,\chi+\sum_{i\in S_{0}}\gamma_{i}\Big{)} on 𝔽e\mathbb{F}_{e}. An elementary calculation shows that

Δ(f)=Δ(f)+iS0(γi/r2)+jS1((γj/r)+12).\Delta(f^{\prime})=\Delta(f)+\sum_{i\in S_{0}}\binom{\gamma_{i}/r}{2}+\sum_{j\in S_{1}}\binom{(\gamma_{j}/r)+1}{2}.

Let M=1+max(iS0γi,jS1γj)/rM=1+\max(\sum_{i\in S_{0}}\gamma_{i},\sum_{j\in S_{1}}\gamma_{j})/r. Since Δ(f)Δ(f)M\Delta(f^{\prime})\geqslant\Delta(f)\gg M, the following lemma ensures that we can choose a line bundle LL on 𝔽e\mathbb{F}_{e} such that f(L)f^{\prime}(L) admits Gaeta resolutions

𝒪(CA)α1𝒪(C)α2𝒪(A)α3𝒪α4\mathcal{O}(-C-A)^{\alpha_{1}^{\prime}}\to\mathcal{O}(-C)^{\alpha_{2}^{\prime}}\oplus\mathcal{O}(-A)^{\alpha_{3}^{\prime}}\oplus\mathcal{O}^{\alpha_{4}^{\prime}}

in which α4=χ(f(L))rM=r+max(iS0γi,jS1γj)\alpha_{4}^{\prime}=\chi(f^{\prime}(L))\geqslant rM=r+\max\Big{(}\sum_{i\in S_{0}}\gamma_{i},\sum_{j\in S_{1}}\gamma_{j}\Big{)}. Here, the inequality follows from the following lemma. Hence, α1max(iS0γi,jS1γj)\alpha_{1}^{\prime}\geqslant\max(\sum_{i\in S_{0}}\gamma_{i},\sum_{j\in S_{1}}\gamma_{j}) as well by comparing ranks. Then a simple calculation shows that the cokernels of Gaeta resolutions

𝒪(CA)α1jγjjS1𝒪(CA+Ej)γj𝒪(C)α2𝒪(A)α3iS0𝒪(Ei)γi𝒪α4iγi\mathcal{O}(-C-A)^{\alpha_{1}^{\prime}-\sum_{j}\gamma_{j}}\oplus\bigoplus_{j\in S_{1}}\mathcal{O}(-C-A+E_{j})^{\gamma_{j}}\to\mathcal{O}(-C)^{\alpha_{2}^{\prime}}\oplus\mathcal{O}(-A)^{\alpha_{3}^{\prime}}\oplus\bigoplus_{i\in S_{0}}\mathcal{O}(-E_{i})^{\gamma_{i}}\oplus\mathcal{O}^{\alpha_{4}^{\prime}-\sum_{i}\gamma_{i}}

have numerical class f(L)f(L). ∎

Lemma 4.6.

On 𝔽e\mathbb{F}_{e}, fix a rank r>0r>0 and a first Chern class c1c_{1} and consider the numerical class f=(r,c1,χ)f=(r,c_{1},\chi). Let MM be a positive real number. Then there are constants C1,C2,C3C_{1},C_{2},C_{3} depending only on ee such that for all χ\chi such that

Δ(f)e+22M2+C1M3/2+C2M+C3,\Delta(f)\geqslant\frac{e+2}{2}M^{2}+C_{1}M^{3/2}+C_{2}M+C_{3},

there is a line bundle LL such that χ(f(L))rM\chi(f(L))\geqslant rM and the class f(L)f(L) admits Gaeta resolutions.

Proof.

We use a setup similar to [CH20, Lemma 4.5]. Consider the curve Q:χ(f(Lx,y))=0Q\colon\chi(f(L_{x,y}))=0 in the xyxy-plane, where Lx,yL_{x,y} is the (in general non-integral) line bundle

Lx,y=xB+yAν(f)+12K𝔽e.L_{x,y}=xB+yA-\nu(f)+\frac{1}{2}K_{\mathbb{F}_{e}}.

Set Δ=Δ(f)\Delta=\Delta(f). By Riemann-Roch,

χ(f(Lx,y))r=x(ye2x)Δ,\frac{\chi(f(L_{x,y}))}{r}=x\left(y-\frac{e}{2}x\right)-\Delta,

so QQ is the hyperbola Δ=x(ye2x)\Delta=x\left(y-\frac{e}{2}x\right), or, as a function of xx, y=Q(x)=Δx+e2xy=Q(x)=\frac{\Delta}{x}+\frac{e}{2}x.

Let Λ\Lambda denote the lattice in the plane of points such that Lx,yL_{x,y} is integral, which is a shift of the standard integral lattice. We say that a point (x,y)Λ(x,y)\in\Lambda is minimal if

  • (x,y)(x,y) is on or above the upper branch Q1Q_{1} of QQ, and

  • (x1,y)(x-1,y) and (x,y1)(x,y-1) are both on or below Q1Q_{1}.

The minimal points exactly correspond to the line bundles Lx,yL_{x,y} such that f(Lx,y)f(L_{x,y}) admits Gaeta resolutions, according to Proposition 4.5(a).

We need to find a minimal point (x,y)(x,y) such that χ(f(Lx,y))/rM\chi(f(L_{x,y}))/r\geqslant M. For this, consider the line y=(e+1)xy=(e+1)x, which intersects Q1Q_{1} at the point

(x,y)=(2Δ/(e+2),(e+1)2Δ/(e+2)).(x^{\prime},y^{\prime})=\left(\sqrt{2\Delta/(e+2)},(e+1)\sqrt{2\Delta/(e+2)}\right).

The tangent line to Q1Q_{1} at this point has equation y=x+2(e+2)Δy=-x+\sqrt{2(e+2)\Delta} and Q1Q_{1} lies above the tangent line. See Figure 1.

Refer to caption
Figure 1. The hyperbola Q1Q_{1}
Refer to caption
Figure 2. Example of minimal points near (x,y)(x^{\prime},y^{\prime})

Let (x0,y0)(x_{0},y_{0}) be the unique minimal point such that ϵx:=xx0\epsilon_{x}:=x^{\prime}-x_{0} satisfies 0ϵx<10\leqslant\epsilon_{x}<1. which lies between Q1Q_{1} and the shift Q1+(0,1)Q_{1}+(0,1). Then ϵy:=y0y\epsilon_{y}:=y_{0}-y^{\prime} satisfies 0ϵy20\leqslant\epsilon_{y}\leqslant 2. Let ϵ:=y0Q(x0)\epsilon:=y_{0}-Q(x_{0}) denote the vertical distance from (x0,y0)(x_{0},y_{0}) to Q1Q_{1}, which satisfies 0ϵ<10\leqslant\epsilon<1. Then χ(f(Lx0,y0))/r=ϵx0\chi(f(L_{x_{0},y_{0}}))/r=\epsilon x_{0}. If ϵx0M\epsilon x_{0}\geqslant M, then Lx0,y0L_{x_{0},y_{0}} gives the result, so assume on the contrary that ϵ<M/x0\epsilon<M/{x_{0}}. Then let mm be a positive integer <x0<x_{0} and consider the lattice point (xm,ym)=(x0m,y0+m)(x_{m},y_{m})=(x_{0}-m,y_{0}+m). We wish to find mm as small as possible such that Q1Q_{1} lies above (xm,ym)(x_{m},y_{m}), as then the point (xm,ym+1)(x_{m},y_{m}+1) will be minimal and ym+1Q(xm)y_{m}+1-Q(x_{m}) will be close to 1. See Figure 2 for an example in which this is achieved with m=4m=4.

To find mm such that Q1Q_{1} lies above (xm,ym)(x_{m},y_{m}), the rise in QQ between x0x_{0} and x0mx_{0}-m should exceed m+ϵm+\epsilon, namely

Q(x0m)Q(x0)mϵ=mΔx02(1+mx0+(mx0)2+)e+22mϵ\displaystyle Q(x_{0}-m)-Q(x_{0})-m-\epsilon=\frac{m\Delta}{x_{0}^{2}}\left(1+\frac{m}{x_{0}}+\left(\frac{m}{x_{0}}\right)^{2}+\cdots\right)-\frac{e+2}{2}m-\epsilon

should be positive. As this quantity exceeds the approximation obtained by truncating the geometric series at the first two terms, and as x0xx_{0}\leqslant x^{\prime}, it suffices to take mm such that

mΔ(x)2(1+mx0)e+22mϵ0,\frac{m\Delta}{(x^{\prime})^{2}}\left(1+\frac{m}{x_{0}}\right)-\frac{e+2}{2}m-\epsilon\geqslant 0,

which yields m2ϵx0/(e+2)m\geqslant\sqrt{2\epsilon x_{0}/(e+2)}. Setting m0=2ϵx0/(e+2)m_{0}=\lceil\sqrt{2\epsilon x_{0}/(e+2)}\rceil, we then have

χ(Lxm0,ym0+1)r\displaystyle\frac{\chi(L_{x_{m_{0}},y_{m_{0}+1}})}{r} =(1+ϵ)x0ϵx(e+1)m0ϵym0e+22m02m0,\displaystyle=(1+\epsilon)x_{0}-\epsilon_{x}(e+1)m_{0}-\epsilon_{y}m_{0}-\frac{e+2}{2}m_{0}^{2}-m_{0},

and for this to be M\geqslant M we need

(1+ϵ)x0e+22m02+(ϵx(e+1)+ϵy+1)m0+M.(1+\epsilon)x_{0}\geqslant\frac{e+2}{2}m_{0}^{2}+(\epsilon_{x}(e+1)+\epsilon_{y}+1)m_{0}+M.

Let δ=m02ϵx0/(e+2)\delta=m_{0}-\sqrt{2\epsilon x_{0}/(e+2)}, which satisfies 0δ<10\leqslant\delta<1. Then the inequality simplifies to

x0(δ(e+2)+ϵx(e+1)+ϵy+1)2ϵx0/(e+2)+(ϵx(e+1)+ϵy+1)δ+e+22δ2+M.x_{0}\geqslant(\delta(e+2)+\epsilon_{x}(e+1)+\epsilon_{y}+1)\sqrt{2\epsilon x_{0}/(e+2)}+(\epsilon_{x}(e+1)+\epsilon_{y}+1)\delta+\frac{e+2}{2}\delta^{2}+M.

Replacing x0x_{0} by xϵxx^{\prime}-\epsilon_{x}, ϵx0\epsilon x_{0} by its upper bound MM, ϵx\epsilon_{x}, ϵy\epsilon_{y}, and δ\delta by their upper bounds, and solving for Δ\Delta, we get a sufficient bound for Δ\Delta:

Δe+22M2+C1M3/2+C2M+C3\Delta\geqslant\frac{e+2}{2}M^{2}+C_{1}M^{3/2}+C_{2}M+C_{3}

for constants C1,C2,C3C_{1},C_{2},C_{3} that depend only on ee. ∎

5. Properties of sheaves with general Gaeta resolutions

Given an exceptional sequence (1,2,,n)(\mathcal{E}_{1},\mathcal{E}_{2},\dots,\mathcal{E}_{n}) from § 3.3 with dd chosen as in Example 4.3 and a sequence of non-negative integers a=(a1,,an)\vec{a}=(a_{1},\dots,a_{n}), consider the vector space

Ha,d=Hom(1a1dad,d+1ad+1nan).H_{\vec{a},d}={\rm Hom}(\mathcal{E}_{1}^{a_{1}}\oplus\cdots\oplus\mathcal{E}_{d}^{a_{d}},\mathcal{E}_{d+1}^{a_{d+1}}\oplus\cdots\oplus\mathcal{E}_{n}^{a_{n}}).

We let R^Ha,d{\hat{R}}\subset H_{\vec{a},d} denote the open subset of injective maps

ϕ:1a1dadd+1ad+1nan,\phi\colon\mathcal{E}_{1}^{a_{1}}\oplus\cdots\oplus\mathcal{E}_{d}^{a_{d}}\to\mathcal{E}_{d+1}^{a_{d+1}}\oplus\cdots\oplus\mathcal{E}_{n}^{a_{n}},

which is non-empty if and only if r:=ad+1++ana1ad0r:=a_{d+1}+\cdots+a_{n}-a_{1}-\cdots-a_{d}\geqslant 0. In this case, we set

Fϕ:=cokerϕ,F_{\phi}:=\operatorname{coker}\phi,

let ff denote the numerical class of these cokernels, which have rank rr, and call the following projectivization the space of Gaeta resolutions:

Rf:=R^(Ha,d).R_{f}:=\mathbb{P}{\hat{R}}\subset\mathbb{P}(H_{\vec{a},d}). (5.1)

Then FϕF_{\phi} satisfies various cohomology vanishing conditions (Proposition 4.4). The purpose of this section is to prove additional properties of sheaves admitting general Gaeta resolutions, including the prioritary condition and a weak Brill-Noether result.

5.1. Basic properties

We begin by proving some basic properties of sheaves admitting a general Gaeta resolution.

Proposition 5.1.
  1. (a)

    On 2\mathbb{P}^{2} and 𝔽e\mathbb{F}_{e}, we have the following for general ϕ\phi:

    1. (i)

      If r=0r=0, then FϕF_{\phi} is torsion supported on the determinant of ϕ\phi (true even if ϕ\phi is not general).

    2. (ii)

      If r=1r=1, then FϕF_{\phi} is torsion-free.

    3. (iii)

      If r2r\geqslant 2, then FϕF_{\phi} is locally free.

    4. (iv)

      If anr+2a_{n}\geqslant r+2, then FϕF_{\phi} is globally generated.

  2. (b)

    On XX an admissible blowup of 𝔽e\mathbb{F}_{e}, using the notation in (b), the same cases are true if we assume, for each iS0i\in S_{0} such that {j:pjpi}\{j\colon p_{j}\succ p_{i}\} is nonempty:

    γi\displaystyle\gamma_{i} j:pjpiγj\displaystyle\geqslant\sum_{j\colon p_{j}\succ p_{i}}\gamma_{j} for (ii),\displaystyle\text{for (ii)},
    γi\displaystyle\gamma_{i} 1+j:pjpiγj\displaystyle\geqslant 1+\sum_{j\colon p_{j}\succ p_{i}}\gamma_{j} for (iii),\displaystyle\text{for (iii)},
    γi\displaystyle\gamma_{i} rα4+j:pjpiγj\displaystyle\geqslant r-\alpha_{4}+\sum_{j\colon p_{j}\succ p_{i}}\gamma_{j} for (iv).\displaystyle\text{for (iv)}.

The key to the proof of this proposition is a Bertini-type statement concerning the codimension on which a general map between vector bundles 𝒜\mathcal{A} and \mathcal{B} drops rank. The case when 𝒜\mathcal{B}\otimes\mathcal{A}^{*} is globally generated is well known.

Proposition 5.2.

On a smooth projective variety YY, consider maps ϕ:𝒜\phi\colon\mathcal{A}\to\mathcal{B}, where 𝒜\mathcal{A}, \mathcal{B} are fixed vector bundles of ranks a,ba,b such that aba\leqslant b. Let

ZY×Hom(𝒜,)Z\subset Y\times\mathbb{P}{\rm Hom}(\mathcal{A},\mathcal{B})

denote the locus of pairs (y,ϕ)(y,\phi) such that the rank of ϕ\phi at yy is <a<a.

  1. (a)

    Suppose 𝒜\mathcal{A}^{*}\otimes\mathcal{B} is globally generated. Then the codimension of ZZ in Y×Hom(𝒜,)Y\times\mathbb{P}{\rm Hom}(\mathcal{A},\mathcal{B}) is ba+1b-a+1.

  2. (b)

    Suppose there are non-trivial decompositions 𝒜=𝒜1𝒜2\mathcal{A}=\mathcal{A}_{1}\oplus\mathcal{A}_{2} and =12\mathcal{B}=\mathcal{B}_{1}\oplus\mathcal{B}_{2} such that each 𝒜ij\mathcal{A}_{i}^{*}\otimes\mathcal{B}_{j} is globally generated except that H0(𝒜21)=0H^{0}(\mathcal{A}_{2}^{*}\otimes\mathcal{B}_{1})=0. Set ai=rk(𝒜i)a_{i}=\mathrm{rk}(\mathcal{A}_{i}) and bj=rk(j)b_{j}=\mathrm{rk}(\mathcal{B}_{j}). If b2<a2b_{2}<a_{2}, then Z=Y×Hom(𝒜,)Z=Y\times\mathbb{P}{\rm Hom}(\mathcal{A},\mathcal{B}). If b2a2b_{2}\geqslant a_{2}, then the codimension of ZZ in Y×Hom(𝒜,)Y\times\mathbb{P}{\rm Hom}(\mathcal{A},\mathcal{B}) is min(b2a2,ba)+1\min(b_{2}-a_{2},b-a)+1.

Proof.

Part (a) is a special case of [Ott95, Teorema 2.8] (see also [Hui16, Proposition 2.6]) and can be proved as follows. Consider the map of global sections

π:Hom(𝒜,)𝒪Yom(𝒜,),\pi\colon{\rm Hom}(\mathcal{A},\mathcal{B})\otimes\mathcal{O}_{Y}\to\mathcal{H}om(\mathcal{A},\mathcal{B}),

which is surjective since 𝒜\mathcal{A}^{*}\otimes\mathcal{B} is globally generated. This map induces a map of projective bundles

ev:Y×Hom(𝒜,)om(𝒜,),(y,ϕ:𝒜)(y,ϕ|y:𝒜|y|y).{\rm ev}\colon Y\times\mathbb{P}{\rm Hom}(\mathcal{A},\mathcal{B})\to\mathbb{P}\mathcal{H}om(\mathcal{A},\mathcal{B}),\qquad(y,\phi\colon\mathcal{A}\to\mathcal{B})\mapsto(y,\phi|_{y}\colon\mathcal{A}|_{y}\to\mathcal{B}|_{y}).

Let Σ\Sigma denote the locus in the target of points (y,ϕy:𝒜|y|y)(y,\phi_{y}\colon\mathcal{A}|_{y}\to\mathcal{B}|_{y}) such that ϕy:𝒜|y|y\phi_{y}\colon\mathcal{A}|_{y}\to\mathcal{B}|_{y} drops rank, and let Z=ev1(Σ)Z={\rm ev}^{-1}(\Sigma). As Σ\Sigma has codimension ba+1b-a+1 in each Hom(𝒜|y,|y)\mathbb{P}{\rm Hom}(\mathcal{A}|_{y},\mathcal{B}|_{y}) and π|y\pi|_{y} is surjective, ZZ has codimension ba+1b-a+1 in Y×Hom(𝒜,)Y\times\mathbb{P}{\rm Hom}(\mathcal{A},\mathcal{B}).

We prove (b) by adapting this argument. In this case, π\pi is not surjective, so the codimension of ZZ may drop if the image of ev{\rm ev} is not transversal to Σ\Sigma. At each point yy, fixing bases, ϕ|y\phi|_{y} is a b×ab\times a matrix of the form

[M0NP]\begin{bmatrix}M&0\\ N&P\end{bmatrix}

in which M,N,PM,N,P are general if ϕ\phi is general. Since bab\geqslant a, this matrix drops rank if and only if the columns are linearly dependent. If b2<a2b_{2}<a_{2}, the columns of PP are always linearly dependent, while if b2a2b_{2}\geqslant a_{2}, there are two cases to consider in which the columns are linearly dependent:

  1. (i)

    The columns of PP are linearly dependent. This occurs in codimension 1+b2a21+b_{2}-a_{2} in the space of such block matrices.

  2. (ii)

    The columns of PP are linearly independent. Then the linear dependence involves a column c\vec{c} of [MN]\begin{bmatrix}M\\ N\end{bmatrix}, hence c\vec{c} is in the span of the remaining columns, which occurs in codimension 1+ba1+b-a. (For such ϕ\phi, Σ\Sigma intersects the image of ev{\rm ev} transversally.)

Thus, ZZ has codimension 1+min(b2a2,ba)1+\min(b_{2}-a_{2},b-a) in Y×Hom(𝒜,)Y\times\mathbb{P}{\rm Hom}(\mathcal{A},\mathcal{B}).

Proof of Proposition 5.1.

For (a), as each ij\mathcal{E}_{i}\otimes\mathcal{E}_{j}^{*} for jdj\leqslant d and i>di>d is globally generated, Proposition 5.2 (a) implies that for general ϕ\phi the locus where ϕ\phi drops rank is either empty or has codimension r+1r+1, which proves (i)(i) and (iii)(iii). Then (ii)(ii) follows from the fact that FϕF_{\phi} cannot have zero-dimensional torsion as it has a two-step resolution by locally free sheaves. For (iv), there is a commutative diagram

𝒪an\textstyle{\mathcal{O}^{a_{n}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒪an\textstyle{\mathcal{O}^{a_{n}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1a1\textstyle{\mathcal{E}_{1}^{a_{1}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ϕ\scriptstyle{\phi}2a2n1an1𝒪an\textstyle{\mathcal{E}_{2}^{a_{2}}\oplus\cdots\oplus\mathcal{E}_{n-1}^{a_{n-1}}\oplus\mathcal{O}^{a_{n}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Fϕ\textstyle{F_{\phi}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

with exact rows, hence 𝒪anFϕ\mathcal{O}^{a_{n}}\to F_{\phi} is surjective if and only if the induced map

1a12a2n1an1\mathcal{E}_{1}^{a_{1}}\to\mathcal{E}_{2}^{a_{2}}\oplus\cdots\oplus\mathcal{E}_{n-1}^{a_{n-1}}

is surjective. This is a map between a bundle of rank a1a_{1} and a bundle of rank

a2++an1=a1+rana12,a_{2}+\cdots+a_{n-1}=a_{1}+r-a_{n}\leqslant a_{1}-2,

hence by dualizing the same Bertini-type statement, a general such map is surjective on all fibers.

For (b), set

\displaystyle\mathcal{F} =𝒪(CA)α1jS1𝒪(CA+Ej)γjand\displaystyle=\mathcal{O}(-C-A)^{\alpha_{1}}\oplus\bigoplus_{j\in S_{1}}\mathcal{O}(-C-A+E_{j})^{\gamma_{j}}\quad\mbox{and}
𝒢\displaystyle\mathcal{G} =𝒪(C)α2𝒪(A)α3iS0𝒪(Ei)γi𝒪α4.\displaystyle=\mathcal{O}(-C)^{\alpha_{2}}\oplus\mathcal{O}(-A)^{\alpha_{3}}\oplus\bigoplus_{i\in S_{0}}\mathcal{O}(-E_{i})^{\gamma_{i}}\oplus\mathcal{O}^{\alpha_{4}}.

The complication is that for each jS1j\in S_{1}, the global sections of the line bundle

L𝒪(C+AEj)L\otimes\mathcal{O}(C+A-E_{j})

vanish on EiEjE_{i}-E_{j}, where ii is the index such that pipjp_{i}\prec p_{j} and LL is any line bundle in 𝒢\mathcal{G} except for 𝒪(Ei)\mathcal{O}(-E_{i}). Still, we can adapt the proof of the Bertini-type statement as follows.

Consider

π:Hom(,𝒢)𝒪Xom(,𝒢),\pi\colon{\rm Hom}(\mathcal{F},\mathcal{G})\otimes\mathcal{O}_{X}\to\mathcal{H}om(\mathcal{F},\mathcal{G}),

which is not surjective on each EiEjE_{i}-E_{j}, and the induced map

ev:X×Hom(,𝒢)om(,𝒢),(y,ϕ:𝒢)(y,ϕ|y:|y𝒢|y).{\rm ev}\colon X\times\mathbb{P}{\rm Hom}(\mathcal{F},\mathcal{G})\to\mathbb{P}\mathcal{H}om(\mathcal{F},\mathcal{G}),\qquad(y,\phi\colon\mathcal{F}\to\mathcal{G})\mapsto(y,\phi|_{y}\colon\mathcal{F}|_{y}\to\mathcal{G}|_{y}).

Let Σ\Sigma denote the locus in the target where the linear maps drop rank and Z=ev1(Σ)Z={\rm ev}^{-1}(\Sigma). At points pp in the open complement UU of the exceptional locus iS0Ei\bigsqcup_{i\in S_{0}}E_{i}, π|p\pi|_{p} is surjective, hence the codimension of Z|UZ|_{U} in U×Hom(,𝒢)U\times\mathbb{P}{\rm Hom}(\mathcal{F},\mathcal{G}) is r+1r+1. On each E~i\tilde{E}_{i}, we apply Proposition 5.2 with

Y=E~i,𝒜2=j:pjpi𝒪(CA+Ej)γj|E~i,2=𝒪(Ei)γi|E~iY=\tilde{E}_{i},\quad\mathcal{A}_{2}=\bigoplus_{j\colon p_{j}\succ p_{i}}\mathcal{O}(-C-A+E_{j})^{\gamma_{j}}|_{\tilde{E}_{i}},\quad\mathcal{B}_{2}=\mathcal{O}(-E_{i})^{\gamma_{i}}|_{\tilde{E}_{i}}

and 𝒜1,1\mathcal{A}_{1},\mathcal{B}_{1} the restrictions of the remaining line bundles. Assuming γij:pjpiγj\gamma_{i}\geqslant\sum_{j\colon p_{j}\succ p_{i}}\gamma_{j} and using the fact that the global sections of each line bundle summand of 𝒜\mathcal{A}^{*}\otimes\mathcal{B} lift to XX, we deduce that the codimension of Z|E~iZ|_{\tilde{E}_{i}} in E~i×Hom(,𝒢)\tilde{E}_{i}\times\mathbb{P}{\rm Hom}(\mathcal{F},\mathcal{G}) is

min(γij:pjpiγj,r)+1.\min(\gamma_{i}-\sum_{j\colon p_{j}\succ p_{i}}\gamma_{j},r)+1.

On EjE_{j}, a similar argument with

Y=Ej,𝒜2=jj:pjpi𝒪(CA+Ej)γj|Ej,2=𝒪(Ei)γi|EjY={E}_{j},\quad\mathcal{A}_{2}=\bigoplus_{j^{\prime}\neq j\colon p_{j^{\prime}}\succ p_{i}}\mathcal{O}(-C-A+E_{j^{\prime}})^{\gamma_{j^{\prime}}}|_{E_{j}},\quad\mathcal{B}_{2}=\mathcal{O}(-E_{i})^{\gamma_{i}}|_{E_{j}}

shows that Z|EjZ|_{E_{j}} has codimension

min(γi+γjj:pjpiγj,r)+1\min(\gamma_{i}+\gamma_{j}-\sum_{j\colon p_{j}\succ p_{i}}\gamma_{j},r)+1

in Ej×Hom(,𝒢)E_{j}\times\mathbb{P}{\rm Hom}(\mathcal{F},\mathcal{G}). Altogether, since E~i\tilde{E}_{i} and EjE_{j} have codimension 1 in XX, we deduce that the general fiber of ZZ over Hom(,𝒢)\mathbb{P}{\rm Hom}(\mathcal{F},\mathcal{G}) is empty or has codimension at least

min(r,γij:pjpiγj+1)+1\min(r,\gamma_{i}-\sum_{j\colon p_{j}\succ p_{i}}\gamma_{j}+1)+1

in XX. As this codimension is 2\geqslant 2 in the case r=1r=1 and 3\geqslant 3 in the case r2r\geqslant 2 assuming γi1+j:pjpiγj\gamma_{i}\geqslant 1+\sum_{j\colon p_{j}\succ p_{i}}\gamma_{j} for all iS0i\in S_{0}, we get the result for (b.ii) and (b.iii).

For (b.iv), we dualize and use a similar argument. ∎

5.2. Prioritary sheaves

We relate sheaves which admit general Gaeta resolutions to prioritary sheaves, which will facilitate the study of stability in § 6. We begin by reviewing the prioritary condition.

Definition 5.3.

Let SS be a smooth surface and DD be a divisor on it. A coherent sheaf FF on SS is DD-prioritary if it is torsion-free and Ext2(F,F(D))=0{\rm Ext}^{2}(F,F(-D))=0. If SS is a Hirzebruch surface or its blowup, AA-prioritary sheaves are simply called prioritary sheaves.

We will need the following lemma ([CH21, Lemma 3.1]) comparing prioritary conditions with respect to different divisors.

Lemma 5.4.

Let SS be a smooth surface, D1D_{1} and D2D_{2} be two divisors such that D1D2D_{1}\geqslant D_{2}. Then D1D_{1}-prioritary sheaves are D2D_{2}-prioritary.

Now let XX denote an admissible blowup of 𝔽e\mathbb{F}_{e} and ff be a fixed class of positive rank admitting Gaeta resolutions.

Proposition 5.5.

For a divisor DD, consider the locus

{ϕRfFϕ is torsion-free and not D-prioritary}.\{\phi\in R_{f}\mid\text{$F_{\phi}$ is torsion-free and not $D$-prioritary}\}.

If D=AD=A, then the locus is empty. If D=A+CD=A+C, then the locus is empty if e=0e=0 and otherwise has codimension iS0γi+α4r+1\geqslant\sum_{i\in S_{0}}\gamma_{i}+\alpha_{4}-r+1 in RfR_{f}.

Proof.

We begin by proving the second statement. Let F=FϕF=F_{\phi} and take D=C+AD=C+A. We need to study Ext2(F,F(D))Hom(F(KD),F){\rm Ext}^{2}(F,F(-D))\cong{\rm Hom}(F(-K-D),F)^{\vee}. We twist the Gaeta resolution of FF by 𝒪(KD)\mathcal{O}(-K-D) and then apply Hom(,F){\rm Hom}(-,F), obtaining the exact sequence

0\displaystyle 0\to Hom(F(KD),F)\displaystyle{\rm Hom}(F(-K-D),F)
\displaystyle\to H0(F(C+K+D))α2H0(F(A+K+D))α3\displaystyle H^{0}(F(C+K+D))^{\alpha_{2}}\oplus H^{0}(F(A+K+D))^{\alpha_{3}}
iS0H0(F(Ei+K+D))γiH0(F(K+D))α4\displaystyle\oplus\bigoplus_{i\in S_{0}}H^{0}(F(E_{i}+K+D))^{\gamma_{i}}\oplus H^{0}(F(K+D))^{\alpha_{4}}
\displaystyle\xrightarrow{h} H0(F(C+A+K+D))α1jS1H0(F(C+A+K+DEj))γj.\displaystyle H^{0}(F(C+A+K+D))^{\alpha_{1}}\oplus\bigoplus_{j\in S_{1}}H^{0}(F(C+A+K+D-E_{j}))^{\gamma_{j}}.

A calculation using the Gaeta resolution for FF shows that H0(F(K+D))H^{0}(F(K+D)) vanishes as, after tensoring by K+DK+D, the line bundles in degree 0 have no global sections and the line bundles in degree 1-1 have no H1H^{1}:

H1(𝒪(AC+K+D))H1(𝒪)0H^{1}(\mathcal{O}(-A-C+K+D))\cong H^{1}(\mathcal{O})^{\vee}\cong 0

and, for jS1j\in S_{1},

H1(𝒪(AC+Ej+K+D))H1(𝒪(Ej))0.H^{1}(\mathcal{O}(-A-C+E_{j}+K+D))\cong H^{1}(\mathcal{O}(-E_{j}))^{\vee}\cong 0.

Similarly, for iS0i\in S_{0}, we get H0(F(Ei+K+D))0H^{0}(F(E_{i}+K+D))\cong 0 since H1(𝒪(Ei))0H^{1}(-\mathcal{O}(E_{i}))\cong 0 and H1(𝒪(EiEj))0H^{1}(\mathcal{O}(-E_{i}-E_{j}))\cong 0. Thus, the map hh in the exact sequence reduces to the map obtained by applying Hom(,F(K+D)){\rm Hom}(-,F(K+D)) to the map

𝒪(CA)α1jS1𝒪(CA+Ej)γj𝜓𝒪(C)α2𝒪(A)α3\mathcal{O}(-C-A)^{\alpha_{1}}\oplus\bigoplus_{j\in S_{1}}\mathcal{O}(-C-A+E_{j})^{\gamma_{j}}\xrightarrow{\psi}\mathcal{O}(-C)^{\alpha_{2}}\oplus\mathcal{O}(-A)^{\alpha_{3}}

from the Gaeta resolution. If this map is surjective, then hh must to be injective. The locus in (Ha,d)\mathbb{P}(H_{\vec{a},d}) of ϕ\phi such that ψ\psi is not surjective has codimension α1+jS1γjα2α3+1=iS0γi+α4r+1\geqslant\alpha_{1}+\sum_{j\in S_{1}}\gamma_{j}-\alpha_{2}-\alpha_{3}+1=\sum_{i\in S_{0}}\gamma_{i}+\alpha_{4}-r+1, giving the desired estimate. If e=0e=0, then B=CB=C and S0S_{0} and S1S_{1} are empty since XX is admissible, from which it follows that H0(F(C+K+D))H0(F(A+K+D))0H^{0}(F(C+K+D))\cong H^{0}(F(A+K+D))\cong 0.

The first statement follows from a similar argument with D=AD=A by checking that H0(F(C+K+D))H^{0}(F(C+K+D)), H0(F(A+K+D))H^{0}(F(A+K+D)), H0(F(Ei+K+D))H^{0}(F(E_{i}+K+D)) for iS0i\in S_{0}, and H0(F(K+D))H^{0}(F(K+D)) all vanish. ∎

The case e=0e=0 (where X=𝔽0X=\mathbb{F}_{0}) is in [Ped21, Proposition 2.20].

Proposition 5.6.

Fix exponents such that r1r\geqslant 1, the condition in Proposition 5.1(b.ii) holds, and iS0γi+α4r\sum_{i\in S_{0}}\gamma_{i}+\alpha_{4}\geqslant r. Then the open family of Gaeta resolutions whose cokernels are torsion-free and (C+A)(C+A)-prioritary is a complete family of (C+A)(C+A)-prioritary sheaves.

By Lemma 5.4, the same statement holds with C+AC+A replaced by any divisor D𝒟D\in\mathcal{D}, where 𝒟\mathcal{D} is defined in (2.3). The inequality iS0γi+α4r\sum_{i\in S_{0}}\gamma_{i}+\alpha_{4}\geqslant r is not needed for the case D=AD=A.

Proof.

If FF has a general Gaeta resolution, then Proposition 5.1(b) implies that FF is torsion-free, and Proposition 5.5 implies that FF is (C+A)(C+A)-prioritary. Thus, the argument proving [CH18, Proposition 3.6] applies here: the sequence being strong full exceptional implies the Kodaira-Spencer map is surjective. ∎

5.3. Weak Brill-Noether result

Let SS be 2\mathbb{P}^{2} or XX an admissible blowup of 𝔽e\mathbb{F}_{e}. Consider Gaeta resolutions 1a1dadd+1ad+1nan\mathcal{E}_{1}^{a_{1}}\oplus\cdots\oplus\mathcal{E}_{d}^{a_{d}}\to\mathcal{E}_{d+1}^{a_{d+1}}\oplus\cdots\oplus\mathcal{E}_{n}^{a_{n}} on SS of the types in Example 4.3.

For integers 1\ell\geqslant 1 and r1r\geqslant 1, set an=ra_{n}=\ell r and assume the remaining exponents a1,,an10a_{1},\dots,a_{n-1}\geqslant 0 satisfy

a1++ad=ad+1++an1+(1)r.a_{1}+\cdots+a_{d}=a_{d+1}+\cdots+a_{n-1}+(\ell-1)r. (5.2)

On XX, we further assume

γij:pjpiγj,for all iS0.\gamma_{i}\geqslant\sum_{j\colon p_{j}\succ p_{i}}\gamma_{j},\qquad\text{for all $i\in S_{0}$}. (5.3)

Consider a general map

ϕ:1a1dadd+1ad+1n1an1𝒪Xr.\phi\colon\mathcal{E}_{1}^{a_{1}}\oplus\cdots\oplus\mathcal{E}_{d}^{a_{d}}\to\mathcal{E}_{d+1}^{a_{d+1}}\oplus\cdots\oplus\mathcal{E}_{n-1}^{a_{n-1}}\oplus\mathcal{O}_{X}^{\ell r}.

By Proposition 5.1, ϕ\phi is injective and the cokernel FϕF_{\phi} is torsion-free of rank rr. Furthermore, it has vanishing higher cohomology, and

𝒪SrH0(Fϕ)𝒪S.\mathcal{O}_{S}^{\ell r}\cong H^{0}(F_{\phi})\otimes\mathcal{O}_{S}.

Then, if ZZ is a zero-dimensional subscheme of length \ell, χ(FϕIZ)=0\chi(F_{\phi}\otimes I_{Z})=0 and we have the following weak Brill-Noether result.

Proposition 5.7.

Let SS denote 2\mathbb{P}^{2} or XX an admissible blowup of 𝔽e\mathbb{F}_{e}. Suppose the sequence of exponents (a1,,an)(a_{1},\dots,a_{n}) satisfies an=ra_{n}=\ell r and (5.2) for some r,1r,\ell\geqslant 1, as well as (5.3) in the case S=XS=X. If ϕ\phi is general and ZS[]Z\in S^{[\ell]} is general, then FϕIZF_{\phi}\otimes I_{Z} has vanishing cohomology in all degrees.

Proof.

As the vanishing of the cohomology of FϕIZF_{\phi}\otimes I_{Z} is an open condition on families of ϕ\phi and of ZZ, it suffices to prove the claim for a single choice of ϕ\phi. We construct ϕ\phi as the direct sum of rr maps {ϕm}1mr\{\phi_{m}\}_{1\leqslant m\leqslant r}, each of which is a general map of the form

1a1dadd+1ad+1n1an1𝒪X,\mathcal{E}_{1}^{a_{1}^{\prime}}\oplus\cdots\oplus\mathcal{E}_{d}^{a_{d}^{\prime}}\to\mathcal{E}_{d+1}^{a_{d+1}^{\prime}}\oplus\cdots\oplus\mathcal{E}_{n-1}^{a_{n-1}^{\prime}}\oplus\mathcal{O}_{X}^{\ell},

where the exponents, which depend on mm, are non-negative and satisfy

a1++ad=ad+1++an1+1.a_{1}^{\prime}+\cdots+a_{d}^{\prime}=a_{d+1}^{\prime}+\cdots+a_{n-1}^{\prime}+\ell-1.

On XX, we need the additional condition that the exponent of 𝒪(Ei)\mathcal{O}(-E_{i}) is at least as large as the sum of the exponents of 𝒪(CA+Ej)\mathcal{O}(-C-A+E_{j}) for jS1j\in S_{1} such that pjpip_{j}\succ p_{i}, which can be ensured due to (5.3). Then, by Proposition 5.1, the cokernel of ϕ\phi is of the form Fϕm=1rLmIZmF_{\phi}\cong\bigoplus_{m=1}^{r}L_{m}\otimes I_{Z_{m}^{\prime}}, where, for each mm, LmL_{m} is a line bundle with vanishing higher cohomology, ZmZ_{m}^{\prime} is a 0-dimensional subscheme, vanishing on ZmZ_{m}^{\prime} imposes independent conditions on H0(Lm)H^{0}(L_{m}), and H0(LmIZm)H^{0}(L_{m}\otimes I_{Z_{m}^{\prime}}) is \ell-dimensional. Choose distinct points Z={q1,,q}Z=\{q_{1},\dots,q_{\ell}\} inductively so that each quq_{u} avoids the base loci of the linear systems of curves in |Lm||L_{m}| that vanish on Zm{q1,,qu1}Z_{m}^{\prime}\sqcup\{q_{1},\dots,q_{u-1}\}. Then FϕIZm=1rLmIZmZF_{\phi}\otimes I_{Z}\cong\bigoplus_{m=1}^{r}L_{m}\otimes I_{Z_{m}^{\prime}\sqcup Z} has no cohomology. ∎

5.4. No sections vanishing on curves

The following result will be used to apply [GL22] in the study of finite Quot schemes in § 8.

Proposition 5.8.

Suppose ϕ\phi is a general Gaeta resolution and FϕF_{\phi} is the cokernel.

  1. (a)

    On 2\mathbb{P}^{2} or 𝔽e\mathbb{F}_{e}, FϕF_{\phi} has no sections vanishing on curves.

  2. (b)

    On XX an admissible blowup of 𝔽e\mathbb{F}_{e}, assume

    1. (i)

      γjα4r\gamma_{j}\geqslant\alpha_{4}-r for all jS1j\in S_{1},

    2. (ii)

      γijS1:pjpiγj+max(0,α4r)\gamma_{i}\geqslant\sum_{j\in S_{1}\colon p_{j}\succ p_{i}}\gamma_{j}+\max(0,\alpha_{4}-r) for all iS0i\in S_{0}.

    Then FϕF_{\phi} has no sections vanishing on curves.

Proof.

Let F=FϕF=F_{\phi}. It suffices to prove that H0(F(D))=0H^{0}(F(-D))=0 for all minimal nonzero effective divisors DD. Twisting the Gaeta resolution by D-D and taking cohomology, we get a long exact sequence

0H0(F(D))H1(1(D)a1d(D)ad)ϕ~H1(d+1(D)ad+1n(D)an).0\to H^{0}(F(-D))\to H^{1}(\mathcal{E}_{1}(-D)^{a_{1}}\oplus\cdots\oplus\mathcal{E}_{d}(-D)^{a_{d}})\xrightarrow{\tilde{\phi}}H^{1}(\mathcal{E}_{d+1}(-D)^{a_{d+1}}\oplus\cdots\oplus\mathcal{E}_{n}(-D)^{a_{n}}).

We need to show that the induced map ϕ~\tilde{\phi} is injective when ϕ\phi is general.

(a) On 2\mathbb{P}^{2}, the only minimal effective divisor is the hyperplane class HH, and H1(𝒪(3))H^{1}(\mathcal{O}(-3)) vanishes, so injectivity of ϕ~\tilde{\phi} is trivial. Similarly, on 𝔽e\mathbb{F}_{e}, the minimal effective divisors are AA and BB, and an easy calculation using § 2.2 shows that 𝒪(C2A)\mathcal{O}(-C-2A) and 𝒪(CAB)\mathcal{O}(-C-A-B) both have vanishing H1H^{1}.

(b) Now consider the sequences on the two-step blowup of 𝔽e\mathbb{F}_{e}. We argue by cases.

Case 1: DD is minimal nonzero effective not equal to any E~i\tilde{E}_{i} for iS0i\in S_{0} or EjE_{j} for jS1j\in S_{1}. In this case, as above, we show that the domain of ϕ~\tilde{\phi} vanishes. Fix a curve YY in the linear equivalence class of DD. As DD is minimal effective, YY must be connected. As C+AC+A is ample on 𝔽e\mathbb{F}_{e} and DD is not contained in Ei\bigcup E_{i}, there is a curve YY^{\prime} in the linear series |C+A||C+A| that is connected, does not contain YY, and intersects YY. Then YYY\cup Y^{\prime} is a connected curve in the linear series |C+A+D||C+A+D|, so H1(𝒪(CAD))=0H^{1}(\mathcal{O}(-C-A-D))=0 by Lemma 5.9.

We use a similar argument to show the vanishing of H1(𝒪(CAEjD))H^{1}(\mathcal{O}(-C-A-E_{j}-D)) for jS1j\in S_{1}. Let iS0i\in S_{0} be the index such that pjpip_{j}\succ p_{i}. By Lemma 2.2,

|C+AEj|=|C+AEi|+(EiEj),|C+A-E_{j}|=|C+A-E_{i}|+(E_{i}-E_{j}),

where |C+AEi||C+A-E_{i}| is basepoint-free (Example 2.5) and contains connected curves that intersect E~i\tilde{E}_{i}, so the union of such a curve and (EiEj)(E_{i}-E_{j}) is connected. Thus, for DD minimal effective not equal to any E~i\tilde{E}_{i} or EjE_{j}, H1(𝒪(CA+EjD))=0H^{1}(\mathcal{O}(-C-A+E_{j}-D))=0 (and this vanishing holds for D=EjD=E_{j} as well).

Case 2: D=EjD=E_{j} for jS1j\in S_{1}. Note that if LL is a line bundle with no cohomology whose restriction to EjE_{j} is trivial, then taking cohomology of the short exact sequence

0L(Ej)LL|Ej00\to L(-E_{j})\to L\to L|_{E_{j}}\to 0

yields an isomorphism H1(L(Ej))H0(𝒪Ej)H^{1}(L(-E_{j}))\cong H^{0}(\mathcal{O}_{E_{j}}). Moreover, because H1(𝒪(Ej))=0H^{1}(\mathcal{O}(-E_{j}))=0 and H1(𝒪(CA+EjEj))=0H^{1}(\mathcal{O}(-C-A+E_{j}-E_{j}))=0, we can view ϕ~\tilde{\phi} as the map obtained from

𝒪(CA)α1jS1{j}𝒪(CA+Ej)γj𝒪(C)α2𝒪(A)α3iS0𝒪(Ei)γi\mathcal{O}(-C-A)^{\alpha_{1}}\oplus\bigoplus_{j^{\prime}\in S_{1}\setminus\{j\}}\mathcal{O}(-C-A+E_{j^{\prime}})^{\gamma_{j^{\prime}}}\to\mathcal{O}(-C)^{\alpha_{2}}\oplus\mathcal{O}(-A)^{\alpha_{3}}\oplus\bigoplus_{i\in S_{0}}\mathcal{O}(-E_{i})^{\gamma_{i}}

by restricting to EjE_{j} and then taking the induced map on global sections. As the restriction to EjE_{j} of each of these exceptional sheaves is trivial, ϕ~\tilde{\phi} is of the form

ϕ~:kα1jS1{j}kγjkα2kα3iS0kγi.\tilde{\phi}\colon{k}^{\alpha_{1}}\oplus\bigoplus_{j^{\prime}\in S_{1}\setminus\{j\}}{k}^{\gamma_{j^{\prime}}}\to{k}^{\alpha_{2}}\oplus{k}^{\alpha_{3}}\oplus\bigoplus_{i\in S_{0}}{k}^{\gamma_{i}}.

Thus, a necessary condition for ϕ~\tilde{\phi} to be injective is α1γj+jS1γjα2+α3+iS0γi\alpha_{1}-\gamma_{j}+\sum_{j^{\prime}\in S_{1}}\gamma_{j^{\prime}}\leqslant\alpha_{2}+\alpha_{3}+\sum_{i\in S_{0}}\gamma_{i}, or equivalently, as α1+jS1γj+r=α2+α3+iS0γi+α4\alpha_{1}+\sum_{j^{\prime}\in S_{1}}\gamma_{j^{\prime}}+r=\alpha_{2}+\alpha_{3}+\sum_{i\in S_{0}}\gamma_{i}+\alpha_{4},

γjα4r.\gamma_{j}\geqslant\alpha_{4}-r.

To find additional sufficient conditions for ϕ~\tilde{\phi} to be injective, we observe that certain blocks of ϕ~\tilde{\phi} are injective when ϕ\phi is general. Let ii denote the index such that pjpip_{j}\succ p_{i}. Then:

  • The block jj:pjpikγjkγi\bigoplus_{j^{\prime}\neq j\colon p_{j^{\prime}}\succ p_{i}}{k}^{\gamma_{j^{\prime}}}\to{k}^{\gamma_{i}} is injective for general ϕ\phi if

    jj:pjpiγjγi\sum_{j^{\prime}\neq j\colon p_{j^{\prime}}\succ p_{i}}\gamma_{j^{\prime}}\leqslant\gamma_{i}

    because the linear series |C+AEiEj||C+A-E_{i}-E_{j^{\prime}}| is basepoint-free. In fact, this inequality is necessary because EjE_{j} is in the base locus of |AEj||A-E_{j^{\prime}}|, |eA+BEj||eA+B-E_{j^{\prime}}|, and |C+AEiEj||C+A-E_{i^{\prime}}-E_{j^{\prime}}| for all jjj^{\prime}\neq j such that pjpip_{j^{\prime}}\succ p_{i} and all iS0{i}i^{\prime}\in S_{0}\setminus\{i\}, hence all other blocks involving these kγjk^{\gamma_{j^{\prime}}} are 0.

  • For a similar reason, for each iS0{i}i^{\prime}\in S_{0}\setminus\{i\}, j:pjpikγjkγi\bigoplus_{j^{\prime}\colon p_{j^{\prime}}\succ p_{i^{\prime}}}{k}^{\gamma_{j^{\prime}}}\to{k}^{\gamma_{i^{\prime}}} is injective for general ϕ\phi if

    j:pjpiγjγi\sum_{j^{\prime}\colon p_{j^{\prime}}\succ p_{i^{\prime}}}\gamma_{j^{\prime}}\leqslant\gamma_{i^{\prime}}

    (though this inequality may not be necessary for ϕ~\tilde{\phi} to be injective).

As all blocks corresponding to kα1{k}^{\alpha_{1}} are general if ϕ\phi is general because the linear systems |A||A|, |C||C|, and |C+AEi||C+A-E_{i^{\prime}}| for iS0i^{\prime}\in S_{0} are all basepoint-free, these inequalities suffice to ensure that ϕ~\tilde{\phi} is injective when ϕ\phi is general.

Case 3: D=E~iD=\tilde{E}_{i} for iS0i\in S_{0}. Similar to the previous case, if LL is a line bundle on XX with no cohomology whose restriction to E~i\tilde{E}_{i} is trivial, then

H1(L(E~i))H0(L|E~i),H^{1}(L(-\tilde{E}_{i}))\cong H^{0}(L|_{\tilde{E}_{i}}),

while H1(𝒪(E~i))=0H^{1}(\mathcal{O}(-\tilde{E}_{i}))=0. Thus, ϕ~\tilde{\phi} can be viewed as the map obtained by restricting

𝒪(CA)α1jS1\displaystyle\mathcal{O}(-C-A)^{\alpha_{1}}\oplus\bigoplus_{j\in S_{1}} 𝒪(CA+Ej)γj\displaystyle\mathcal{O}(-C-A+E_{j})^{\gamma_{j}} (*)
𝒪(C)α2𝒪(A)α3iS0{i}𝒪(Ei)γi𝒪(Ei)γi\displaystyle\to\mathcal{O}(-C)^{\alpha_{2}}\oplus\mathcal{O}(-A)^{\alpha_{3}}\oplus\bigoplus_{i^{\prime}\in S_{0}\setminus\{i\}}\mathcal{O}(-E_{i^{\prime}})^{\gamma_{i^{\prime}}}\oplus\mathcal{O}(-E_{i})^{\gamma_{i}}

to E~i\tilde{E}_{i} and then taking the induced map on global sections. The restriction of each of the exceptional bundles in ()(*) to E~i\tilde{E}_{i} is 𝒪E~i\mathcal{O}_{\tilde{E}_{i}}, except for 𝒪(Ei)|E~i𝒪E~i(1)\mathcal{O}(-E_{i})|_{\tilde{E}_{i}}\cong\mathcal{O}_{\tilde{E}_{i}}(1) and 𝒪(CA+Ej)|E~i𝒪E~i(1)\mathcal{O}(-C-A+E_{j})|_{\tilde{E}_{i}}\cong\mathcal{O}_{\tilde{E}_{i}}(1) for pjpip_{j}\succ p_{i}, each of which has a two-dimensional space of global sections. Thus, the restriction of (*) to E~i\tilde{E}_{i} is

𝒪E~iα1j:pjpi𝒪E~iγjj:pjpi𝒪E~i(1)γj𝒪E~iα2𝒪E~iα3iS0{i}𝒪E~iγi𝒪E~i(1)γi\mathcal{O}_{\tilde{E}_{i}}^{\alpha_{1}}\oplus\bigoplus_{j\colon p_{j}\not\succ p_{i}}\mathcal{O}_{\tilde{E}_{i}}^{\gamma_{j}}\oplus\bigoplus_{j\colon p_{j}\succ p_{i}}\mathcal{O}_{\tilde{E}_{i}}(1)^{\gamma_{j}}\to\mathcal{O}_{\tilde{E}_{i}}^{\alpha_{2}}\oplus\mathcal{O}_{\tilde{E}_{i}}^{\alpha_{3}}\oplus\bigoplus_{i^{\prime}\in S_{0}\setminus\{i\}}\mathcal{O}_{\tilde{E}_{i}}^{\gamma_{i^{\prime}}}\oplus\mathcal{O}_{\tilde{E}_{i}}(1)^{\gamma_{i}}

and hence ϕ~\tilde{\phi} is of the form

ϕ~:kα1j:pjpikγjj:pjpiH0(𝒪E~i(1))γjkα2kα3iS0{i}kγiH0(𝒪E~i(1))γi.\tilde{\phi}\colon{k}^{\alpha_{1}}\oplus\bigoplus_{j\colon p_{j}\not\succ p_{i}}{k}^{\gamma_{j}}\oplus\bigoplus_{j\colon p_{j}\succ p_{i}}H^{0}(\mathcal{O}_{\tilde{E}_{i}}(1))^{\gamma_{j}}\to{k}^{\alpha_{2}}\oplus{k}^{\alpha_{3}}\oplus\bigoplus_{i^{\prime}\in S_{0}\setminus\{i\}}{k}^{\gamma_{i^{\prime}}}\oplus H^{0}(\mathcal{O}_{\tilde{E}_{i}}(1))^{\gamma_{i}}.

A necessary inequality for ϕ~\tilde{\phi} to be injective is α1+jS1γj+jS1:pjpiγjα2+α3+iS0γi+γi\alpha_{1}+\sum_{j\in S_{1}}\gamma_{j}+\sum_{j\in S_{1}\colon p_{j}\succ p_{i}}\gamma_{j}\leqslant\alpha_{2}+\alpha_{3}+\sum_{i^{\prime}\in S_{0}}\gamma_{i^{\prime}}+\gamma_{i}, or equivalently,

γiα4r+j:pjpiγj.\gamma_{i}\geqslant\alpha_{4}-r+\sum_{j\colon p_{j}\succ p_{i}}\gamma_{j}.

As in the previous case, we obtain sufficient conditions for ϕ~\tilde{\phi} to be injective when ϕ\phi is general by looking at various blocks.

  • The block j:pjpiH0(𝒪E~i(1))γjH0(𝒪E~i(1))γi\bigoplus_{j\colon p_{j}\succ p_{i}}H^{0}(\mathcal{O}_{\tilde{E}_{i}}(1))^{\gamma_{j}}\to H^{0}(\mathcal{O}_{\tilde{E}_{i}}(1))^{\gamma_{i}} is injective for general ϕ\phi if

    γij:pjpiγj\gamma_{i}\geqslant\sum_{j\colon p_{j}\succ p_{i}}\gamma_{j}

    because the linear system |C+AEiEj||C+A-E_{i}-E_{j}| is basepoint-free. As there are no nonzero maps 𝒪E~i(1)𝒪E~i\mathcal{O}_{\tilde{E}_{i}}(1)\to\mathcal{O}_{\tilde{E}_{i}}, this condition is necessary for ϕ~\tilde{\phi} to be injective as all other blocks involving j:pjpiH0(𝒪E~i(1))γj\bigoplus_{j\colon p_{j}\succ p_{i}}H^{0}(\mathcal{O}_{\tilde{E}_{i}}(1))^{\gamma_{j}} are 0.

  • For iS0{i}i^{\prime}\in S_{0}\setminus\{i\}, the block j:pjpikγjkγi\bigoplus_{j\colon p_{j}\succ p_{i^{\prime}}}{k}^{\gamma_{j}}\to{k}^{\gamma_{i^{\prime}}} is injective for general ϕ\phi if

    γij:pjpiγj\gamma_{i^{\prime}}\geqslant\sum_{j\colon p_{j}\succ p_{i^{\prime}}}\gamma_{j}

    (but this condition may not be necessary).

The blocks involving kα1{k}^{\alpha_{1}} are all general for general ϕ\phi as the linear systems |A||A|, |C||C|, |C+AEi||C+A-E_{i^{\prime}}| for iii^{\prime}\neq i are all basepoint-free and the curves in |C+A||C+A| containing pip_{i} have no fixed tangent direction at pip_{i}, so the above inequalities are sufficient. ∎

Lemma 5.9.

Let DD be a nonzero effective divisor on a rational surface and |D||D| denote its linear series. The following are equivalent:

  1. (a)

    |D||D| contains a connected curve;

  2. (b)

    Every curve in |D||D| is connected;

  3. (c)

    H1(𝒪(D))=0H^{1}(\mathcal{O}(-D))=0.

Proof.

Let YY be a curve in the linear equivalence class of DD and consider the corresponding short exact sequence

0𝒪(D)𝒪𝒪Y0.0\to\mathcal{O}(-D)\to\mathcal{O}\to\mathcal{O}_{Y}\to 0.

Taking cohomology and using H1(𝒪)=0H^{1}(\mathcal{O})=0, we see that H1(𝒪(D))=0H^{1}(\mathcal{O}(-D))=0 if and only if H0(𝒪)H0(𝒪Y)H^{0}(\mathcal{O})\to H^{0}(\mathcal{O}_{Y}) is an isomorphism, which is true if and only if H0(𝒪Y)H^{0}(\mathcal{O}_{Y}) is one-dimensional, which holds exactly when YY is connected. (If DD is ample, (c)(c) also follows from Kodaira vanishing.) ∎

6. Gaeta resolutions and stability

For XX an admissible blowup, we study the connection between the existence of Gaeta resolutions and stability of a sheaf, which allows Gaeta resolutions to be applied in the study of moduli problems. First, we describe conditions ensuring that general stable sheaves in M(f)M(f) admit Gaeta resolutions (Proposition 1.2). Then, by imposing stronger conditions on ff and on the polarization HH, we show the locus of maps in the resolution space whose cokernels are unstable has codimension 2\geqslant 2 (Proposition 6.5), as well as the parallel statement in the moduli space that the locus of sheaves not admitting Gaeta resolutions has codimension 2\geqslant 2 (Theorem 1.3). In particular, the latter results imply that M(f)M(f) is non-empty and that general stable sheaves away from a locus of codimension 2\geqslant 2 satisfy various nice properties (Corollary 1.4).

We have discussed about prioritary conditions in the previous section. One of the motivations to consider the prioritary condition is the following statement, which is essentially in the proof of [Wal98, Theorem 1].

Lemma 6.1.

Over a smooth projective surface SS, if DD is a divisor and HH is a polarization such that H(KS+D)<0H\cdot(K_{S}+D)<0, then any HH-semistable torsion-free sheaf is DD-prioritary.

Proof.

If FF is torsion-free and HH-semistable, then by Serre duality, Ext2(F,F(D))Hom(F,F(KS+D))0{\rm Ext}^{2}(F,F(-D))\cong{\rm Hom}(F,F(K_{S}+D))\cong 0 by semistability and the fact that H(KS+D)<0H\cdot(K_{S}+D)<0. ∎

As a warm-up, we prove Proposition 1.2. Recall that XX is an admissible blowup of 𝔽e\mathbb{F}_{e}, HH satisfies H(KX+A)<0H\cdot(K_{X}+A)<0, and fK(X)f\in K(X) is a class of rank r>0r>0, Euler characteristic χ0\chi\geqslant 0, and first Chern class satisfying the inequalities in Propositions 4.5(a) and 5.1(b.ii).

Proof of Proposition 1.2.

When non-empty, the moduli space M(f)M(f) is irreducible ([Wal98, Theorem 1]), using the fact that the stack of prioritary sheaves is irreducible. Since semistability is an open property in families, Propositions 4.5(a), 5.6, and Lemma 6.1 show that a general semistable sheaf in M(f)M(f) has a Gaeta resolution. ∎

Using Proposition 4.5(b), we obtain a formulation similar to Proposition 1.2.

Proposition 6.2.

Let XX and HH be as in Proposition 1.2. Assume the class ff has fixed rank r>0r>0 and first Chern class, and its discriminant is sufficiently large. Then there is a line bundle LL pulled back from 𝔽e\mathbb{F}_{e} such that for general semistable sheaves FF of class ff, FLF\otimes L admits a Gaeta resolution.

Since the space RfR_{f} of Gaeta resolutions is rational, we immediately deduce the following special cases of [Bal87, Theorem 2.2].

Corollary 6.3.

For XX, HH, and fK(X)f\in K(X) as in Propositions 1.2 or 6.2, M(f)M(f) is unirational if it is nonempty.

Over Hirzebruch surfaces, similar results were proved in [CH20, Theorem 2.10, Proposition 4.4].

The rest of this section is dedicated to the proof of Theorem 1.3, which requires additional technical conditions on the exponents of the Gaeta resolutions and on the polarization. Before stating these conditions, we set some notation. For simplicity, we write the Gaeta resolutions of the form (b) on XX as

01a1dadϕd+1ad+1d+1ad+1Fϕ0.0\to\mathcal{E}_{1}^{a_{1}}\oplus\cdots\oplus\mathcal{E}_{d}^{a_{d}}\xrightarrow{\phi}\mathcal{E}_{d+1}^{a_{d+1}}\oplus\cdots\oplus\mathcal{E}_{d+1}^{a_{d+1}}\to F_{\phi}\to 0.

Let R^Ha,d{\hat{R}}\subset H_{\vec{a},d} denote the locus of injective maps ϕ\phi as in § 5. Let pp and qq denote the projections from R^×X{\hat{R}}\times X to the first and second factors, respectively. Let 𝔽\mathbb{F} denote the cokernel of the tautological map over R^×X{\hat{R}}\times X:

0i=1dqiaij=d+1nqjaj𝔽0.\displaystyle 0\to\bigoplus_{i=1}^{d}q^{*}\mathcal{E}_{i}^{a_{i}}\to\bigoplus_{j=d+1}^{n}q^{*}\mathcal{E}_{j}^{a_{j}}\to\mathbb{F}\to 0. (6.1)

Because it is the family of cokernels of injective maps, 𝔽\mathbb{F} is flat over R^{\hat{R}}.

The technical conditions on the exponents are as follows. We assume

r2andγij:pjpiγj+1iS0r\geqslant 2\quad\text{and}\quad\gamma_{i}\geqslant\sum_{j\colon p_{j}\succ p_{i}}\gamma_{j}+1\;\quad\forall i\in S_{0} (6.2)

to ensure that the complement of R^{\hat{R}} in Ha,dH_{\vec{a},d} has codimension 2\geqslant 2 and that the cokernels FϕF_{\phi} are torsion-free away from a locus of codimension 2\geqslant 2 in R^{\hat{R}} (the proof is similar to the proof of Proposition 5.1, but we consider the image of the locus ZZ in Proposition 5.2 in Hom(𝒜,)\mathbb{P}{\rm Hom}(\mathcal{A},\mathcal{B})). We make the additional assumption

iS0γi+α4r+1,\sum_{i\in S_{0}}\gamma_{i}+\alpha_{4}\geqslant r+1, (6.3)

which ensures that the FϕF_{\phi} are (C+A)(C+A)-prioritary away from a locus of codimension 2\geqslant 2 in R^{\hat{R}} (Proposition 5.5), require all of the exponents aia_{i} to be strictly positive, and require the discriminant of the sheaves FϕF_{\phi} to be sufficiently large in the sense of (6.9).

To state the conditions on the ample divisor HH, which we assume is general, we write

H=uA+vCiS0S1diEiH=uA+vC-\sum_{i\in S_{0}\cup S_{1}}d_{i}E_{i}

for rational numbers u,v,di>0u,v,d_{i}>0 222As scaling HH does not affect stability, these weights could be taken as integers as well.. We assume that

u>jS1dj and v>iS0di>jS1dj,u>\sum_{j\in S_{1}}d_{j}\mbox{ and }v>\sum_{i\in S_{0}}d_{i}>\sum_{j\in S_{1}}d_{j}, (6.4)

namely that HH is a positive linear combination of all divisors in 𝒟\mathcal{D}, where 𝒟\mathcal{D} is defined in (2.3). Note that by Proposition 2.8, (6.4) implies that HH is ample. This condition implies that

H(KX+D)<0,D𝒟.H\cdot(K_{X}+D)<0,\ \forall D\in\mathcal{D}.

as well as the condition H(KX+2A)<0H\cdot(K_{X}+2A)<0 that appears in Theorem 1.5. Moreover, we assume that no did_{i} can be too close to vv in the sense that

divλλ+1anddiv2λtfor all i, where λ=uv+e2 
.
\frac{d_{i}}{v}\leqslant\frac{\lambda}{\lambda+1}\quad\text{and}\quad\frac{d_{i}}{v}\leqslant\sqrt{\frac{2\lambda}{t}}\quad\text{for all $i$, where $\lambda=\frac{u}{v}+\frac{e}{2}$ \\ }.
(6.5)
Remark 6.4.

Our arguments can be extended to allow 𝒟\mathcal{D} to include other divisors whose linear systems are basepoint-free with general member isomorphic to 1\mathbb{P}^{1}, for instance other divisors of the form C+AiIEiC+A-\sum_{i\in I}E_{i} for I{1,,t}I\subset\{1,\dots,t\} such that #Ie+2\#I\leqslant e+2, which could expand the range of HH depending on the configuration of the blown-up points.

6.1. Locus of unstable sheaves in the space of Gaeta resolutions.

In this subsection we prove the following proposition.

Proposition 6.5.

Assume that a\vec{a} and HH satisfy the conditions (6.2-6.5). Then the complement of {ϕR^Fϕ is H-semistable}\{\phi\in{\hat{R}}\mid F_{\phi}\mbox{ is $H$-semistable}\} in Ha,dH_{\vec{a},d} has codimension 2\geqslant 2. In particular, the moduli space M(f)M(f) is non-empty.

As observed above, the complement of R^{\hat{R}} has codimension at least 22 in Ha,dH_{\vec{a},d}, as does the locus of ϕ\phi such that FϕF_{\phi} is not torsion-free. Moreover, by Proposition 5.5, the locus of ϕ\phi such that FϕF_{\phi} is not (C+A)(C+A)-prioritary also has codimension 2\geqslant 2 in Ha,dH_{\vec{a},d}, so it suffices to show that the locus of unstable torsion-free sheaves that are (C+A)(C+A)-prioritary has codimension at least 2. We will do this by showing various Harder-Narasimhan strata have codimension 2\geqslant 2.

For ϕR^\phi\in{\hat{R}} whose cokernel FϕF_{\phi} is (C+A)(C+A)-prioritary and unstable, let

0=G0G1G=Fϕ0=G_{0}\subsetneqq G_{1}\subsetneqq\cdots\subsetneqq G_{\ell}=F_{\phi}

be the Harder-Narasimhan filtration of FϕF_{\phi} with respect to Gieseker stability, let gri=Gi/Gi1\operatorname{gr}_{i}=G_{i}/G_{i-1} denote the grading and let rir_{i}, νi\nu_{i}, and Δi\Delta_{i} denote the rank, total slope, and discriminant of gri\operatorname{gr}_{i}. Then ν1HνH\nu_{1}\cdot H\geqslant\cdots\geqslant\nu_{\ell}\cdot H and ri1r_{i}\geqslant 1 and Δi0\Delta_{i}\geqslant 0 for all ii.

Lemma 6.6.

The following subset of R^{\hat{R}} has codimension at least 22:

{ϕR^|Fϕ is unstable and (C+A)-prioritary;(νiνj)D>2 for some i<j and some D𝒟{C+A}}.\left\{\phi\in{\hat{R}}\,\middle|\begin{array}[]{l}F_{\phi}\mbox{ is unstable and $(C+A)$-prioritary};\\ \mbox{$(\nu_{i}-\nu_{j})\cdot D>2$ for some $i<j$ and some $D\in\mathcal{D}\cup\{C+A\}$}\end{array}\right\}.
Proof.

All D𝒟{C+A}D\in\mathcal{D}\cup\{C+A\} are basepoint-free, so Bertini’s Theorem implies that general divisors in the corresponding linear systems are nonsingular. Moreover, these general divisors are isomorphic to 1\mathbb{P}^{1}. Let DD be general in the linear system |D||D| so that it avoids the singularities of FϕF_{\phi}. The restrictions of the cokernels in a neighborhood of ϕ\phi, which are locally free on DD. As (C+A)(C+A)-prioritary implies DD-prioritary, FϕF_{\phi} is DD-prioritary, so the Kodaira-Spencer map

TϕR^ExtX1(Fϕ,Fϕ)ExtD1(Fϕ|D,Fϕ|D)\operatorname{T}_{\phi}{\hat{R}}\to{\rm Ext}_{X}^{1}(F_{\phi},F_{\phi})\to{\rm Ext}^{1}_{D}(F_{\phi}|_{D},F_{\phi}|_{D})

is surjective, according to [LP97, Corollary 15.4.4] or [CH20, Proposition 2.6]. The restrictions to DD provide a complete family of vector bundles. On the other hand, the inequality implies that over DD, μmax(Fϕ|D)μmin(Fϕ|D)>2\mu_{\max{}}(F_{\phi}|_{D})-\mu_{\min{}}(F_{\phi}|_{D})>2. Thus, the subset has codimension at least 2, according to [LP97, Corollary 15.4.3]. ∎

The last step is to show that the locus of ϕ\phi in R^{\hat{R}} satisfying the following two conditions has codimension 2\geqslant 2:

  1. (a)

    FϕF_{\phi} is unstable and (C+A)(C+A)-prioritary;

  2. (b)

    For all i<ji<j, the inequality (νiνj)D2(\nu_{i}-\nu_{j})\cdot D\leqslant 2 holds for all D𝒟{C+A}D\in\mathcal{D}\cup\{C+A\}.

We will do this for each locally-closed stratum of this locus of fixed Harder-Narasimhan type, namely for an integer 2\ell\geqslant 2 and polynomials P1,,PP_{1},\dots,P_{\ell}, we let YP1,,PY_{P_{1},\dots,P_{\ell}} denote the locus of ϕ\phi in R^{\hat{R}} such that (a) and (b) hold and the Harder-Narasimhan filtration of FϕF_{\phi} has length \ell and the Hilbert polynomial of gri\operatorname{gr}_{i} is PiP_{i}. Our strategy for showing the codimension of YP1,,PY_{P_{1},\dots,P_{\ell}} in R^{\hat{R}} is 2\geqslant 2 is based on similar ideas for 2\mathbb{P}^{2} in [LP05, Chapter 15].

We begin with the following observation:

Lemma 6.7.
  1. (i)

    For i<ji<j, Hom(gri,grj)=0{\rm Hom}(\operatorname{gr}_{i},\operatorname{gr}_{j})=0.

  2. (ii)

    Under the conditions (a) and (b) above, Ext2(gri,grj)=0{\rm Ext}^{2}(\operatorname{gr}_{i},\operatorname{gr}_{j})=0 for all ii and jj.

Proof.

Part (i) follows from semistability.

For (ii), (b) implies that when i<ji<j, (νiνj+K)D0(\nu_{i}-\nu_{j}+K)\cdot D\leqslant 0 for all D𝒟D\in\mathcal{D} and (νiνj+K)(C+A)e2<0(\nu_{i}-\nu_{j}+K)\cdot(C+A)\leqslant-e-2<0. As HH is a positive linear combination of all divisors in 𝒟\mathcal{D}, letting mm denote the minimum of the weights of AA and CC, HH can be written as m(C+A)m(C+A) plus a non-negative linear combination of divisors in 𝒟\mathcal{D}. Thus, (νiνj+K)H<0(\nu_{i}-\nu_{j}+K)\cdot H<0 and this inequality is also true when iji\geqslant j since (6.4) implies HK<0H\cdot K<0. Thus, Ext2(gri,grj)Hom(grj,griK)=0{\rm Ext}^{2}(\operatorname{gr}_{i},\operatorname{gr}_{j})\cong{\rm Hom}(\operatorname{gr}_{j},\operatorname{gr}_{i}\otimes K)^{\vee}=0 for all ii and jj. ∎

Let Flag=Flag(𝔽/R^;P1,,P)R^\operatorname{Flag}=\operatorname{Flag}(\mathbb{F}/{\hat{R}};P_{1},\dots,P_{\ell})\to{\hat{R}} be the relative flag scheme of filtrations whose grading gri\operatorname{gr}_{i} has Hilbert polynomial PiP_{i}. Given ϕYP1,,P\phi\in Y_{P_{1},\dots,P_{\ell}} and a point p=(ϕ,(G1,,G))p=(\phi,(G_{1},\dots,G_{\ell})) of the fiber over ϕ\phi, there is an exact sequence

0Ext+0(Fϕ,Fϕ)TpFlagTϕR^ω+Ext+1(Fϕ,Fϕ),0\to{\rm Ext}^{0}_{+}(F_{\phi},F_{\phi})\to\operatorname{T}_{p}\operatorname{Flag}\to\operatorname{T}_{\phi}{\hat{R}}\xrightarrow[]{\omega_{+}}{\rm Ext}^{1}_{+}(F_{\phi},F_{\phi}), (6.6)

[LP97, Proposition 15.4.1] realizing the vertical tangent space as the group Ext+0(Fϕ,Fϕ){\rm Ext}^{0}_{+}(F_{\phi},F_{\phi}) and the normal space of YP1,,PY_{P_{1},\dots,P_{\ell}} in R^{\hat{R}} at ϕ\phi as the image of ω+\omega_{+}. Here the groups Ext+i{\rm Ext}^{i}_{+} are defined with respect to the filtration of FϕF_{\phi}, and ω+\omega_{+} is the composite map

TϕR^KSExt1(Fϕ,Fϕ)h+Ext+1(Fϕ,Fϕ).\operatorname{T}_{\phi}{\hat{R}}\xrightarrow{\operatorname{KS}}{\rm Ext}^{1}(F_{\phi},F_{\phi})\xrightarrow{h_{+}}{\rm Ext}^{1}_{+}(F_{\phi},F_{\phi}). (6.7)

The Kodaira-Spencer map KS\operatorname{KS} is surjective since R^{\hat{R}} parametrizes a complete family of prioritary sheaves. The idea is to show that h+h_{+} is also surjective and that ext+1(Fϕ,Fϕ)2{\rm ext}^{1}_{+}(F_{\phi},F_{\phi})\geqslant 2, which imply that YP1,,PY_{P_{1},\dots,P_{\ell}} has codimension 2\geqslant 2 in R^{\hat{R}}. For foundational material on the groups Ext±i{\rm Ext}^{i}_{\pm}, see [DLP85].

There is a canonical exact sequence

Ext1(Fϕ,Fϕ)h+Ext+1(Fϕ,Fϕ)Ext2(Fϕ,Fϕ)\dots\to{\rm Ext}^{1}(F_{\phi},F_{\phi})\xrightarrow{h_{+}}{\rm Ext}^{1}_{+}(F_{\phi},F_{\phi})\to{\rm Ext}^{2}_{-}(F_{\phi},F_{\phi})\to\dots

showing that surjectivity of h+h_{+} is guaranteed by the vanishing of Ext2(Fϕ,Fϕ){\rm Ext}^{2}_{-}(F_{\phi},F_{\phi}). This group can be calculated using the spectral sequence

E1p,q={0if p<0,iExtp+q(gri,grip)otherwise,E_{1}^{p,q}=\begin{cases}0&\text{if $p<0$,}\\ \bigoplus_{i}{\rm Ext}^{p+q}(\operatorname{gr}_{i},\operatorname{gr}_{i-p})&\text{otherwise,}\end{cases}

converging to Extp+q(Fϕ,Fϕ){\rm Ext}^{p+q}_{-}(F_{\phi},F_{\phi}). By Lemma 6.7 (b), E1p,qE_{1}^{p,q} vanishes when p+q=2p+q=2, hence Ext2(Fϕ,Fϕ)=0{\rm Ext}_{-}^{2}(F_{\phi},F_{\phi})=0, so h+h_{+} is surjective.

To calculate Ext+1(Fϕ,Fϕ){\rm Ext}_{+}^{1}(F_{\phi},F_{\phi}), we use the spectral sequence

E1p,q={iExtp+q(gri,grip)if p<0,0otherwise,E_{1}^{p,q}=\begin{cases}\bigoplus_{i}{\rm Ext}^{p+q}(\operatorname{gr}_{i},\operatorname{gr}_{i-p})&\text{if $p<0$,}\\ 0&\text{otherwise,}\end{cases}

converging to Ext+p+q(Fϕ,Fϕ){\rm Ext}^{p+q}_{+}(F_{\phi},F_{\phi}). By Lemma 6.7, Ext+0(Fϕ,Fϕ)=Ext+2(Fϕ,Fϕ)=0{\rm Ext}_{+}^{0}(F_{\phi},F_{\phi})={\rm Ext}_{+}^{2}(F_{\phi},F_{\phi})=0 and the spectral sequence degenerates on the first page, yielding

Ext+1(Fϕ,Fϕ)i<jExt1(gri,grj).{\rm Ext}^{1}_{+}(F_{\phi},F_{\phi})\cong\bigoplus_{i<j}{\rm Ext}^{1}(\operatorname{gr}_{i},\operatorname{gr}_{j}).

Using (4.3), we thus calculate

ext+1(Fϕ,Fϕ)=i<jext1(gri,grj)=i<jχ(gri,grj)=i<jrirj(Δi+ΔjP(νjνi)),{\rm ext}_{+}^{1}(F_{\phi},F_{\phi})=\sum_{i<j}{\rm ext}^{1}(\operatorname{gr}_{i},\operatorname{gr}_{j})=-\sum_{i<j}\chi(\operatorname{gr}_{i},\operatorname{gr}_{j})=\sum_{i<j}r_{i}r_{j}(\Delta_{i}+\Delta_{j}-P(\nu_{j}-\nu_{i})), (6.8)

where PP is the Hilbert polynomial of 𝒪X\mathcal{O}X, as in § 4.3.

To finish the proof of the proposition, we will show that ext+1(Fϕ,Fϕ)2{\rm ext}^{1}_{+}(F_{\phi},F_{\phi})\geqslant 2 given the conditions (νiνj)D2(\nu_{i}-\nu_{j})\cdot D\leqslant 2 for all D𝒟{C+A}D\in\mathcal{D}\cup\{C+A\} and (νiνj)H0(\nu_{i}-\nu_{j})\cdot H\geqslant 0 for all i<ji<j. As χ(gri,grj)0\chi(\operatorname{gr}_{i},\operatorname{gr}_{j})\leqslant 0 for each i<ji<j, we see that

i<jχ(gri,grj)χ(gr1,1<jgrj),-\sum_{i<j}\chi(\operatorname{gr}_{i},\operatorname{gr}_{j})\geqslant-\chi(\operatorname{gr}_{1},\oplus_{1<j}\operatorname{gr}_{j}),

which allows us to reduce to the case where the Harder-Narasimhan filtration has length 2. To see that the conditions (a) and (b) still hold, note that ν(1<jgrj)=1<jrjνj/1<jrj\nu(\oplus_{1<j}\operatorname{gr}_{j})=\sum_{1<j}r_{j}\nu_{j}/\sum_{1<j}r_{j} is a weighted average of the νj\nu_{j}, hence we have (ν1ν(1<jgrj))H0(\nu_{1}-\nu(\oplus_{1<j}\operatorname{gr}_{j}))\cdot H\geqslant 0 and (ν1ν(1<jgrj))D2(\nu_{1}-\nu(\oplus_{1<j}\operatorname{gr}_{j}))\cdot D\leqslant 2 for all D𝒟{C+A}D\in\mathcal{D}\cup\{C+A\}. As we know Δ10\Delta_{1}\geqslant 0 but no such inequality is guaranteed for the discriminant of 1<jgrj\oplus_{1<j}\operatorname{gr}_{j}, we need 1<jrjr1\sum_{1<j}r_{j}\geqslant r_{1} to apply the lemma below; if this does not hold, then a similar setup using i<gri\oplus_{i<\ell}\operatorname{gr}_{i} and gr\operatorname{gr}_{\ell}, which satisfies i<ri>r\sum_{i<\ell}r_{i}>r_{\ell} and Δ0\Delta_{\ell}\geqslant 0, meets the conditions of the lemma. Thus, it suffices to prove the following:

Lemma 6.8.

Assume the condition (6.5) on HH. Suppose the class (r,ν,Δ)(r,\nu,\Delta) is the sum of classes (r1,ν1,Δ1)(r_{1},\nu_{1},\Delta_{1}) and (r2,ν2,Δ2)(r_{2},\nu_{2},\Delta_{2}) of positive rank with the property that (ν1ν2)D2(\nu_{1}-\nu_{2})\cdot D\leqslant 2 for D=A,C,A+CD=A,C,A+C, that (ν1ν2)H0(\nu_{1}-\nu_{2})\cdot H\geqslant 0, that Δ10\Delta_{1}\geqslant 0 if r2r1r_{2}\geqslant r_{1}, and that Δ20\Delta_{2}\geqslant 0 if r1r2r_{1}\geqslant r_{2}. Then the condition

Δ(λ+1)24λ+t8+1r,where λ=uv+e2,\Delta\geqslant\frac{(\lambda+1)^{2}}{4\lambda}+\frac{t}{8}+\frac{1}{r},\qquad\text{where $\lambda=\frac{u}{v}+\frac{e}{2}$}, (6.9)

is sufficient to ensure that r1r2(Δ1+Δ2P(ν2ν1))2r_{1}r_{2}(\Delta_{1}+\Delta_{2}-P(\nu_{2}-\nu_{1}))\geqslant 2.

Proof.

Note that r1+r2=rr_{1}+r_{2}=r, r1ν1+r2ν2=rνr_{1}\nu_{1}+r_{2}\nu_{2}=r\nu, and r1Δ1+r2Δ2=rΔ+r1r22r(ν2ν1)2r_{1}\Delta_{1}+r_{2}\Delta_{2}=r\Delta+\frac{r_{1}r_{2}}{2r}(\nu_{2}-\nu_{1})^{2}. Using this, in the case r2r1r_{2}\geqslant r_{1}, we write

r1r2(Δ1+Δ2P(ν2ν1))\displaystyle r_{1}r_{2}(\Delta_{1}+\Delta_{2}-P(\nu_{2}-\nu_{1})) =r1(r1Δ1+r2Δ2)+(r2r1)r1Δ1r1r2P(ν2ν1)\displaystyle=r_{1}(r_{1}\Delta_{1}+r_{2}\Delta_{2})+(r_{2}-r_{1})r_{1}\Delta_{1}-r_{1}r_{2}P(\nu_{2}-\nu_{1})
=(r2r1)r1Δ1+r1rΔr1r2(r22r(ν2ν1)212(ν2ν1)K+1).\displaystyle=(r_{2}-r_{1})r_{1}\Delta_{1}+r_{1}r\Delta-r_{1}r_{2}(\tfrac{r_{2}}{2r}(\nu_{2}-\nu_{1})^{2}-\tfrac{1}{2}(\nu_{2}-\nu_{1})\cdot K+1).

As Δ10\Delta_{1}\geqslant 0, it suffices to find an upper bound for r22r(ν2ν1)212(ν2ν1)K+1\frac{r_{2}}{2r}(\nu_{2}-\nu_{1})^{2}-\frac{1}{2}(\nu_{2}-\nu_{1})\cdot K+1. Setting ξ=r/r2\xi=r/r_{2} and writing

ν1ν2=aA+bBieiEi,\nu_{1}-\nu_{2}=aA+bB-\sum_{i}e_{i}E_{i},

this equals

ξ1((aξe2b)(bξ)12iei(eiξ))+1ξ.\xi^{-1}\left((a-\xi-\frac{e}{2}b)(b-\xi)-\frac{1}{2}\sum_{i}e_{i}(e_{i}-\xi)\right)+1-\xi.

The lemma below shows that given (6.5), an upper bound is

ξ1(ξ2(λ+1)24λ+tξ28)+1ξ,\xi^{-1}\left(\frac{\xi^{2}(\lambda+1)^{2}}{4\lambda}+\frac{t\xi^{2}}{8}\right)+1-\xi,

where λ=uv+e2>0\lambda=\tfrac{u}{v}+\tfrac{e}{2}>0 since if e=0e=0 then HH ample ensures that u,v>0u,v>0. This yields the bound

r1r2(Δ1+Δ2P(ν2ν1))r1rΔr1r2(ξ(λ+1)24λ+tξ8+1ξ),r_{1}r_{2}(\Delta_{1}+\Delta_{2}-P(\nu_{2}-\nu_{1}))\geqslant r_{1}r\Delta-r_{1}r_{2}\left(\frac{\xi(\lambda+1)^{2}}{4\lambda}+\frac{t\xi}{8}+1-\xi\right),

and, as r2ξ=rr_{2}\xi=r, we can guarantee that the right side is 2\geqslant 2 by assuming Δ\Delta satisfies (6.9).

The argument in the case r1r2r_{1}\geqslant r_{2} is similar. ∎

Before stating and proving the lemma below, we introduce some useful notation. Set λ=uv+e2\lambda=\frac{u}{v}+\frac{e}{2} as above and consider the change of variables

J=a+uvb.J=a+\frac{u}{v}b.

The conditions (ν1ν2)D2(\nu_{1}-\nu_{2})\cdot D\leqslant 2 can be written as a2a\leqslant 2, b2b\leqslant 2, a+b2a+b\leqslant 2, and the condition (ν1ν2)H0(\nu_{1}-\nu_{2})\cdot H\geqslant 0 is JidiveiJ\geqslant\sum_{i}\tfrac{d_{i}}{v}e_{i}. Moreover, thinking of JJ as fixed,

(aξe2b)(bξ)=λb2+(J+ξ(λ1))bξ(Jξ)(a-\xi-\frac{e}{2}b)(b-\xi)=-\lambda b^{2}+(J+\xi(\lambda-1))b-\xi(J-\xi)

is a quadratic function with maximum value

(Jξ(λ+1))24λoccurring atb=J+ξ(λ1)2λ.\frac{(J-\xi(\lambda+1))^{2}}{4\lambda}\qquad\text{occurring at}\quad b=\frac{J+\xi(\lambda-1)}{2\lambda}. (6.10)

In particular, assuming J0J\geqslant 0 and the constraints a2a\leqslant 2, b2b\leqslant 2, and a+b2a+b\leqslant 2, an upper bound is obtained by taking J=0J=0, as the constraints imply J2max{u/v,1}J\leqslant 2\max\{u/v,1\} and this upper bound for JJ yields a smaller value since max{u/v,1}<ξ(λ+1)\max\{u/v,1\}<\xi(\lambda+1).

Lemma 6.9.

Assume HH satisfies (6.5). Given the constraints JidiveiJ\geqslant\sum_{i}\tfrac{d_{i}}{v}e_{i}, a2a\leqslant 2, b2b\leqslant 2, and a+b2a+b\leqslant 2, we have

(aξe2b)(bξ)12i=1tei(eiξ)ξ2(λ+1)24λ+tξ28.(a-\xi-\frac{e}{2}b)(b-\xi)-\frac{1}{2}\sum_{i=1}^{t}e_{i}(e_{i}-\xi)\leqslant\frac{\xi^{2}(\lambda+1)^{2}}{4\lambda}+\frac{t\xi^{2}}{8}.
Proof.

For J0J\geqslant 0, the result follows by bounding the first term on the left side by setting J=0J=0 in (6.10), as well as the fact that 12iei(eiξ)tξ2/8-\frac{1}{2}\sum_{i}e_{i}(e_{i}-\xi)\leqslant t\xi^{2}/8 as the maximum value of ei(eiξ)-e_{i}(e_{i}-\xi) is ξ2/4\xi^{2}/4.

Now suppose that J<0J<0. Then idiveiJ\sum_{i}\tfrac{d_{i}}{v}e_{i}\leqslant J is also negative. By (6.10),

(aξe2b)(bξ)(Jξ(λ+1))24λ=ξ2(λ+1)24λ+ξ(λ+1)(J)2λ+J24λ.(a-\xi-\frac{e}{2}b)(b-\xi)\leqslant\frac{(J-\xi(\lambda+1))^{2}}{4\lambda}=\frac{\xi^{2}(\lambda+1)^{2}}{4\lambda}+\frac{\xi(\lambda+1)(-J)}{2\lambda}+\frac{J^{2}}{4\lambda}.

Using the inequality Jdiv(ei)iIdivei-J\leqslant\sum\tfrac{d_{i}}{v}(-e_{i})\leqslant-\sum_{i\in I}\tfrac{d_{i}}{v}e_{i}, where I{1,,t}I\subset\{1,\dots,t\} is the subset of indices for which eie_{i} is negative, as well as the Cauchy-Schwarz inequality to deduce (iIdivei)2#IiI(div)2ei2(\sum_{i\in I}\frac{d_{i}}{v}e_{i})^{2}\leqslant\#I\sum_{i\in I}(\frac{d_{i}}{v})^{2}e_{i}^{2}, the two terms involving JJ are bounded above by

ξ(λ+1)2λiIdivei+#I4λiI(div)2ei2.-\frac{\xi(\lambda+1)}{2\lambda}\sum_{i\in I}\frac{d_{i}}{v}e_{i}+\frac{\#I}{4\lambda}\sum_{i\in I}\left(\frac{d_{i}}{v}\right)^{2}e_{i}^{2}. (6.11)

Bounding each ei(eiξ)-e_{i}(e_{i}-\xi) where ei0e_{i}\geqslant 0 by its maximum value ξ2/4\xi^{2}/4, we see that 12iei(eiξ)-\frac{1}{2}\sum_{i}e_{i}(e_{i}-\xi) is bounded above by

ξ2iIei12iIei2+(t#I)ξ28.\frac{\xi}{2}\sum_{i\in I}e_{i}-\frac{1}{2}\sum_{i\in I}e_{i}^{2}+\frac{(t-\#I)\xi^{2}}{8}. (6.12)

As ei<0e_{i}<0 for iIi\in I, the conditions on di/vd_{i}/v in (6.5) imply that the sum of (6.11) and (6.12) is bounded by (t#I)ξ2/8(t-\#I)\xi^{2}/8, which completes the proof. ∎

By a similar argument, we can show the following statement whose proof will be sketched. The statement is similar to [LP97, Corollary 15.4.6], but the proof is more involved since the Picard group is not as simple as \mathbb{Z}.

Lemma 6.10.

Suppose HH is an ample divisor satisfying (6.4) and (6.5). Consider a complete family of {Fy}yY\{F_{y}\}_{y\in Y} of semistable sheaves of a fixed class on XX, parametrized by a smooth algebraic variety YY. When the discriminant is large, say as in (6.9), the set of points yYy\in Y such that FyF_{y} is strictly semistable forms a closed subset of codimension 2\geqslant 2.

Proof.

For yYy\in Y such that FyF_{y} is strictly semistable, consider one of its Jordan-Hölder filtrations and let gri\operatorname{gr}_{i} be the corresponding sub-quotients and νi=ν(gri)\nu_{i}=\nu(\operatorname{gr}_{i}). The proof of Lemma 6.6 also shows that the set

{yY|Fy is strictly semistable and (νiνj)D>2 for some i,j and some D𝒟{C+A}}\left\{y\in Y\middle|\begin{array}[]{l}F_{y}\mbox{ is strictly semistable and }\\ \mbox{$(\nu_{i}-\nu_{j})\cdot D>2$ for some $i,j$ and some $D\in\mathcal{D}\cup\{C+A\}$}\end{array}\right\}

has codimension 2\geqslant 2.

We next consider yYy\in Y such that (a) FyF_{y} is strictly semistable and (b) (νiνj)D2(\nu_{i}-\nu_{j})\cdot D\leqslant 2, D𝒟{C+A}\forall D\in\mathcal{D}\cup\{C+A\}. Note that Ext2(gri,grj)Hom(grj,griKX)=0{\rm Ext}^{2}(\operatorname{gr}_{i},\operatorname{gr}_{j})\cong{\rm Hom}(\operatorname{gr}_{j},\operatorname{gr}_{i}\otimes K_{X})^{\vee}=0, for all i,ji,j. Thus, Ext2(Fy,Fy)=0{\rm Ext}^{2}_{-}(F_{y},F_{y})=0, which is calculated with respect to the fixed Jordan-Hölder filtration. Consider the relative flag scheme Flag\operatorname{Flag} of filtrations of the same type as the Jordan-Hölder-filtration. We have

TpFlagTyYExt+1(Fy,Fy).\operatorname{T}_{p}\operatorname{Flag}\to\operatorname{T}_{y}Y\twoheadrightarrow{}{\rm Ext}_{+}^{1}(F_{y},F_{y}).

Moreover, the vanishing of Ext2(gri,grj){\rm Ext}^{2}(\operatorname{gr}_{i},\operatorname{gr}_{j}) implies Ext+2(Fy,Fy)=0{\rm Ext}^{2}_{+}(F_{y},F_{y})=0. The codimension of set of yYy\in Y satifying conditions (a) and (b) is bounded below by ext+1(Fs,Fs)i<jχ(gri,grj)2{\rm ext}_{+}^{1}(F_{s},F_{s})\geqslant\sum_{i<j}-\chi(\operatorname{gr}_{i},\operatorname{gr}_{j})\geqslant 2. ∎

Using this lemma, we can immediately strengthen Proposition 6.5 replacing “semistable” by “stable”.

6.2. Locus of semistable sheaves not admitting Gaeta resolutions

This subsection is devoted to the proof of Theorem 1.3.

First, the subset ZZ is indeed closed. As in the construction of the moduli space using geometric invariant theory [MFK94], let Quot(𝒪X(m)N,f){\rm Quot}(\mathcal{O}_{X}(-m)^{\oplus N},f) be the Quot scheme such that M(f)M(f) is a good quotient of the semistable locus Quotss(𝒪X(m)N,f){\rm Quot}^{\operatorname{ss}}(\mathcal{O}_{X}(-m)^{\oplus N},f) with respect to the action by GL(N,)\operatorname{GL}(N,\mathbb{C}). According to Proposition 4.4 and upper semicontinuity, the subset of quotient sheaves that do not admit Gaeta resolutions is closed and invariant under the action of GL(N,)\operatorname{GL}(N,\mathbb{C}). Under the good quotient map, the image of this subset is closed and is exactly ZZ. 333If the Jordan-Hölder grading of a semistable sheaf admits a Gaeta resolution, the sheaf does as well.

Let

G=i=1nGL(ai,),G=\prod_{i=1}^{n}\operatorname{GL}(a_{i},\mathbb{C}),

which acts on R^{\hat{R}} ([Ped21, § 4.3]). Let G¯=G/(id,,id)\bar{G}=G/\mathbb{C}^{*}({\rm id},\dots,{\rm id}). There is an induced action of G¯\bar{G} on R^{\hat{R}}. The universal cokernel (6.1) induces a map

λ𝔽:K(X)PicG(R^),wdet(p!([𝔽]qw)).\displaystyle\lambda_{\mathbb{F}}\colon\operatorname{K}(X)\to\operatorname{Pic}^{{G}}({\hat{R}}),\quad w\mapsto\det\left(p_{!}\left([\mathbb{F}]\cdot q^{*}w\right)\right). (6.13)

which will be shown to be an isomorphism.

Since R^{\hat{R}} is an open subset in a vector space and its complement has codimension 2\geqslant 2, PicG(R^)\operatorname{Pic}^{G}({\hat{R}}) is isomorphic to the character group Char(G)n\operatorname{Char}(G)\cong\mathbb{Z}^{n} and PicG¯(R^)Char(G¯)n1\operatorname{Pic}^{\bar{G}}({\hat{R}})\cong\operatorname{Char}(\bar{G})\cong\mathbb{Z}^{n-1}. The character groups can be explicitly described as follows:

n\displaystyle\mathbb{Z}^{n} Char(G),\displaystyle\xrightarrow{\cong}\operatorname{Char}(G),
(x1,,xn)\displaystyle(x_{1},\dots,x_{n}) [(M1,,Mn)i=1ndet(Mi)xi],\displaystyle\mapsto[(M_{1},\dots,M_{n})\mapsto\prod_{i=1}^{n}\det(M_{i})^{x_{i}}],

and under this isomorphism,

Char(G¯)={(x1,,xn)n|i=1naixi=0}.\displaystyle\operatorname{Char}(\bar{G})=\left\{(x_{1},\dots,x_{n})\in\mathbb{Z}^{n}\,\middle|\,\sum_{i=1}^{n}a_{i}x_{i}=0\right\}. (6.14)
Proposition 6.11.

The map λ𝔽\lambda_{\mathbb{F}} in (6.13) is an isomorphism and it induces an isomorphism fPicG¯(R^)f^{\perp}\xrightarrow{\cong}\operatorname{Pic}^{\bar{G}}({\hat{R}}).

This is similar to [LP97, Lemma 18.5.1] and [Ped21, Proposition 4.2].

Proof.

We calculate λ𝔽\lambda_{\mathbb{F}} using the isomorphism PicG(R^)Char(G)n\operatorname{Pic}^{G}({\hat{R}})\cong\operatorname{Char}(G)\cong\mathbb{Z}^{n}. In K(X)\operatorname{K}(X), for j=1,,nj=1,\dots,n, let 𝐞j\mathbf{e}_{j} be the class [j][\mathcal{E}_{j}^{\vee}] of the dual bundle. Then the 𝐞j\mathbf{e}_{j} form a basis of K(X)\operatorname{K}(X). Let VjV_{j} be a complex vector space of dimension aja_{j} for j=1,,nj=1,\dots,n. We can calculate λ𝔽(𝐞j)\lambda_{\mathbb{F}}(\mathbf{e}_{j}) using the GG-equivariant short exact sequence (6.1):

λ𝔽(𝐞j)=det(jid[𝒪R^ViHom(j,i)]+max{j,d+1}in[𝒪R^ViHom(j,i)]).\displaystyle\lambda_{\mathbb{F}}(\mathbf{e}_{j})=\det\left(-\sum_{j\leqslant i\leqslant d}[\mathcal{O}_{\hat{R}}\otimes V_{i}\otimes{\rm Hom}(\mathcal{E}_{j},\mathcal{E}_{i})]+\sum_{\max\{j,d+1\}\leqslant i\leqslant n}[\mathcal{O}_{\hat{R}}\otimes V_{i}\otimes{\rm Hom}(\mathcal{E}_{j},\mathcal{E}_{i})]\right).

The map λ𝔽\lambda_{\mathbb{F}} takes the following matrix form:

[sgn(id12)χ(j,i)]1i,jn.\left[\operatorname{sgn}\left(i-d-\frac{1}{2}\right)\chi(\mathcal{E}_{j},\mathcal{E}_{i})\right]_{1\leqslant i,j\leqslant n}.

Since the matrix is lower triangular with ±1\pm 1 on the diagonal, λ𝔽\lambda_{\mathbb{F}} is an isomorphism. For w=iwi𝐞iK(X)w=\sum_{i}w_{i}\mathbf{e}_{i}\in\operatorname{K}(X), wfw\in f^{\perp} if and only if iwiχ(i,f)=0\sum_{i}w_{i}\chi(\mathcal{E}_{i},f)=0, if and only if λ𝔽(w)Char(G¯)\lambda_{\mathbb{F}}(w)\in\operatorname{Char}(\bar{G}). ∎

Let R^ssR^{\hat{R}}^{\operatorname{ss}}\subset{\hat{R}} denote the subset of cokernels which are semistable. According to Proposition 6.5, R^R^ss{\hat{R}}\setminus{\hat{R}}^{\operatorname{ss}} has codimension 2\geqslant 2 in R^{\hat{R}}. Let R^sR^ss{\hat{R}}^{\operatorname{s}}\subset{\hat{R}}^{\operatorname{ss}} denote the subset of cokernels which are stable. The coarse moduli property provides a map

π:R^sM(f),\pi\colon{\hat{R}}^{\operatorname{s}}\to M(f),

which factors through U=M(f)ZU=M(f)\setminus Z. According to Lemma 6.10, the restriction map induces isomorphisms PicG(R^)PicG(R^s)\operatorname{Pic}^{{G}}({\hat{R}})\cong\operatorname{Pic}^{{G}}({\hat{R}}^{\operatorname{s}}) and PicG¯(R^)PicG¯(R^s)\operatorname{Pic}^{\bar{G}}({\hat{R}})\cong\operatorname{Pic}^{\bar{G}}({\hat{R}}^{\operatorname{s}}). The Donaldson map λM:fPic(M(f))\lambda_{M}\colon f^{\perp}\to\operatorname{Pic}(M(f)) is an isomorphism according to Theorem 1.5, which will be proved at the end of the subsection. By the functoriality of the determinant line bundle construction, we have the following commutative diagram:

f{f^{\perp}}PicG¯(R^){\operatorname{Pic}^{\bar{G}}({\hat{R}})}Pic(M(f)){\operatorname{Pic}(M(f))}Pic(U){\operatorname{Pic}(U)}PicG¯(R^s).{\operatorname{Pic}^{\bar{G}}({\hat{R}}^{\operatorname{s}}).}λ𝔽\scriptstyle{\lambda_{\mathbb{F}}}\scriptstyle{\cong}λM\scriptstyle{\lambda_{M}}\scriptstyle{\cong}\scriptstyle{\cong}

It is clear from the diagram that the restriction map Pic(M(f))Pic(U)\operatorname{Pic}(M(f))\to\operatorname{Pic}(U) is also injective. Since the polarization is general, M(f)M(f) is locally factorial ([Yos96, Corollary 3.4]). Therefore, ZZ has codimension 2\geqslant 2 in M(f)M(f). We have proven Theorem 1.3.

Proof of Corollary 1.4.

The result follows directly from Propositions 1.2, 4.4(b), 5.1(b), and Theorem 1.3. ∎

We are left to provide

Proof of Theorem 1.5.

Let f=(r,c1,c2)K(S)f=(r,c_{1},c_{2})\in\operatorname{K}(S) be the corresponding class. According to [Yos96, Corollary 3.4], there is a surjective map PicSfPicM(f)\mathbb{Z}\oplus\operatorname{Pic}S\cong f^{\perp}\twoheadrightarrow\operatorname{Pic}M(f). On the other hand, under our assumption on c2c_{2}, PicM(f)\operatorname{Pic}M(f) contains a subgroup PicS\mathbb{Z}\oplus\operatorname{Pic}S, as shown in [HL10, Example 8.1.6]. ∎

7. Strange duality

In this section, we review Le Potier’s strange duality conjecture for rational surfaces over \mathbb{C} and use our study of Gaeta resolutions to prove Theorem 1.6, which states that the strange morphism is injective in various cases on 2\mathbb{P}^{2} and on XX an admissible two-step blowup of 𝔽e\mathbb{F}_{e}. The argument is similar to what was shown on 2\mathbb{P}^{2} in [BGJ16]. We assume Theorem 1.7, which will be proved in § 8.

7.1. Strange morphism

Let SS be a smooth projective rational surface over \mathbb{C} and HH be an ample divisor. Let σ\sigma and ρ\rho denote two classes in the Grothendieck group K(S)\operatorname{K}(S). On K(S)\operatorname{K}(S), there is a pairing given by χ(σρ)\chi(\sigma\cdot\rho).

Let M(σ)M(\sigma) and M(ρ)M(\rho) be the moduli spaces of HH-semistable sheaves of class σ\sigma and ρ\rho respectively. For the moment, suppose there are no strictly semistable sheaves, namely M(σ)=Ms(σ)M(\sigma)=M^{\operatorname{s}}(\sigma), and there is a universal family 𝒲\mathcal{W} over M(σ)×SM(\sigma)\times S. Let FF be a sheaf of class ρ\rho, then we have a determinant line bundle on M(σ)M(\sigma),

Θρ:=det(p!(𝒲LqF)).\Theta_{\rho}:=\det\left({p}_{!}\left(\mathcal{W}\stackrel{{\scriptstyle L}}{{\otimes}}{q}^{*}F\right)\right)^{*}.

Here, p{p} and q{q} are projections from M(σ)×SM(\sigma)\times S to the first and second factors respectively. The isomorphism class of Θρ\Theta_{\rho} does not depend on FF but only on its class in K(S)\operatorname{K}(S), so we are justified in using the subscript ρ\rho. Two universal families may differ by a line bundle pulled back from M(σ)M(\sigma), but if we assume χ(σρ)=0\chi(\sigma\cdot\rho)=0, then they will provide isomorphic determinant line bundles on the moduli space. Thus, from now on, we assume χ(σρ)=0\chi(\sigma\cdot\rho)=0, so that Θρ\Theta_{\rho} is independent of the choice of 𝒲\mathcal{W}.

Even if there does not exist a universal family, we can still define Θρ\Theta_{\rho} by carrying out the construction on the Quot scheme coming from the GIT construction where there is a universal family, and then showing that it descends to M(σ)M(\sigma). If there are strictly semistable sheaves of class σ\sigma, we need to further require ρ{1,h,h2}\rho\in\{1,h,h^{2}\}^{\perp\perp} for h=[𝒪H]K(S)h=[\mathcal{O}_{H}]\in\operatorname{K}(S); see [LP92, (2.9)]. These conditions will be satisfied in our setting. Similarly, we can construct a determinant line bundle

ΘσM(ρ).\Theta_{\sigma}\to M(\rho).

Orthogonal classes σ\sigma and ρ\rho are candidates for strange duality if the moduli spaces M(σ)M(\sigma) and M(ρ)M(\rho) are non-empty, and if the following conditions on pairs (W,F)M(σ)×M(ρ)(W,F)\in M(\sigma)\times M(\rho) are satisfied:

  1. (a)

    h2(WF)=0h^{2}(W\otimes F)=0 and 𝒯or1(W,F)=𝒯or2(W,F)=0\operatorname{\mathcal{T}or}^{1}(W,F)=\operatorname{\mathcal{T}or}^{2}(W,F)=0 for all (W,F)(W,F) away from a codimension 2\geqslant 2 subset in M(σ)×M(ρ)M(\sigma)\times M(\rho), and

  2. (b)

    h0(WF)=0h^{0}(W\otimes F)=0 for some (W,F)(W,F).

Under these conditions, there is a line bundle

Θσ,ρM(σ)×M(ρ)\Theta_{\sigma,\rho}\to M(\sigma)\times M(\rho)

with a canonical section whose zero locus is given by

{(W,F)h0(WF)>0}M(σ)×M(ρ),\{\,(W,F)\mid h^{0}(W\otimes F)>0\,\}\subset M(\sigma)\times M(\rho),

(see [LP05, Proposition 9]). The see-saw theorem implies that

Θσ,ρΘρΘσ,\Theta_{\sigma,\rho}\cong\Theta_{\rho}\boxtimes\Theta_{\sigma},

see [LP05, Lemme 8]. Then, using the Künneth formula, the canonical section of Θσ,ρ\Theta_{\sigma,\rho} induces a linear map

SDσ,ρ:H0(M(ρ),Θσ))H0(M(σ),Θρ)\operatorname{SD}_{\sigma,\rho}:H^{0}(M(\rho),\Theta_{\sigma}))^{*}\rightarrow H^{0}\big{(}M(\sigma),\Theta_{\rho})

that is well-defined up to a non-zero scalar. Following Le Potier, we call this the strange morphism.

Conjecture 7.1 (Le Potier).

If SDσ,ρ\mathrm{SD}_{\sigma,\rho} is nonzero, then it is an isomorphism.

We focus on the case when SS is 2\mathbb{P}^{2} or XX, an admissible blowup of 𝔽e\mathbb{F}_{e}. For the former, we take HH to be the hyperplane class, while for the latter, we assume HH is general and satisfies (6.4, 6.5). Let

ρ=(1,0,1)\rho=(1,0,1-\ell)

be the numerical class of an ideal sheaf of 1\ell\geqslant 1 points, so that M(ρ)=S[]M(\rho)=S^{[\ell]} is the Hilbert scheme of points on SS. Clearly ρ{1,h,h2}\rho\in\{1,h,h^{2}\}^{\perp\perp}. Every numerical class orthogonal to ρ\rho has the form

σ=(r,L,χ=r).\sigma=(r,L,\chi=r\ell).

We assume rr and \ell are fixed, that r2r\geqslant 2, and that

L is sufficiently positive.\text{$L$ is sufficiently positive}.

The condition on rr ensures that general sheaves in M(σ)M(\sigma) are locally free, and it is no restriction in the study of strange duality as strange duality is known over all surfaces in the case when σ\sigma and ρ\rho both have rank one. Assumptions about the positivity of LL are required to apply many results in this paper, and we assume here that LL is sufficiently positive such that the positivity assumptions of every result we need are met. A precise statement of the positivity conditions we impose on LL can be found in the appendix.

In particular, since LL is sufficiently positive, σ\sigma admits Gaeta resolutions by Proposition 4.5(a), the discriminant condition (6.9) holds, and M(σ)M(\sigma) is non-empty by Proposition 6.5, so general sheaves in M(σ)M(\sigma) admit Gaeta resolutions by Proposition 1.2 or Theorem 1.3. In this situation, conditions (a) and (b) on p. 7.1 are satisfied. Let WM(σ)W\in M(\sigma) and IZM(ρ)I_{Z}\in M(\rho) be such that singWZ=\operatorname{sing}W\cap Z=\emptyset, which holds away from codimension r+1r+1. Under this condition 𝒯ori(W,IZ)=0\operatorname{\mathcal{T}or}_{i}(W,I_{Z})=0 for i=1,2i=1,2. Furthermore,

h2(WIZ)=h2(W)=hom(W,KS)h^{2}(W\otimes I_{Z})=h^{2}(W)=\hom(W,K_{S})

vanishes by semistability. Condition (b) also holds by Proposition 5.7. Thus, σ\sigma and ρ\rho are candidates for strange duality, and we will study the strange morphism

SDσ,ρ:H0(S[],Θσ)H0(M(σ),Θρ).\operatorname{SD}_{\sigma,\rho}\colon H^{0}(S^{[\ell]},\Theta_{\sigma})^{*}\to H^{0}(M(\sigma),\Theta_{\rho}).

We begin with the determinant line bundle Θσ\Theta_{\sigma} on the Hilbert scheme of points.

7.2. Determinant line bundles on the Hilbert scheme of points

We first review some general results about line bundles on the Hilbert scheme of points on surfaces. The Hilbert scheme of points, S[]S^{[\ell]}, is a resolution of singularities of the symmetric product S()S^{(\ell)}. The resolution, which we denote by π:S[]S()\pi\colon S^{[\ell]}\to S^{(\ell)}, is called the Hilbert-Chow morphism. Given a line bundle MM on SS, let MM^{\boxtimes\ell} be i=1priM\otimes_{i=1}^{\ell}\operatorname{pr}_{i}^{*}M on the \ell-fold product S=S××SS^{\ell}=S\times\cdots\times S. There is an 𝔖\mathfrak{S}_{\ell}-action on SS^{\ell} such that MM^{\boxtimes\ell} is 𝔖\mathfrak{S}_{\ell}-equivariant. The line bundle MM^{\boxtimes\ell} descends onto S()S^{(\ell)}, giving a line bundle M()M^{(\ell)}. We denote its pullback to S[]S^{[\ell]} as

M=πM().M_{\ell}=\pi^{*}M^{(\ell)}.

Via this construction, we can view PicS\operatorname{Pic}S as a subgroup of PicS[]\operatorname{Pic}S^{[\ell]}, by sending MM to MM_{\ell}. Fogarty [Fog73] showed that under this inclusion

PicS[]PicSE2.\operatorname{Pic}S^{[\ell]}\cong\operatorname{Pic}S\oplus\mathbb{Z}\frac{E}{2}.

Here, EE is the exceptional divisor of the Hilbert-Chow morphism, which parametrizes non-reduced subschemes. Furthermore, if MM is ample, then MM_{\ell} is nef, and the canonical divisor on S[]S^{[\ell]} is KS[]=(KS)K_{S^{[\ell]}}=(K_{S})_{\ell}.

The determinant line bundle on S[]S^{[\ell]} induced by σ\sigma is

Θσ=Lr2E.\Theta_{\sigma}=L_{\ell}-\frac{r}{2}E.

By the Kodaira vanishing theorem, if ΘσKS[]=L(KS)r2E\Theta_{\sigma}-K_{S^{[\ell]}}=L_{\ell}-(K_{S})_{\ell}-\frac{r}{2}E is ample on S[]S^{[\ell]}, then Θσ\Theta_{\sigma} has no higher cohomology. Results of Beltrametti, Sommese, Catanese, and Göttsche [BS91, CG90] show that if MM is a line bundle on SS, then M12EM_{\ell}-\frac{1}{2}E is nef if MM is (1\ell-1)-very ample and very ample if MM is \ell-very ample. Thus, we deduce the following:

Lemma 7.2.

Suppose LL is sufficiently positive, for instance LKSL-K_{S} is the tensor product of an \ell-very ample line bundle and r1r-1 (\ell-1)-very ample line bundles. Then Θσ\Theta_{\sigma} has vanishing higher cohomology.

Thus, the vanishing of the higher cohomology of Θσ\Theta_{\sigma} follows from the assumption that LL is sufficiently positive. A precise statement of a sufficient condition on LL is (A.5) in the appendix.

7.3. Injectivity of the strange morphism

To prove Theorem 1.6, we will make use of a Quot scheme argument. As above, let σ=(r,L,r)\sigma=(r,L,r\ell) and ρ=(1,0,1)\rho=(1,0,1-\ell) be orthogonal classes, where 1\ell\geqslant 1 and r2r\geqslant 2 are fixed, so sheaves of class σ\sigma are expected to be locally free, and we assume LL is sufficiently positive in the sense explained in the appendix. Consider the class

v=σ+ρ.v=\sigma+\rho.

The first part of the argument is to show that if VV is a vector bundle of class vv that admits a general Gaeta resolution, then Quot(V,ρ){\rm Quot}(V^{*},\rho) is finite and reduced, which will be proved in § 8. The strategy is to show that the relative Quot scheme over the space of Gaeta resolutions has relative dimension 0. This is known for 2\mathbb{P}^{2} by [BGJ16], so we write the argument for XX, an admissible blowup of 𝔽e\mathbb{F}_{e}, though it works for 2\mathbb{P}^{2} as well. Since σ\sigma admits Gaeta resolutions, an easy calculation using Proposition 4.5(a) shows that vv also admits Gaeta resolutions, with the same exponents γi\gamma_{i} and γj\gamma_{j} and with α1,α2,α3\alpha_{1},\alpha_{2},\alpha_{3} each larger by \ell and α4\alpha_{4} smaller by 1\ell-1.

The starting point is the following result.

Lemma 7.3.

There is a vector bundle VV of class vv such that VV has a Gaeta resolution and Quot(V,ρ){\rm Quot}(V^{*},\rho) contains an isolated point.

Proof.

Choose a vector bundle WW of class σ\sigma that admits a Gaeta resolution and a general ideal sheaf IZI_{Z} of class ρ\rho such that H0(WIZ)=0H^{0}(W\otimes I_{Z})=0, which is possible by Proposition 5.7. Since h0(W)=rh^{0}(W)=r\ell\geqslant\ell and ZZ is general, we can choose a quotient W𝒪ZW\to\mathcal{O}_{Z} such that H0(W)H0(𝒪Z)H^{0}(W)\to H^{0}(\mathcal{O}_{Z}) is surjective. Let JJ denote the kernel of W𝒪ZW\to\mathcal{O}_{Z}. Since LL is sufficiently positive, Ext1(J,𝒪){\rm Ext}^{1}(J,\mathcal{O}) is large (for instance, (A.1) suffices). Then a general extension VV of JJ by 𝒪\mathcal{O} is locally free (see the proof of [BGJ16, Lemma 5.9]) and there are short exact sequences

0JW𝒪Z0,\displaystyle 0\to J\to W\to\mathcal{O}_{Z}\to 0,
0𝒪VJ0,\displaystyle 0\to\mathcal{O}\to V\to J\to 0,

which together give a long exact sequence 0𝒪VW𝒪Z00\to\mathcal{O}\to V\to W\to\mathcal{O}_{Z}\to 0 whose dual

0WVIZ00\to W^{*}\to V^{*}\to I_{Z}\to 0

is a point of Quot(V,ρ){\rm Quot}(V^{*},\rho). As the tangent space at this point is Hom(W,IZ)H0(WIZ)=0{\rm Hom}(W^{*},I_{Z})\cong H^{0}(W\otimes I_{Z})=0, this is an isolated point of Quot(V,ρ){\rm Quot}(V^{*},\rho). Thus, all that remains is to show that VV admits a Gaeta resolution.

We do this in two steps using Proposition 4.4, by first showing that JJ admits a Gaeta resolution. First, we see that Hp(J)=0H^{p}(J)=0 for p=1,2p=1,2 using the corresponding vanishings for WW and the fact that H0(W)H0(𝒪Z)H^{0}(W)\to H^{0}(\mathcal{O}_{Z}) is surjective. Similarly, H1(J(Ei))=0H^{1}(J(E_{i}))=0 for all iS0S1i\in S_{0}\cup S_{1}, as the induced map H0(W(Ei))H0(𝒪Z(Ei))H^{0}(W(E_{i}))\to H^{0}(\mathcal{O}_{Z}(E_{i})) is still surjective. Finally, the vanishing of Hp(J(D))H^{p}(J(D)) for p1p\neq 1 and DD one of the divisors appearing in Proposition 4.4 (b.ii) follows from the same vanishing for WW and the vanishing of the higher cohomology of 𝒪Z\mathcal{O}_{Z}.

Second, since Hp(𝒪)=0H^{p}(\mathcal{O})=0 for p=1,2p=1,2, Hp(𝒪Ei)=0H^{p}(\mathcal{O}_{E_{i}})=0. Using the vanishings for JJ, we obtain that Hp(V)=0H^{p}(V)=0 for p=1,2p=1,2 and that H1(V(Ei))=0H^{1}(V(E_{i}))=0 for all iS0S1i\in S_{0}\cup S_{1}. For each divisor DD appearing in Proposition 4.4 (b.ii), Hp(𝒪(D))=0H^{p}(\mathcal{O}(D))=0 for all pp, so Hp(V(D))Hp(J(D))H^{p}(V(D))\cong H^{p}(J(D)) and this vanishes for p1p\neq 1. Therefore, VV also admits a Gaeta resolution. ∎

The lemma establishes that the relative Quot scheme contains a point with vanishing relative tangent space, which can be deformed to give an open set with this property. The main technical task is to prove that the relative Quot scheme cannot have any other components of the same dimension, which is carried out in the next section for S=XS=X, and which was done for 2\mathbb{P}^{2} in [BGJ16]. The proof requires strong positivity assumptions on LL.

The second part of the argument is to count the points of Quot(V,ρ){\rm Quot}(V^{*},\rho) and compare the result to h0(S[],Θσ)h^{0}(S^{[\ell]},\Theta_{\sigma}). In previous work [GL22], we showed that if the Quot scheme is finite and reduced, then the number of points is

#Quot(V,ρ)=S[]c2(V[]),\#{\rm Quot}(V^{*},\rho)=\int_{S^{[\ell]}}c_{2\ell}({V}^{[\ell]}),

where V[]{V}^{[\ell]} is the tautological vector bundle defined as in (1.2), whose rank is (r+1)(r+1)\ell and whose fiber at a point [IZ]S[][I_{Z}]\in S^{[\ell]} is H0(V𝒪Z)H^{0}(V\otimes\mathcal{O}_{Z}). By the following theorem, this top Chern class calculates the Euler characteristic of the determinant line bundle:

Theorem 7.4.

Let SS be a smooth projective surface with χ(𝒪S)=1\chi(\mathcal{O}_{S})=1. Then

S[]c2(V[])=χ(S[],Θσ).\int_{S^{[\ell]}}c_{2\ell}({V}^{[\ell]})=\chi(S^{[\ell]},\Theta_{\sigma}).

Over Enriques surfaces, this statement was proved by Marian-Oprea-Pandharipande [MOP22, Proposition 3.2] . In general, it was conjectured by Johnson and shown to be equivalent to another conjecture ([Joh18, Conjecture 1.3, Theorem 4.1]). The second conjecture was proven by the works of Marian-Opera-Pandharipande [MOP19, MOP22] and Göttsche-Mellit [GM22] combined. See for example [GM22, Corollary 1.2].

Now, using the fact that LL sufficiently positive ensures that Θσ\Theta_{\sigma} has vanishing higher cohomology, we can deduce the injectivity of the strange morphism.

Proof of Theorem 1.6.

Above, we checked the conditions for the strange morphism SDσ,ρ\mathrm{SD}_{\sigma,\rho} to be well-defined. Let VV be a vector bundle of class σ+ρ\sigma+\rho that admits a general Gaeta resolution. By Theorem 1.7, Quot(V,ρ){\rm Quot}(V^{*},\rho) is finite and reduced and the points of Quot(V,ρ){\rm Quot}(V^{*},\rho) are short exact sequences

0WiVIZi00\to W_{i}^{*}\to V^{*}\to I_{Z_{i}}\to 0

for sheaves WiW_{i} of class σ\sigma that are locally free and semistable. Since the Quot scheme is finite and reduced, the tangent space Hom(Wi,IZi)H0(WiIZi){\rm Hom}(W_{i}^{*},I_{Z_{i}})\cong H^{0}(W_{i}\otimes I_{Z_{i}}) is 0, while if iji\neq j then Hom(Wi,IZj)0{\rm Hom}(W_{i}^{*},I_{Z_{j}})\neq 0 follows from semistability, as otherwise the induced map WiVIZjW_{i}^{*}\to V^{*}\to I_{Z_{j}} would be zero, hence WiVW_{i}^{*}\to V^{*} would factor through WjVW_{j}^{*}\to V^{*}, yielding an equality Wi=WjW_{i}^{*}=W_{j}^{*} as subsheaves of VV^{*}, identifying the two points of the Quot scheme. It follows that the hyperplanes in H0(M(ρ),Θσ)H^{0}(M(\rho),\Theta_{\sigma}) determined by the points IZiI_{Z_{i}} map under SDσ,ρ\mathrm{SD}_{\sigma,\rho} to linearly independent lines ΘZiH0(M(σ),Θρ)\Theta_{Z_{i}}\in\mathbb{P}H^{0}(M(\sigma),\Theta_{\rho}). Thus, the rank of SDσ,ρ\mathrm{SD}_{\sigma,\rho} is at least

#Quot(V,ρ)=S[]c2(V[])=χ(S[],Θσ)=h0(S[],Θσ),\#{\rm Quot}(V^{*},\rho)=\int_{S^{[\ell]}}c_{2\ell}({V}^{[\ell]})=\chi(S^{[\ell]},\Theta_{\sigma})=h^{0}(S^{[\ell]},\Theta_{\sigma}),

namely SDσ,ρ\mathrm{SD}_{\sigma,\rho} is injective. ∎

8. Finite Quot schemes

This section is devoted to the proof of Theorem 1.7. We will show that for XX an admissible blowup of 𝔽e\mathbb{F}_{e} over kk, under appropriate conditions, Quot schemes are indeed finite and reduced. This is an extension of the corresponding result in [BGJ16] for 2\mathbb{P}^{2}. We then apply our previous work [GL22] to enumerate the finite Quot scheme.

We first sketch the ideas. For a vector bundle VV admitting a Gaeta resolution, instead of directly studying ideal sheaf quotients VIZV^{*}\twoheadrightarrow I_{Z}, we replace VV^{*} by the dual of the Gaeta resolution of VV and IZI_{Z} by the canonical map 𝒪𝒪Z\mathcal{O}\twoheadrightarrow\mathcal{O}_{Z}. Namely, we consider commutative diagrams of the form (8.2), which we can view as a family over an open subset RR of the resolution space RvR_{v} (8.3). We prove in § 8.2 that the main component of the family has the same dimension as RR, hence we deduce that an open subscheme of the main component is isomorphic to a relative Quot scheme over an open subset of RR. The fibers of the relative Quot scheme have dimension 0, and generically the relative Zariski tangent space has dimension 0. We conclude that when VV is general with appropriate numerical constraints, the Quot scheme is finite and reduced.

We prove Theorem 1.7 in § 8.1, except for leaving the technical dimension count argument for § 8.2.

8.1. Finite Quot schemes

Over XX an admissible blowup of 𝔽e\mathbb{F}_{e}, as in the previous section, let ρ=(1,0,1)\rho=(1,0,1-\ell) be the class of an ideal sheaf of \ell points for some fixed 1\ell\geqslant 1 and σ=(r,L,χ=r)\sigma=(r,L,\chi=r\ell) be an orthogonal class for some fixed r2r\geqslant 2 and LL satisfying the positivity conditions in the appendix. In particular, the need for the positivity condition (A.2) will be seen in the proof of Proposition 8.1. Then v=σ+ρv=\sigma+\rho admits Gaeta resolutions, with exponents denoted α1,{γj}jS1,α2,α3,{γi}iS0,α4\alpha_{1},\{\gamma_{j}\}_{j\in S_{1}},\alpha_{2},\alpha_{3},\{\gamma_{i}\}_{i\in S_{0}},\alpha_{4} as in (b). Since LL is sufficiently positive, all these exponents are large except for α4=(r1)+1\alpha_{4}=(r-1)\ell+1.

8.1.1. Morphisms vs. chain maps

Consider a vector bundle VV of class vv that admits a general Gaeta resolution. Then VV is locally free, so its dual has a resolution

0VΛ𝜑Ω0,0\to V^{*}\to\Lambda\xrightarrow{\varphi}\Omega\to 0, (8.1)

where

Λ\displaystyle\Lambda =𝒪α4iS0𝒪(Ei)γi𝒪(A)α3𝒪(C)α2, and\displaystyle=\mathcal{O}^{\alpha_{4}}\oplus\bigoplus_{i\in S_{0}}\mathcal{O}(E_{i})^{\gamma_{i}}\oplus\mathcal{O}(A)^{\alpha_{3}}\oplus\mathcal{O}(C)^{\alpha_{2}},\mbox{ and }
Ω\displaystyle\Omega =jS1𝒪(C+AEj)γj𝒪(C+A)α1.\displaystyle=\bigoplus_{j\in S_{1}}\mathcal{O}(C+A-E_{j})^{\gamma_{j}}\oplus\mathcal{O}(C+A)^{\alpha_{1}}.

We wish to study quotients of VV^{*} of class ρ\rho, which are expected to be ideal sheaves of points. Instead of considering maps VIZV^{*}\twoheadrightarrow I_{Z}, we replace them by chain maps of complexes using (8.1) and the short exact sequence 0IZ𝒪1Z𝒪Z00\to I_{Z}\to\mathcal{O}\xrightarrow{1_{Z}}\mathcal{O}_{Z}\to 0. Namely, we consider commutative diagrams

Λ\textstyle{\Lambda\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}φ\scriptstyle{\varphi}π\scriptstyle{\pi}Ω\textstyle{\Omega\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ\scriptstyle{\psi}𝒪\textstyle{\mathcal{O}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1Z\scriptstyle{1_{Z}}𝒪Z\textstyle{\mathcal{O}_{Z}} (8.2)

where φ\varphi is surjective and ZZ has length \ell. Letting RRvR\subset R_{v} denote the open subset of the space of Gaeta resolutions for which the cokernel VV is locally free, dualization yields an inclusion

RHom(Λ,Ω)=:R\subset\mathbb{P}{\rm Hom}(\Lambda,\Omega)=:\mathbb{P} (8.3)

and a universal sequence

0𝒱πXΛπR𝒪R(1)πXΩ0,0\to\mathcal{V}^{*}\to\pi_{X}^{*}\Lambda\to\pi_{R}^{*}\mathcal{O}_{R}(1)\otimes\pi_{X}^{*}\Omega\to 0, (8.4)

where πR\pi_{R} and πX\pi_{X} denote the projections from R×XR\times X to the factors.

Let Ω[]\Omega^{*[\ell]} be the Fourier-Mukai transform of Ω\Omega^{*} over X[]X^{[\ell]} and

Ξ={(φ,ψ,π)ψφ=1Zπ up to a non-zero scalar}×(Ω[])×α41\Xi=\{(\varphi,\psi,\pi)\mid\psi\circ\varphi=1_{Z}\circ\pi\mbox{ up to a non-zero scalar}\}\subset\mathbb{P}\times\mathbb{P}(\Omega^{*[\ell]})\times\mathbb{P}^{\alpha_{4}-1}

be the locus of diagrams (8.2), which are commutative, with the reduced induced scheme structure. We will show in Proposition 8.1 that

dimΞ=dim.\dim\Xi=\dim\mathbb{P}.

8.1.2. Relative Quot schemes

On the other hand, we consider the relative Quot scheme QuotπR:=QuotπR(𝒱,ρ){\rm Quot}_{\pi_{R}}:={\rm Quot}_{\pi_{R}}(\mathcal{V}^{*},\rho). Letting

UQQuotπRU_{Q}\subset{\rm Quot}_{\pi_{R}}

denote the subset where the relative Zariski tangent space has dimension 0, UQU_{Q} is open (by upper semicontinuity), non-empty (by Lemma 7.3), and smooth.

We claim that quotients in UQU_{Q} can only be ideal sheaves of points. A sheaf FF of class ρ\rho could take the following forms: FF is isomorphic to an ideal sheaf IZI_{Z}, FF contains dimension 0 torsion, or FF contains dimension 11 torsion. The second case cannot occur as it violates our assumption on the relative tangent space. In the third case, let TT denote the torsion subsheaf of FF, so that F/TF/T is torsion free. Then the quotient FF/TF\twoheadrightarrow F/T provides a nonzero morphism V𝒪(D)V^{*}\to\mathcal{O}(-D) where DD is a non-trivial effective curve given by c1(T)c_{1}(T). But this cannot happen under the conditions in Proposition 5.8, which are guaranteed when LL is sufficiently positive, see the appendix.

8.1.3. Finite Quot schemes

We next define a regular map

ι:UQΞ\iota\colon U_{Q}\to\Xi

over \mathbb{P} by associating to each quotient VIZV^{*}\twoheadrightarrow I_{Z} a diagram of the form (8.2). It is enough to define it on the level of functor of points. Furthermore, it is enough to consider morphisms from an affine scheme. Let SS be the spectrum of some kk-algebra. A morphism s:SQuotπRs\colon S\to{\rm Quot}_{\pi_{R}} is equivalent to a family of quotients sX𝒱𝒵s_{X}^{*}\mathcal{V}^{*}\to\mathcal{I}_{\mathscr{Z}} over S×XS\times X. Here, sX=s×idXs_{X}=s\times{\rm id}_{X} and 𝒵\mathcal{I}_{\mathscr{Z}} is the ideal sheaf of a subscheme 𝒵S×X\mathscr{Z}\subset S\times X. We denote the projection maps on S×XS\times X by πS\pi_{S} and πX\pi_{X}. Applying the functor omπS(,𝒪S×X)\mathcal{H}om_{\pi_{S}}(-,\mathcal{O}_{S\times X}) to the pull-back of the sequence (8.4) via sXs_{X}, we have the following exact sequence

0\displaystyle 0 omπS(πXΩ,𝒪S×X)omπS(πXΛ,𝒪S×X)omπS(sX𝒱,𝒪S×X)\displaystyle\to\mathcal{H}om_{\pi_{S}}(\pi_{X}^{*}\Omega,\mathcal{O}_{S\times X})\to\mathcal{H}om_{\pi_{S}}(\pi_{X}^{*}\Lambda,\mathcal{O}_{S\times X})\to\mathcal{H}om_{\pi_{S}}(s_{X}^{*}\mathcal{V}^{*},\mathcal{O}_{S\times X})
xtπS1(πXΩ,𝒪S×X).\displaystyle\to\mathcal{E}xt^{1}_{\pi_{S}}(\pi_{X}^{*}\Omega,\mathcal{O}_{S\times X}).

According to our choice of Ω\Omega, the first term and the last term are zero. Therefore, the family sX𝒱𝒵s_{X}^{*}\mathcal{V}\to\mathcal{I}_{\mathscr{Z}} can be completed to a commutative diagram

0{0}sX𝒱{s_{X}^{*}\mathcal{V}^{*}}πXΛ{\pi_{X}^{*}\Lambda}πXΩ{\pi_{X}^{*}\Omega}0{0}0{0}𝒵{\mathcal{I}_{\mathscr{Z}}}𝒪S×X{\mathcal{O}_{S\times X}}𝒪𝒵{\mathcal{O}_{\mathscr{Z}}}0{0}

Then the square on the right provides a morphism SΞS\to\Xi. Clearly, the square uniquely determines the left-most vertical morphism. We have obtained an injective morphism ι:UQΞ\iota\colon U_{Q}\to\Xi whose image is contained in the unique component with the maximal dimension.

The complement Ξimι\Xi\setminus{\rm im\,}\iota of the image has dimension <dim<\dim\mathbb{P}. Then URU\subset R, the complement of the image of ΞimιR\Xi\setminus{\rm im\,}\iota\to R, is non-empty and open in RR. For each sheaf VV parametrized by UU, VV^{*} has an isolated quotient. By the definition of UU, each fiber of Ξ\Xi over [V]U[V]\in U, which parametrizes all non-zero maps of the form VIZV^{*}\to I_{Z}, is entirely contained in the image of ι\iota. In particular, the maps have to be surjective. Therefore, ι\iota induces an isomorphism UQ|UΞ|UU_{Q}|_{U}\to\Xi|_{U}.

We have thus proved Theorem 1.7(a) for every sheaf VV parametrized by UU.

8.1.4. Genericity of kernels and cokernels

We have seen that Quot schemes for VV in the nonempty open set URvU\subset R_{v} are finite and reduced and that the quotient sheaves are all ideal sheaves IZI_{Z}. Moreover, in the proof of Lemma 7.3, the isolated point [VIZ][V^{*}\twoheadrightarrow I_{Z}] of Quot(V,ρ){\rm Quot}(V^{*},\rho) has the property that ZZ is general, hence we can shrink UU further if necessary to ensure all the ideal sheaves arising as quotients are general. Similarly, by Proposition 1.2, general sheaves in M(σ)M(\sigma) admit Gaeta resolutions, and since M(σ)M(\sigma) is nonempty, we may choose WW in the proof of Lemma 7.3 to be semistable, hence the isolated quotient [VIZ][V\twoheadrightarrow I_{Z}] has kernel WW^{*}, which is also semistable. As semistability is an open condition in families, and as the relative Quot scheme can be viewed as a family of sheaves with invariants σ\sigma^{*}, there is a non-empty open set in the relative Quot scheme of quotients for which the kernel is semistable. Shrinking UU if necessary, we may assume that all kernels of the finite Quot schemes are semistable. This proves Theorem 1.7(b).

8.1.5. Length of finite Quot schemes

According to [GL22, Proposition 1.2], when the Quot scheme Quot(V,ρ){\rm Quot}(V^{*},\rho) is finite and reduced, it is isomorphic to the moduli space S(V,ρ)S(V^{*},\rho) of limit stable pairs, which consist of torsion free sheaves FF of class ρ\rho together with nonzero morphisms VFV^{*}\to F. Then the virtual fundamental class of S(V,ρ)S(V^{*},\rho) agrees with the fundamental class, and its degree is as stated in Theorem 1.7(c), by [GL22, Theorem 1.1]. We refer the reader to [Lin18] for foundational material on the moduli space of limit stable pairs.

8.2. Dimension count

We prove that dimΞ=dim\dim\Xi=\dim\mathbb{P} by counting diagrams of the form (8.2) for fixed π\pi (nonzero) and ψ\psi (surjective). For fixed π\pi and ψ\psi, the locus of φ\varphi is (ψ)1(1Zπ)(\psi_{*})^{-1}(1_{Z}\circ\pi) in the exact sequence

0Hom(Λ,kerψ)Hom(Λ,Ω)ψHom(Λ,𝒪Z)Ext1(Λ,kerψ)0.0\to{\rm Hom}(\Lambda,\ker\psi)\to{\rm Hom}(\Lambda,\Omega)\xrightarrow{\psi_{*}}{\rm Hom}(\Lambda,\mathcal{O}_{Z})\to{\rm Ext}^{1}(\Lambda,\ker\psi)\to 0.

This locus, if it is non-empty, is an affine space in =Hom(Λ,Ω)\mathbb{P}=\mathbb{P}{\rm Hom}(\Lambda,\Omega) that is isomorphic to Hom(Λ,kerψ){\rm Hom}(\Lambda,\ker\psi). Since π\pi is fixed, we may assume it is nonzero only on the first 𝒪\mathcal{O}-summand of Λ\Lambda. For the locus to be non-empty, we need 1Z1_{Z} to be in the image of the middle map of

0Hom(𝒪,kerψ)Hom(𝒪,Ω)Hom(𝒪,𝒪Z)Ext1(𝒪,kerψ)0.0\to{\rm Hom}(\mathcal{O},\ker\psi)\to{\rm Hom}(\mathcal{O},\Omega)\to{\rm Hom}(\mathcal{O},\mathcal{O}_{Z})\to{\rm Ext}^{1}(\mathcal{O},\ker\psi)\to 0.

Thus, for the purpose of a dimension count we may assume that the image of Hom(𝒪,Ω)Hom(𝒪,𝒪Z){\rm Hom}(\mathcal{O},\Omega)\to{\rm Hom}(\mathcal{O},\mathcal{O}_{Z}) contains 1Z1_{Z}.

The idea is to control the dimension of Ext1(Λ,kerψ){\rm Ext}^{1}(\Lambda,\ker\psi). In particular, if Ext1(Λ,kerψ)=0{\rm Ext}^{1}(\Lambda,\ker\psi)=0, which is true for general ψ\psi as we show below, then the dimension of (ψ)1(1Zπ)(\psi_{*})^{-1}(1_{Z}\circ\pi) is hom(Λ,Ω)hom(Λ,𝒪Z)\hom(\Lambda,\Omega)-\hom(\Lambda,\mathcal{O}_{Z}). Adding to this the dimension of the parameter spaces for π\pi and ψ\psi, we get

hom(Λ,Ω)hom(Λ,𝒪Z)+(α41)+(2+hom(Ω,𝒪Z)1).\hom(\Lambda,\Omega)-\hom(\Lambda,\mathcal{O}_{Z})+(\alpha_{4}-1)+(2\ell+\hom(\Omega,\mathcal{O}_{Z})-1). (8.5)

Using the fact that hom(Λ,𝒪Z)=rk(Λ)\hom(\Lambda,\mathcal{O}_{Z})=\ell\,\mathrm{rk}(\Lambda), hom(Ω,𝒪Z)=rk(Ω)\hom(\Omega,\mathcal{O}_{Z})=\ell\,\mathrm{rk}(\Omega), rk(Λ)rk(Ω)=r+1{\rm rk\,}(\Lambda)-{\rm rk\,}(\Omega)=r+1, and α4=(r1)+1\alpha_{4}=(r-1)\ell+1, we see that (8.5) equals hom(Λ,Ω)1\hom(\Lambda,\Omega)-1, which equals dim\dim\mathbb{P}. Thus, the main technical difficulty is to control the dimension of Ext1(Λ,kerψ){\rm Ext}^{1}(\Lambda,\ker\psi) in the case when ψ\psi is not general.

Proposition 8.1.

Let

M=max{m(+r+1m)1m}M=\max\{\,m(\ell+r+1-m)\mid 1\leqslant m\leqslant\ell\,\}

and assume

α2,α3M,γjM+ for jS1,andγiM++j:pjpiγj for iS0.\alpha_{2},\alpha_{3}\geqslant M,\quad\gamma_{j}\geqslant M+\ell\mbox{ for }j\in S_{1},\quad\text{and}\quad\gamma_{i}\geqslant M+\ell+\sum_{j\colon p_{j}\succ p_{i}}\gamma_{j}\mbox{ for }i\in S_{0}.

Then the space Ξ\Xi of commutative diagrams has the expected dimension, namely dimΞ=dim\dim\Xi=\dim\mathbb{P}. Furthermore, there is only one component with this dimension, while other components have strictly smaller dimension.

Proof.

Choose an open set UXU\subset X containing ZZ and trivializations on UU of 𝒪(C+A)\mathcal{O}(C+A) and 𝒪(C+AEj)\mathcal{O}(C+A-E_{j}) for jS1j\in S_{1}. The components of the map ψ:Ω𝒪Z\psi\colon\Omega\to\mathcal{O}_{Z} generate subspaces of Hom(𝒪(C+A),𝒪Z){\rm Hom}(\mathcal{O}(C+A),\mathcal{O}_{Z}) and Hom(𝒪(C+AEj),𝒪Z){\rm Hom}(\mathcal{O}(C+A-E_{j}),\mathcal{O}_{Z}), which we identify with subspaces T,Ts+1,,TtT,T_{s+1},\dots,T_{t} of H0(𝒪Z)H^{0}(\mathcal{O}_{Z}) using the trivializations on UU. For D=0,E1,,Es,A,D=0,E_{1},\dots,E_{s},A, or CC, we calculate bounds on ext1(𝒪(D),kerψ){\rm ext}^{1}(\mathcal{O}(D),\ker\psi) by considering the dimension of the image of

Hom(𝒪(D),Ω)Hom(𝒪(D),𝒪Z).{\rm Hom}(\mathcal{O}(D),\Omega)\to{\rm Hom}(\mathcal{O}(D),\mathcal{O}_{Z}).

Picking a trivialization of 𝒪(D)\mathcal{O}(D) on UU (shrinking UU if necessary), this image is the image of the multiplication map

mD:(H0(𝒪(C+AD))T)jS1(H0(𝒪(C+AEjD))Tj)H0(𝒪Z)m_{D}\colon\left(H^{0}(\mathcal{O}(C+A-D))\otimes T\right)\oplus\bigoplus_{j\in S_{1}}\left(H^{0}(\mathcal{O}(C+A-E_{j}-D))\otimes T_{j}\right)\to H^{0}(\mathcal{O}_{Z})

defined by (ϕx,ϕjxj)ϕ|Zx+ϕj|Zxj(\phi\otimes x,\sum\phi_{j}\otimes x_{j})\mapsto\phi|_{Z}\cdot x+\sum\phi_{j}|_{Z}\cdot x_{j}. As |C+AD||C+A-D| is basepoint-free, there is a section ϕDH0(𝒪(C+AD))\phi_{D}\in H^{0}(\mathcal{O}(C+A-D)) that is nowhere zero on the support of ZZ, hence multiplying by ϕD\phi_{D} is injective, which implies that the image of H0(𝒪(C+AD))TH^{0}(\mathcal{O}(C+A-D))\otimes T has dimension dimT\geqslant\dim T. Similarly, as |C+AEiEj||C+A-E_{i}-E_{j}| is basepoint-free when pjpip_{j}\succ p_{i} (Example 2.6), we can choose ϕjH0(𝒪(C+AEiEj))\phi_{j}\in H^{0}(\mathcal{O}(C+A-E_{i}-E_{j})) that is nowhere zero on the support of ZZ, so the image of H0(𝒪(C+AEiEj))TjH^{0}(\mathcal{O}(C+A-E_{i}-E_{j}))\otimes T_{j} has dimension dimTj\geqslant\dim T_{j}.

Define the stratum

Wλ,{λj}jS1,{λi}iS0Hom(Ω,𝒪Z)W_{\lambda,\{\lambda_{j}\}_{j\in S_{1}},\{\lambda_{i}\}_{i\in S_{0}}}\subset{\rm Hom}(\Omega,\mathcal{O}_{Z})

by the conditions that dimT=λ\dim T=\ell-\lambda, dimTj=λj\dim T_{j}=\ell-\lambda_{j} for jS1j\in S_{1}, and the sum ϕEi|ZT+j:pjpiϕj|ZTj\phi_{E_{i}}|_{Z}\cdot T+\sum_{j\colon p_{j}\succ p_{i}}\phi_{j}|_{Z}\cdot T_{j} has dimension λi\ell-\lambda_{i} for iS0i\in S_{0}. By the above, the stratum is empty unless 0λiλ,λj0\leqslant\lambda_{i}\leqslant\lambda,\lambda_{j}\leqslant\ell when pjpip_{j}\succ p_{i}. For ψ\psi in this stratum, the image of mDm_{D} has dimension λ\geqslant\ell-\lambda and the image of mEim_{E_{i}} has dimension λi\geqslant\ell-\lambda_{i}. Thus, ext1(𝒪(D),kerψ)λ{\rm ext}^{1}(\mathcal{O}(D),\ker\psi)\leqslant\lambda and ext1(𝒪(Ei),kerψ)λi{\rm ext}^{1}(\mathcal{O}(E_{i}),\ker\psi)\leqslant\lambda_{i}, which yields the estimate

ext1(Λ,kerψ)λ(α2+α3+α4)+iS0λiγi=λ(α1+jS1γj+r+1)iS0(λλi)γi.{\rm ext}^{1}(\Lambda,\ker\psi)\leqslant\lambda(\alpha_{2}+\alpha_{3}+\alpha_{4})+\sum_{i\in S_{0}}\lambda_{i}\gamma_{i}=\lambda\Big{(}\alpha_{1}+\sum_{j\in S_{1}}\gamma_{j}+r+1\Big{)}-\sum_{i\in S_{0}}(\lambda-\lambda_{i})\gamma_{i}.

For strata with λ=0\lambda=0, then also λi=0\lambda_{i}=0, so ext1(Λ,kerψ)=0{\rm ext}^{1}(\Lambda,\ker\psi)=0 and the dimension of the space of commutative diagrams for ψ\psi in the union of such strata is equal to the expected dimension. Thus, it suffices to show that for strata with λ>0\lambda>0, the codimension of the stratum in Hom(Ω,𝒪Z){\rm Hom}(\Omega,\mathcal{O}_{Z}) is at least as large as ext1(Λ,kerψ){\rm ext}^{1}(\Lambda,\ker\psi).

When λ>0\lambda>0, we will compare this estimate for ext1(Λ,kerψ){\rm ext}^{1}(\Lambda,\ker\psi) to the codimension of the stratum in Hom(Ω,𝒪Z){\rm Hom}(\Omega,\mathcal{O}_{Z}). A map ψ\psi in this stratum can be obtained by first choosing a subspace TiH0(𝒪Z)T_{i}\subset H^{0}(\mathcal{O}_{Z}) of dimension λi\ell-\lambda_{i}, then choosing subspaces T(ϕEi|Z)1TiT\subset(\phi_{E_{i}}|_{Z})^{-1}\cdot T_{i} and Tj(ϕj|Z)1TiT_{j}\subset(\phi_{j}|_{Z})^{-1}\cdot T_{i} of dimension λ\ell-\lambda and λj\ell-\lambda_{j}, and finally using T,TjT,T_{j} to define the map ψ\psi. Comparing a dimension count based on this description to hom(Ω,𝒪Z)\hom(\Omega,\mathcal{O}_{Z}), we see that the stratum has codimension

λ(α1+λ)+jS1λj(γj+λj)+iS0λi(λiλ+j:pjpi(λj)).\lambda(\alpha_{1}+\lambda-\ell)+\sum_{j\in S_{1}}\lambda_{j}(\gamma_{j}+\lambda_{j}-\ell)+\sum_{i\in S_{0}}\lambda_{i}\left(\lambda_{i}-\lambda+\sum_{j\colon p_{j}\succ p_{i}}(\ell-\lambda_{j})\right).

The difference between the upper bound for ext1(Λ,kerψ){\rm ext}^{1}(\Lambda,\ker\psi) and this codimension is

Δ=λ(+r+1λ)+iS0j:pjpi(λjλi)(λj)jS1(λjλ)γjiS0(λλi)(γiλi).\Delta=\lambda(\ell+r+1-\lambda)+\sum_{i\in S_{0}}\sum_{j\colon p_{j}\succ p_{i}}(\lambda_{j}-\lambda_{i})(\ell-\lambda_{j})-\sum_{j\in S_{1}}(\lambda_{j}-\lambda)\gamma_{j}-\sum_{i\in S_{0}}(\lambda-\lambda_{i})(\gamma_{i}-\lambda_{i}).

Since γiM++j:pjpiγj\gamma_{i}\geqslant M+\ell+\sum_{j\colon p_{j}\succ p_{i}}\gamma_{j}, we obtain

Δλ(+r+1λ)iS0j:pjpi(λjλi)(γj+λj)iS0(λλi)(M+λi).\Delta\leqslant\lambda(\ell+r+1-\lambda)-\sum_{i\in S_{0}}\sum_{j\colon p_{j}\succ p_{i}}(\lambda_{j}-\lambda_{i})(\gamma_{j}+\lambda_{j}-\ell)-\sum_{i\in S_{0}}(\lambda-\lambda_{i})(M+\ell-\lambda_{i}).

As α2,α3,γiMλ(+r+1λ)\alpha_{2},\alpha_{3},\gamma_{i}\geqslant M\geqslant\lambda(\ell+r+1-\lambda), we may assume that the image of mDm_{D} for D=A,CD=A,C is exactly ϕD|ZT\phi_{D}|_{Z}\cdot T and that the image of mEim_{E_{i}} has dimension exactly λi\ell-\lambda_{i}, as otherwise we can improve the upper bound for ext1(Λ,kerψ){\rm ext}^{1}(\Lambda,\ker\psi) by subtracting MM, leading to a new Δ\Delta that is non-positive. Similarly, as M+λiM+\ell-\lambda_{i} and γj+λj\gamma_{j}+\lambda_{j}-\ell are each M\geqslant M, Δ\Delta is non-positive unless λi=λj=λ\lambda_{i}=\lambda_{j}=\lambda. Thus, all that remains is the case when ϕEi|ZT=ϕj|ZTj\phi_{E_{i}}|_{Z}\cdot T=\phi_{j}|_{Z}\cdot T_{j} for all i,ji,j such that pjpip_{j}\succ p_{i} and the image of each mDm_{D} has dimension λ\ell-\lambda. In other words, up to multiplication by units, TT is stable under multiplication by rational functions ζ\zeta in H0(𝒪(C+AD))H^{0}(\mathcal{O}(C+A-D)) for D=A,C,EiD=A,C,E_{i} and H0(𝒪(C+AEiEj)H^{0}(\mathcal{O}(C+A-E_{i}-E_{j}) for pjpip_{j}\succ p_{i}.

This final case cannot happen. Since 1Z1_{Z} is in the image of Hom(𝒪,Ω)Hom(𝒪,𝒪Z){\rm Hom}(\mathcal{O},\Omega)\to{\rm Hom}(\mathcal{O},\mathcal{O}_{Z}), TT must contain an element α\alpha that restricts to a unit at each point of the support of ZZ. But the linear systems |A||A|, |C||C|, |C+AEi||C+A-E_{i}| for iS0i\in S_{0}, and |C+AEiEj||C+A-E_{i}-E_{j}| for pjpip_{j}\succ p_{i} collectively separate points and tangents on XX (compare with the proof of Proposition 2.8), hence multiplying α\alpha by the functions ζ\zeta generates all of H0(𝒪Z)H^{0}(\mathcal{O}_{Z}), so the only way TT can be stable under multiplication by all ζ\zeta is if T=H0(𝒪Z)T=H^{0}(\mathcal{O}_{Z}). This cannot happen as λ>0\lambda>0.

We finish the proof by arguing that there is a unique component with the maximal dimension. As shown above, over a general point (ψ,π)(Ω[])×α41(\psi,\pi)\in\mathbb{P}\left(\Omega^{*[\ell]}\right)\times\mathbb{P}^{\alpha_{4}-1}, the fiber of Ξ\Xi is an affine space in \mathbb{P} isomorphic to Hom(Λ,kerψ){\rm Hom}(\Lambda,\ker\psi). The fiber of any component with maximal dimension must contain a non-empty open set in this affine space for dimension reasons, but since any two non-empty open sets in an affine space must intersect, there can only be one such component. ∎

Appendix: Positivity Conditions

We rephrase the conditions on the exponents of Gaeta resolutions in Proposition 4.5(a) to explain the connection with the positivity of the first Chern class. We then summarize the inequalities required for the proof of Theorems 1.6 and 1.7.

Given the numerical class

f=(r,αA+δCiS0γiEijS1γjEj,χ)f=\Big{(}r,\alpha A+\delta C-\sum_{i\in S_{0}}\gamma_{i}E_{i}-\sum_{j\in S_{1}}\gamma_{j}E_{j},\chi\Big{)}

on XX an admissible blowup of 𝔽e\mathbb{F}_{e}, the Euler characteristics in Proposition 4.5(a) can be written more explicitly as

α1\displaystyle\alpha_{1} =δ+α+rχiS0γijS1γj;\displaystyle=\delta+\alpha+r-\chi-\sum_{i\in S_{0}}\gamma_{i}-\sum_{j\in S_{1}}\gamma_{j};
α2\displaystyle\alpha_{2} =α+rχiS0γi;\displaystyle=\alpha+r-\chi-\sum_{i\in S_{0}}\gamma_{i};
α3\displaystyle\alpha_{3} =δ+rχiS0γi.\displaystyle=\delta+r-\chi-\sum_{i\in S_{0}}\gamma_{i}.

Thus, rephrasing Proposition 4.5(a) in the case when χr\chi\geqslant r and γij:pjpiγj\gamma_{i}\geqslant\sum_{j\colon p_{j}\succ p_{i}}\gamma_{j} for all ii, we can say that ff admits Gaeta resolutions if and only if γi,γj,χ\gamma_{i},\gamma_{j},\chi are all 0\geqslant 0 and

α,δiS0γi+χr.\alpha,\delta\geqslant\sum_{i\in S_{0}}\gamma_{i}+\chi-r. (A.1)

Now, as in § 7, consider the orthogonal classes ρ=(1,0,1)\rho=(1,0,1-\ell) and

σ=(r,L=αA+δCiS0γiEijS1γjEj,χ=r),\sigma=\Big{(}r,L=\alpha A+\delta C-\sum_{i\in S_{0}}\gamma_{i}E_{i}-\sum_{j\in S_{1}}\gamma_{j}E_{j},\chi=r\ell\Big{)},

where 1\ell\geqslant 1 and r2r\geqslant 2 are fixed, and set v=σ+ρv=\sigma+\rho. The assumption that LL is sufficiently positive in § 7 includes three conditions. Let

M=max{m(+r+1m)1m}M=\max\{\,m(\ell+r+1-m)\mid 1\leqslant m\leqslant\ell\,\}

The first set of conditions used in the proof of Theorem 1.6 is:

γjMfor all jS1;γiM+j:pjpiγjfor all iS0;α,δM+iS0γi+r(1).\displaystyle\begin{split}\gamma_{j}&\geqslant M\qquad\qquad\quad\quad\>\>\text{for all $j\in S_{1}$};\\ \gamma_{i}&\geqslant M+\sum_{j\colon p_{j}\succ p_{i}}\gamma_{j}\qquad\>\text{for all $i\in S_{0}$};\\ \alpha,\delta&\geqslant M+\sum_{i\in S_{0}}\gamma_{i}+r(\ell-1).\end{split} (A.2)

These conditions ensure that σ\sigma and vv admit Gaeta resolutions (Proposition 4.5(a)), general Gaeta resolutions for σ\sigma and vv are locally free (Proposition 5.1), general Gaeta resolutions for vv have no sections vanishing on curves (Proposition 5.8(b)), Weak Brill-Noether holds (Proposition 5.7), and dimΞ=dim\dim\Xi=\dim\mathbb{P} (Proposition 8.1).

The inequalities (A.2) imply that LL is MM-very ample, as LL can be expressed as a tensor product of MM very ample line bundles by Proposition 2.8 and mm-very ampleness is additive under tensor products ([HTT05]).

The second condition is the discriminant condition (6.9), which in this context is

P(1rL)(λ+1)24λ+t8+1r+,where λ=uv+e2.P\left(\frac{1}{r}L\right)\geqslant\frac{(\lambda+1)^{2}}{4\lambda}+\frac{t}{8}+\frac{1}{r}+\ell,\qquad\text{where $\lambda=\frac{u}{v}+\frac{e}{2}$}. (A.3)

The last condition is that

Θσ=Lr2E has vanishing higher cohomology,\text{$\Theta_{\sigma}=L_{\ell}-\frac{r}{2}E$ has vanishing higher cohomology}, (A.4)

which by Proposition 2.8 and Lemma 7.2 can be ensured by the following inequalities:

γjr(1)for all jS1;γir(1)+j:pjpi(γj+1)for all iS0;δr(1)1+iS0(γi+1);αr(1)1+e+iS0(γi+1).\displaystyle\begin{split}\gamma_{j}&\geqslant r(\ell-1)\hskip 113.81102pt\text{for all $j\in S_{1}$};\\ \gamma_{i}&\geqslant r(\ell-1)+\sum_{j\colon p_{j}\succ p_{i}}(\gamma_{j}+1)\hskip 37.55785pt\text{for all $i\in S_{0}$};\\ \delta&\geqslant r(\ell-1)-1+\sum_{i\in S_{0}}(\gamma_{i}+1);\\ \alpha&\geqslant r(\ell-1)-1+e+\sum_{i\in S_{0}}(\gamma_{i}+1).\end{split} (A.5)

The sufficiency of these conditions follows from computing LKXL-K_{X} and observing that it can be decomposed as the tensor product of an \ell-very ample line bundle and r1r-1 (1)(\ell-1)-very ample line bundles. For e,s,te,s,t not too large relative to \ell and rr, (A.5) is implied by (A.2).

References

  • [Abe10] Takeshi Abe. Deformation of rank 2 quasi-bundles and some strange dualities for rational surfaces. Duke Math. J., 155(3):577–620, 2010.
  • [Bae88] Dagmar Baer. Tilting sheaves in representation theory of algebras. Manuscripta Math., 60(3):323–347, 1988.
  • [Bal87] Edoardo Ballico. On moduli of vector bundles on rational surfaces. Arch. Math. (Basel), 49(3):267–272, 1987.
  • [BGJ16] Aaron Bertram, Thomas Goller, and Drew Johnson. Le potier’s strange duality, quot schemes, and multiple point formulas for del pezzo surfaces. arXiv preprint arXiv:1610.04185, 2016.
  • [BMOY17] Barbara Bolognese, Alina Marian, Dragos Oprea, and Kota Yoshioka. On the strange duality conjecture for abelian surfaces II. J. Algebraic Geom., 26(3):475–511, 2017.
  • [Bon89] A. I. Bondal. Representations of associative algebras and coherent sheaves. Izv. Akad. Nauk SSSR Ser. Mat., 53(1):25–44, 1989.
  • [Bot57] Raoul Bott. Homogeneous vector bundles. Ann. of Math. (2), 66:203–248, 1957.
  • [BS91] M. Beltrametti and A. J. Sommese. Zero cycles and kkth order embeddings of smooth projective surfaces. In Problems in the theory of surfaces and their classification (Cortona, 1988), Sympos. Math., XXXII, pages 33–48. Academic Press, London, 1991. With an appendix by Lothar Göttsche.
  • [CG90] Fabrizio Catanese and Lothar Gœttsche. dd-very-ample line bundles and embeddings of Hilbert schemes of 0-cycles. Manuscripta Math., 68(3):337–341, 1990.
  • [CH18] Izzet Coskun and Jack Huizenga. Weak Brill-Noether for rational surfaces. In Local and global methods in algebraic geometry, volume 712 of Contemp. Math., pages 81–104. Amer. Math. Soc., [Providence], RI, [2018] ©2018.
  • [CH20] Izzet Coskun and Jack Huizenga. Brill-Noether theorems and globally generated vector bundles on Hirzebruch surfaces. Nagoya Math. J., 238:1–36, 2020.
  • [CH21] Izzet Coskun and Jack Huizenga. Existence of semistable sheaves on Hirzebruch surfaces. Adv. Math., 381:Paper No. 107636, 96, 2021.
  • [Dan02] Gentiana Danila. Résultats sur la conjecture de dualité étrange sur le plan projectif. Bull. Soc. Math. France, 130(1):1–33, 2002.
  • [DLP85] J.-M. Drezet and J. Le Potier. Fibrés stables et fibrés exceptionnels sur 𝐏2{\bf P}_{2}. Ann. Sci. École Norm. Sup. (4), 18(2):193–243, 1985.
  • [Dre86] J.-M. Drezet. Fibrés exceptionnels et suite spectrale de Beilinson généralisée sur 𝐏2(𝐂){\bf P}_{2}({\bf C}). Math. Ann., 275(1):25–48, 1986.
  • [Fog73] J. Fogarty. Algebraic families on an algebraic surface. II. The Picard scheme of the punctual Hilbert scheme. Amer. J. Math., 95:660–687, 1973.
  • [GH78] Phillip Griffiths and Joseph Harris. Principles of algebraic geometry. Pure and Applied Mathematics. Wiley-Interscience [John Wiley & Sons], New York, 1978.
  • [GK04] A. L. Gorodentsev and S. A. Kuleshov. Helix theory. Mosc. Math. J., 4(2):377–440, 535, 2004.
  • [GL22] Thomas Goller and Yinbang Lin. Rank-one sheaves and stable pairs on surfaces. Adv. Math., 401:Paper No. 108322, 32, 2022.
  • [GM22] Lothar Göttsche and Anton Mellit. Refined Verlinde and Segre formula for Hilbert schemes. arXiv e-prints, page arXiv:2210.01059, October 2022.
  • [Gro61] Alexander Grothendieck. Techniques de construction et théorèmes d’existence en géométrie algébrique. IV. Les schémas de Hilbert. In Séminaire Bourbaki, volume 6 Exp. No. 221, pages 249–276. Soc. Math. France, Paris, 1960-1961.
  • [Har77] Robin Hartshorne. Algebraic geometry. Springer-Verlag, New York-Heidelberg, 1977. Graduate Texts in Mathematics, No. 52.
  • [HL10] Daniel Huybrechts and Manfred Lehn. The geometry of moduli spaces of sheaves. Cambridge Mathematical Library. Cambridge University Press, Cambridge, second edition, 2010.
  • [HP11] Lutz Hille and Markus Perling. Exceptional sequences of invertible sheaves on rational surfaces. Compos. Math., 147(4):1230–1280, 2011.
  • [HTT05] Yukitoshi Hinohara, Kazuyoshi Takahashi, and Hiroyuki Terakawa. On tensor products of kk-very ample line bundles. Proc. Amer. Math. Soc., 133(3):687–692, 2005.
  • [Hui16] Jack Huizenga. Effective divisors on the Hilbert scheme of points in the plane and interpolation for stable bundles. J. Algebraic Geom., 25(1):19–75, 2016.
  • [Huy06] D. Huybrechts. Fourier-Mukai transforms in algebraic geometry. Oxford Mathematical Monographs. The Clarendon Press, Oxford University Press, Oxford, 2006.
  • [Joh18] Drew Johnson. Universal Series for Hilbert Schemes and Strange Duality. International Mathematics Research Notices, 05 2018.
  • [Lin18] Yinbang Lin. Moduli spaces of stable pairs. Pacific J. Math., 294(1):123–158, 2018.
  • [LP92] J. Le Potier. Fibré déterminant et courbes de saut sur les surfaces algébriques. In Complex projective geometry (Trieste, 1989/Bergen, 1989), volume 179 of London Math. Soc. Lecture Note Ser., pages 213–240. Cambridge Univ. Press, Cambridge, 1992.
  • [LP97] J. Le Potier. Lectures on vector bundles, volume 54 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 1997. Translated by A. Maciocia.
  • [LP98] J. Le Potier. Problème de Brill-Noether et groupe de Picard de l’espace de modules des faisceaux semi-stables sur le plan projectif. In Algebraic geometry (Catania, 1993/Barcelona, 1994), volume 200 of Lecture Notes in Pure and Appl. Math., pages 75–90. Dekker, New York, 1998.
  • [LP05] J. Le Potier. Dualité étrange sur le plan projectif. Unpublished, 2005.
  • [MFK94] D. Mumford, J. Fogarty, and F. Kirwan. Geometric invariant theory, volume 34 of Ergebnisse der Mathematik und ihrer Grenzgebiete (2) [Results in Mathematics and Related Areas (2)]. Springer-Verlag, Berlin, third edition, 1994.
  • [MO07a] Alina Marian and Dragos Oprea. The level-rank duality for non-abelian theta functions. Invent. Math., 168(2):225–247, 2007.
  • [MO07b] Alina Marian and Dragos Oprea. Virtual intersections on the Quot scheme and Vafa-Intriligator formulas. Duke Math. J., 136(1):81–113, 2007.
  • [MOP19] Alina Marian, Dragos Oprea, and Rahul Pandharipande. The combinatorics of Lehn’s conjecture. J. Math. Soc. Japan, 71(1):299–308, 2019.
  • [MOP22] Alina Marian, Dragos Oprea, and Rahul Pandharipande. Higher rank Segre integrals over the Hilbert scheme of points. J. Eur. Math. Soc. (JEMS), 24(8):2979–3015, 2022.
  • [Orl92] D. O. Orlov. Projective bundles, monoidal transformations, and derived categories of coherent sheaves. Izv. Ross. Akad. Nauk Ser. Mat., 56(4):852–862, 1992.
  • [Ott95] G. Ottaviani. Varietà proiettive di codimensione piccola. Ist. nazion. di alta matematica F. Severi. Aracne, 1995.
  • [Ped21] Dmitrii Pedchenko. The Picard Group of the Moduli Space of Sheaves on a Quadric Surface. International Mathematics Research Notices, 09 2021. rnab175.
  • [Wal98] Charles Walter. Irreducibility of moduli spaces of vector bundles on birationally ruled surfaces. In Algebraic geometry (Catania, 1993/Barcelona, 1994), volume 200 of Lecture Notes in Pure and Appl. Math., pages 201–211. Dekker, New York, 1998.
  • [Yos96] Kōta Yoshioka. A note on a paper of J.-M. Drézet on the local factoriality of some moduli spaces: “Points non factoriels des variétés de modules de faisceaux semi-stables sur une surface rationnelle” [J. Reine Angew. Math. 413 (1991), 99–126; MR1089799 (92d:14009)]. Internat. J. Math., 7(6):843–858, 1996.
  • [Yua21] Yao Yuan. Strange duality on 2\mathbb{P}^{2} via quiver representations. Adv. Math., 377:Paper No. 107469, 35, 2021.