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

Higher rank Brill-Noether theory on 2\mathbb{P}^{2}

Ben Gould1,∗, Yeqin Liu2, and Woohyung Lee3
Abstract.

Let M2(𝐯)M_{\mathbb{P}^{2}}(\mathbf{v}) be a moduli space of semistable sheaves on 2\mathbb{P}^{2}, and let Bk(𝐯)M2(𝐯)B^{k}(\mathbf{v})\subseteq M_{\mathbb{P}^{2}}(\mathbf{v}) be the Brill-Noether locus of sheaves EE with h0(2,E)kh^{0}(\mathbb{P}^{2},E)\geq k. In this paper we develop the foundational properties of Brill-Noether loci on 2\mathbb{P}^{2}. Set r=r(E)r=r(E) to be the rank and c1,c2c_{1},c_{2} the Chern classes. The Brill-Noether loci have natural determinantal scheme structures and expected dimensions dimBk(𝐯)=dimM2(𝐯)k(kχ(E))\dim B^{k}(\mathbf{v})=\dim M_{\mathbb{P}^{2}}(\mathbf{v})-k(k-\chi(E)). When c1>0c_{1}>0, we show that the Brill-Noether locus Br(𝐯)B^{r}(\mathbf{v}) is nonempty. When c1=1c_{1}=1, we show all of the Brill-Noether loci are irreducible and of the expected dimension. We show that when μ=c1/r>1/2\mu=c_{1}/r>1/2 is not an integer and c20c_{2}\gg 0, the Brill-Noether loci are reducible and describe distinct irreducible components of both expected and unexpected dimension.

1,2,3: Department of Mathematics, Statistics, and Computer Science, University of Illinois at Chicago, Science and Engineering Offices, 851 South Morgan Street, Chicago, IL 60607, USA

Correspondence to be sent to: email: [email protected]

1. Introduction

On a polarized variety (X,H)(X,H) the moduli spaces MX,H(𝐯)M_{X,H}(\mathbf{v}) of HH-semistable sheaves of numerical type 𝐯\mathbf{v} carry Brill-Noether loci Bk(𝐯)MX,H(𝐯)B^{k}(\mathbf{v})\subseteq M_{X,H}(\mathbf{v}) whose members EE satisfy h0(X,E)kh^{0}(X,E)\geq k. In this paper we develop the foundational properties of Brill-Noether loci on X=2X=\mathbb{P}^{2}. Set r=ch0(𝐯)r=\text{ch}_{0}(\mathbf{v}), μ=μ(𝐯)=ch1(𝐯)ch0(𝐯)\mu=\mu(\mathbf{v})=\frac{\text{ch}_{1}(\mathbf{v})}{\text{ch}_{0}(\mathbf{v})} and Δ=Δ(𝐯)=12μ(𝐯)2ch2(𝐯)ch0(𝐯)\Delta=\Delta(\mathbf{v})=\frac{1}{2}\mu(\mathbf{v})^{2}-\frac{\text{ch}_{2}(\mathbf{v})}{\text{ch}_{0}(\mathbf{v})} to be the rank, slope, and discriminant of 𝐯\mathbf{v}. The Brill-Noether loci have natural determinantal scheme structures and expected dimensions expdimBk(𝐯)=dimM2(𝐯)k(kχ(E))\text{expdim}B^{k}(\mathbf{v})=\dim M_{\mathbb{P}^{2}}(\mathbf{v})-k(k-\chi(E)). When μ>0\mu>0 we show that Br(𝐯)B^{r}(\mathbf{v}) is nonempty. When ch1(𝐯)=1\text{ch}_{1}(\mathbf{v})=1, we show each Brill-Noether locus Bk(𝐯)B^{k}(\mathbf{v}) is irreducible and of the expected dimension as a determinantal variety, dimBk(𝐯)=dimM2(𝐯)k(kχ(𝐯))\dim B^{k}(\mathbf{v})=\dim M_{\mathbb{P}^{2}}(\mathbf{v})-k(k-\chi(\mathbf{v})). We prove that when r,ch1>0r,\text{ch}_{1}>0 and μ>1/2\mu>1/2 is not an integer with Δ0\Delta\gg 0, then Br(𝐯)B^{r}(\mathbf{v}) is reducible, and describe distinct irreducible components.

When X=CX=C is a smooth curve and the rank is 1, MX(𝐯)=Picd(C)M_{X}(\mathbf{v})=\text{Pic}^{d}(C) is the space of line bundles of a given degree dd, and the Brill-Noether loci Bk(𝐯)=Wdk(C)B^{k}(\mathbf{v})=W^{k}_{d}(C) have been studied for over a century [ACGH85]. Much is known about their geometry when CC is general: when the expected dimension of Wdk(C)W^{k}_{d}(C) as a determinantal variety is positive, it is nonempty of that dimension and irreducible. A general LPicd(C)L\in\text{Pic}^{d}(C) has h0(L)=χ(L)=dg+1h^{0}(L)=\chi(L)=d-g+1 when this is non-negative. In higher rank the general bundle still has h0(E)=χ(E)h^{0}(E)=\chi(E) when this is non-negative, and the Brill-Noether loci have been studied in detail (for a survey see, e.g., [New21] and its bibliography).

On algebraic surfaces, however, much less is known about Brill-Noether theory. The basic theory has begun to be worked out on Hirzebruch surfaces (see [CMR10]) and in rank 2 on 2\mathbb{P}^{2} (see [RLTLZ21]). On 2\mathbb{P}^{2}, the moduli spaces M2(𝐯)M_{\mathbb{P}^{2}}(\mathbf{v}) of semistable bundles of any rank are irreducible, and a well-known theorem of Göttsche-Hirschowitz describes the global sections of a general bundle EM2(𝐯)E\in M_{\mathbb{P}^{2}}(\mathbf{v}) of rank at least two: if the slope μ(E)\mu(E) is positive, then h0(E)=max{0,χ(E)}h^{0}(E)=\max\{0,\chi(E)\}, which is determined by 𝐯\mathbf{v}. More generally, a general bundle EE of any slope has at most one nonzero cohomology group, and by semicontinuity there is an open dense subset of M2(𝐯)M_{\mathbb{P}^{2}}(\mathbf{v}) of bundles with this cohomology. The Brill-Noether loci Bk(𝐯)B^{k}(\mathbf{v}) with k>χ(𝐯)k>\chi(\mathbf{v}) form its complement. These foundational results make the study of the Brill-Noether loci Bk(𝐯)B^{k}(\mathbf{v}) on 2\mathbb{P}^{2} approachable.

1.1. Geometry of Brill-Noether loci

The moduli spaces M(𝐯)=M2(𝐯)M(\mathbf{v})=M_{\mathbb{P}^{2}}(\mathbf{v}) of semistable sheaves on 2\mathbb{P}^{2} with Chern character 𝐯\mathbf{v} are irreducible projective algebraic varieties. We call a Chern character 𝐯K(2)\mathbf{v}\in K(\mathbb{P}^{2}) stable if it is the Chern character of a stable sheaf, so that M(𝐯)M(\mathbf{v})\neq\emptyset. We are interested in the Brill-Noether loci Bk(𝐯)M(𝐯)B^{k}(\mathbf{v})\subseteq M(\mathbf{v}). As in the case of line bundles on curves, the Brill-Noether loci are constructed as determinantal varieties, so each has an expected dimension, which is a lower bound for the dimension of an irreducible component of Bk(𝐯)B^{k}(\mathbf{v}); it is

expdimBk(𝐯)=dimM2(𝐯)k(kχ(𝐯)).\text{expdim}B^{k}(\mathbf{v})=\dim M_{\mathbb{P}^{2}}(\mathbf{v})-k(k-\chi(\mathbf{v})).

Our main results are summarized as follows.

Theorem 1.1.

Let 𝐯K(2)\mathbf{v}\in K(\mathbb{P}^{2}) be a stable Chern character and M(𝐯)M(\mathbf{v}) the associated moduli space of semistable sheaves. Suppose that μ(𝐯)>0\mu(\mathbf{v})>0 and set r=ch0(𝐯)r=\text{ch}_{0}(\mathbf{v}).

  1. (1)

    (Emptiness) For any EM(𝐯)E\in M(\mathbf{v}),

    (1) h0(E)max{12c1(E)2+32c1(E)+1,r(E)}.h^{0}(E)\leq\max\left\{\frac{1}{2}c_{1}(E)^{2}+\frac{3}{2}c_{1}(E)+1,r(E)\right\}.

    Equivalently, if kmax{12c1(E)2+32c1(E)+1,r(E)}k\geq\max\left\{\frac{1}{2}c_{1}(E)^{2}+\frac{3}{2}c_{1}(E)+1,r(E)\right\}, then Bk(𝐯)B^{k}(\mathbf{v}) is empty.

    When r(E)12c1(E)2+32c1(E)+1r(E)\geq\frac{1}{2}c_{1}(E)^{2}+\frac{3}{2}c_{1}(E)+1, this bound is sharp, and Br(𝐯)B^{r}(\mathbf{v}) contains a component of the expected dimension. See Theorem 4.1 and Theorem 4.6.

  2. (2)

    (Nonemptiness) Br(𝐯)B^{r}(\mathbf{v}) is nonempty. See Theorem 4.7.

  3. (3)

    (Irreducibility) If ch1(𝐯)=1\text{ch}_{1}(\mathbf{v})=1, all of the nonempty Brill-Noether loci on M(𝐯)M(\mathbf{v}) are irreducible and of the expected dimension. See Theorem 5.1.

  4. (4)

    (Reducibility) Suppose ch1(𝐯)>1\text{ch}_{1}(\mathbf{v})>1. If μ(𝐯)>1/2\mu(\mathbf{v})>1/2 is not an integer and Δ(𝐯)0\Delta(\mathbf{v})\gg 0, then Br(𝐯)B^{r}(\mathbf{v}) is reducible and contains components of both expected and unexpected dimensions. See Theorem 5.8.

Since the Brill-Noether loci are nested, Theorem 1.1 (2) implies the Brill-Noether loci Bk(𝐯)B^{k}(\mathbf{v}) are nonempty for all 0kr0\leq k\leq r. The bound (1) is also sharp for the line bundles 𝒪(d)\mathcal{O}(d) and the Lazarsfeld-Mukai-type bundles Md=coker(𝒪(d)𝒪h0(𝒪(d)))M_{d}=\text{coker}(\mathcal{O}(-d)\rightarrow\mathcal{O}^{h^{0}(\mathcal{O}(d))}), including M1=T2(1)M_{1}=T_{\mathbb{P}^{2}}(-1).

Our main technical tools for studying Brill-Noether loci are parametrizations of certain moduli spaces M(𝐯)M(\mathbf{v}) and Brill-Noether loci Bk(𝐯)B^{k}(\mathbf{v}) by projective bundles whose fibers are (projectivizations of) extension spaces. The applications to Brill-Noether loci are often of the following form. A sheaf EE on 2\mathbb{P}^{2} has an evaluation map on global sections

evE:H0(2,E)𝒪2E.\text{ev}_{E}:H^{0}(\mathbb{P}^{2},E)\otimes\mathcal{O}_{\mathbb{P}^{2}}\rightarrow E.

Assume that the general EM(𝐯)E\in M(\mathbf{v}) has mm global sections, i.e., h0(2,E)=mh^{0}(\mathbb{P}^{2},E)=m, and mr(E)m\leq r(E). When evE\text{ev}_{E} has full rank EE sits as an extension class

[0H0(2,E)𝒪2evEEE0]Ext1(E,H0(2,E)𝒪2).[0\rightarrow H^{0}(\mathbb{P}^{2},E)\otimes\mathcal{O}_{\mathbb{P}^{2}}\stackrel{{\scriptstyle\text{ev}_{E}}}{{\rightarrow}}E\rightarrow E^{\prime}\rightarrow 0]\in\text{Ext}^{1}(E^{\prime},H^{0}(\mathbb{P}^{2},E^{\prime})\otimes\mathcal{O}_{\mathbb{P}^{2}}).

If EE^{\prime} is semistable with Chern character 𝐯K(2)\mathbf{v}^{\prime}\in K(\mathbb{P}^{2}), we may form the projective bundle \mathbb{P} over M(𝐯)M(\mathbf{v}^{\prime}) whose fiber over EE^{\prime} is the projective space Ext1(E,H0(2,E)𝒪2)\mathbb{P}\text{Ext}^{1}(E^{\prime},H^{0}(\mathbb{P}^{2},E^{\prime})\otimes\mathcal{O}_{\mathbb{P}^{2}}). If the general such extension is stable, we obtain a classifying rational map ϕ:M(𝐯)\phi:\mathbb{P}\dashrightarrow M(\mathbf{v}), which we call the extension parametrization of M(𝐯)M(\mathbf{v}).

The critical technical challenge in constructing such parametrizations is in proving the stability of extension sheaves. When the rank of EE^{\prime} is 0, so χ(E)=r(E)\chi(E)=r(E), we accomplish this by studying smooth curves in 2\mathbb{P}^{2} realized as the support of EE^{\prime} and the moduli spaces Pic𝒞/Ud\text{Pic}^{d}_{\mathcal{C}/U} of these sheaves (Sections 4 and 5). The central inputs to our results are the classification of stable Chern characters on 2\mathbb{P}^{2} by Drézet-Le Potier ([DLP85, Théorème C]) and the computation of the cohomology of the general stable bundle in M(𝐯)M(\mathbf{v}) by Göttsche-Hirschowitz (Theorem 2.1). Our methods likely extend to other surfaces for which these two notions are understood, e.g., Hirzebruch surfaces and K3 surfaces of Picard rank 1 (see [CH21] and [CNY21] respectively).

1.2. Organization of the paper

In Section 2 we recall basic facts about moduli spaces of sheaves on 2\mathbb{P}^{2}. In Section 3 we introduce extension parametrizations, which are the main tools used in further sections.

In Section 4 we prove emptiness and non-emptiness statements for Brill-Noether loci. In Section 5 we prove irreducibility and reducibility statements for Brill-Noether loci.

2. Preliminaries

2.1. Stability on 2\mathbb{P}^{2}

In this section we collect basic facts about stability for coherent sheaves on 2\mathbb{P}^{2}. We refer the reader to the books by Le Potier [LP97] and Huybrechts-Lehn [HL10] for further details.

2.1.1. Numerical invariants & stability

All sheaves in this paper will be coherent, but not necessarily torsion-free. Let EE be a coherent sheaf on 2\mathbb{P}^{2}. The Hilbert polynomial PE(m)=χ(E(m))P_{E}(m)=\chi(E(m)) is of the form

PE(m)=αdmdd!+O(md1),P_{E}(m)=\alpha_{d}\frac{m^{d}}{d!}+O(m^{d-1}),

and we define the reduced Hilbert polynomial pE(m)p_{E}(m) to be

pE(m)=PE(m)/αd.p_{E}(m)=P_{E}(m)/\alpha_{d}.

A pure-dimensional coherent sheaf EE on 2\mathbb{P}^{2} is (semi)stable if for all nontrivial subsheaves FEF\hookrightarrow E we have pF<()pEp_{F}<(\leq)p_{E}, where polynomials are compared for sufficiently large mm. We will assume throughout this paper that a semistable sheaf has positive rank.

Given a character 𝐯K(2)\mathbf{v}\in K(\mathbb{P}^{2}), we define, respectively, the slope and the discriminant of 𝐯\mathbf{v} by

μ(𝐯)=ch1(𝐯)ch0(𝐯),Δ(𝐯)=12μ(𝐯)2ch2(𝐯)ch0(𝐯).\mu(\mathbf{v})=\frac{\text{ch}_{1}(\mathbf{v})}{\text{ch}_{0}(\mathbf{v})},\quad\Delta(\mathbf{v})=\frac{1}{2}\mu(\mathbf{v})^{2}-\frac{\text{ch}_{2}(\mathbf{v})}{\text{ch}_{0}(\mathbf{v})}.

On 2\mathbb{P}^{2} these classes may be considered as rational numbers. When 𝐯=ch(E)\mathbf{v}=\text{ch}(E) for a coherent sheaf EE we set μ(E)=μ(𝐯)\mu(E)=\mu(\mathbf{v}) and Δ(E)=Δ(𝐯)\Delta(E)=\Delta(\mathbf{v}). We have

μ(EF)=μ(E)+μ(F),Δ(EF)=Δ(E)+Δ(F)\mu(E\otimes F)=\mu(E)+\mu(F),\quad\Delta(E\otimes F)=\Delta(E)+\Delta(F)

for coherent sheaves E,FE,F on 2\mathbb{P}^{2}.

In terms of these invariants, the Riemann-Roch theorem takes the form

χ(E)=r(E)(p(μ(E))Δ(E)),\chi(E)=r(E)(p(\mu(E))-\Delta(E)),

where

p(x)=p𝒪2(x)=12(x2+3x+2).p(x)=p_{\mathcal{O}_{\mathbb{P}^{2}}}(x)=\frac{1}{2}(x^{2}+3x+2).

More generally for sheaves E,FE,F on 2\mathbb{P}^{2}, we set exti(E,F)=dimExti(E,F)\text{ext}^{i}(E,F)=\dim\text{Ext}^{i}(E,F). The relative Riemann-Roch theorem states

χ(E,F)=(1)iexti(E,F)=r(E)r(F)(p(μ(E)μ(F))Δ(E)Δ(F)),\chi(E,F)=\sum(-1)^{i}\text{ext}^{i}(E,F)=r(E)r(F)(p(\mu(E)-\mu(F))-\Delta(E)-\Delta(F)),

and similarly for Chern characters.

Additionally, we may define slope-stability for sheaves on 2\mathbb{P}^{2}: EE is slope-(semi)stable if for all subsheaves FEF\hookrightarrow E of smaller rank, we have μ(F)<()μ(E)\mu(F)<(\leq)\mu(E). Slope-stability implies stability, and is often easier to check.

2.1.2. Classification of stable bundles

The Chern characters of stable bundles on 2\mathbb{P}^{2} have been classified (see, e.g., [DLP85], [LP97]). The classification is important for our results, and we sketch it in this subsection.

Recall that we say a character 𝐯K(2)\mathbf{v}\in K(\mathbb{P}^{2}) is stable if there is a stable sheaf of Chern character 𝐯\mathbf{v}, and that we assume stable characters have ch0(𝐯)>0\text{ch}_{0}(\mathbf{v})>0. Stability imposes strong conditions on the numerical data of 𝐯\mathbf{v}. These conditions are determined by exceptional bundles on 2\mathbb{P}^{2}. A sheaf EE on 2\mathbb{P}^{2} is exceptional if Exti(E,E)=0\text{Ext}^{i}(E,E)=0 for i>0i>0.

By a theorem of Drézet [Dré86], every exceptional sheaf on 2\mathbb{P}^{2} is a stable vector bundle. A stable bundle EE is exceptional if and only if Δ(E)<1/2\Delta(E)<1/2, by Riemann-Roch [LP97, Proposition 16.1.1]. By definition, exceptional bundles are rigid and their moduli spaces consist of a single reduced point [LP97, Corollary 16.1.5]. Exceptional bundles on 2\mathbb{P}^{2} include the line bundles 𝒪2(k)\mathcal{O}_{\mathbb{P}^{2}}(k) and the tangent bundle T2T_{\mathbb{P}^{2}}, and all others can be formed from these by a process called mutation [Dré86].

The slopes of exceptional bundles can be completely described [LP97, §16.3]; set \mathcal{E} to be the set of slopes of exceptional bundles. For an exceptional bundle EE of slope α\alpha we write E=EαE=E_{\alpha} and Δ(E)=Δα\Delta(E)=\Delta_{\alpha}. We define the Drézet-Le Potier curve to be the following piecewise-polynomial curve in the (μ,Δ)(\mu,\Delta)-plane:

δ(μ)=supα:|μα|<3(p(|μα|)Δα).\delta(\mu)=\sup_{\alpha\in\mathcal{E}:|\mu-\alpha|<3}(p(-|\mu-\alpha|)-\Delta_{\alpha}).

A plot of an approximation of the Drézet-Le Potier curve between 0 and 1 is shown in Figure 1. The precise relationship of exceptional bundles to stable characters is due to Drézet and Le Potier ([DLP85, Théorème C], [LP97]): a character 𝐯K(2)\mathbf{v}\in K(\mathbb{P}^{2}) is stable if and only if c1:=rμc_{1}:=r\mu\in\mathbb{Z}, χ:=r(P(μ)Δ)\chi:=r(P(\mu)-\Delta)\in\mathbb{Z}, and Δ(𝐯)δ(μ(𝐯))\Delta(\mathbf{v})\geq\delta(\mu(\mathbf{v})), or 𝐯\mathbf{v} is exceptional. A character satisfying the first two conditions in the theorem is called integral. Each exceptional bundle EαE_{\alpha} of slope α\alpha determines two “branches” of the Drézet-Le Potier curve, on the left and right sides of the vertical line μ=α\mu=\alpha. The characters 𝐯\mathbf{v} on the branch to the right of μ=α\mu=\alpha satisfy χ(𝐯,Eα)=0\chi(\mathbf{v},E_{\alpha})=0, and we call this branch the Eα{}^{\perp}E_{\alpha}-branch of the curve (or left orthogonal branch), and characters 𝐯\mathbf{v} on the branch to the left of μ=α\mu=\alpha satisfy χ(Eα,𝐯)=0\chi(E_{\alpha},\mathbf{v})=0 and we call this the EαE_{\alpha}^{\perp}-branch (or right orthogonal branch).

Refer to caption
Figure 1. The Drézet-Le Potier curve Δ=δ(μ)\Delta=\delta(\mu)

Let 𝐯K(2)\mathbf{v}\in K(\mathbb{P}^{2}) be a character and set ch0(𝐯)=r\text{ch}_{0}(\mathbf{v})=r, μ(𝐯)=μ\mu(\mathbf{v})=\mu, and Δ(𝐯)=Δ\Delta(\mathbf{v})=\Delta. When 𝐯\mathbf{v} is stable with r>0r>0, Drézet and Le Potier showed the moduli space of stable sheaves with Chern character 𝐯\mathbf{v} is a normal, irreducible, factorial projective variety of dimension

dimM(𝐯)=r2(2Δ1)+1\dim M(\mathbf{v})=r^{2}(2\Delta-1)+1

([LP97, Theorem 17.0.1]). When 𝐯\mathbf{v} is a non-exceptional stable character of rank >1>1, the general bundle EM(𝐯)E\in M(\mathbf{v}) is slope-stable ([DLP85, Corollaire 4.12]). Slope-stability allows us to use elementary modifications:

2.2. Elementary modifications

If Δ0\Delta_{0} is the discriminant of a stable non-exceptional sheaf EE, then by integrality of the Euler characteristic one can see that the discriminant of another stable sheaf of slope μ(E)\mu(E) and rank r(E)r(E) differs from Δ0\Delta_{0} by an integral multiple of 1r(E)\frac{1}{r(E)}. One can in fact obtain any such discriminant larger than or equal to δ(μ(E))\delta(\mu(E)) by taking elementary modifications: choosing a point p2p\in\mathbb{P}^{2} and a surjection E𝒪pE\rightarrow\mathcal{O}_{p} one forms the exact sequence

0EE𝒪p0.0\rightarrow E^{\prime}\rightarrow E\rightarrow\mathcal{O}_{p}\rightarrow 0.

The sheaf EE^{\prime} is not locally free, but it has r(E)=r(E)r(E^{\prime})=r(E), μ(E)=μ(E)\mu(E^{\prime})=\mu(E), and Δ(E)=Δ(E)+1/r(E)\Delta(E^{\prime})=\Delta(E)+1/r(E). One can check that when EE is slope-stable, so is EE^{\prime} (but this is not true for stability). In this way one can construct slope-stable bundles of discriminant Δ0+k/r(E)\Delta_{0}+k/r(E) for any non-negative integer kk. See [CH20, Lemma 2.7] for details.

2.3. Brill-Noether loci

We now define the Brill-Noether loci and endow them with a natural determinantal structure, which leads to a lower bound on the dimension of their components.

The expected value of h0(E)h^{0}(E) when EE is stable is provided by the following well-known theorem due to Göttsche-Hirschowitz.

Theorem 2.1 ([GH98]).

When ch0(𝐯)2\text{ch}_{0}(\mathbf{v})\geq 2, the general sheaf EM(𝐯)E\in M(\mathbf{v}) has at most one nonzero cohomology group:

  1. (1)

    if χ(E)0\chi(E)\geq 0 and μ(E)>3\mu(E)>-3, then h1(E)=h2(E)=0h^{1}(E)=h^{2}(E)=0;

  2. (2)

    if χ(E)0\chi(E)\geq 0 and μ(E)<3\mu(E)<-3, then h0(E)=h1(E)=0h^{0}(E)=h^{1}(E)=0;

  3. (3)

    if χ(E)<0\chi(E)<0, h0(E)=h2(E)=0h^{0}(E)=h^{2}(E)=0.

When a sheaf FM(𝐯)F\in M(\mathbf{v}) has cohomology groups as prescribed in Theorem 2.1, we will say that it is cohomologically general or has general cohomology. By semicontinuity, when r2r\geq 2 the locus of sheaves with general cohomology forms an open subset of the moduli space; the Brill-Noether loci Bk(𝐯)B^{k}(\mathbf{v}) with k>χ(𝐯)k>\chi(\mathbf{v}) make up the complement of this open set. Let Ms(𝐯)M(𝐯)M^{s}(\mathbf{v})\subseteq M(\mathbf{v}) denote the locus of strictly stable sheaves.

Definition 2.2.

For a stable Chern character 𝐯K(2)\mathbf{v}\in K(\mathbb{P}^{2}), the kkth Brill-Noether locus Bk(𝐯)M(𝐯)B^{k}(\mathbf{v})\subseteq M(\mathbf{v}) is defined as a set to be the closure

Bk(𝐯)={EMs(𝐯):h0(E)k}¯M(𝐯).B^{k}(\mathbf{v})=\overline{\{E\in M^{s}(\mathbf{v}):h^{0}(E)\geq k\}}\subseteq M(\mathbf{v}).

Bk(𝐯)B^{k}(\mathbf{v}) is the locus where h0h^{0} jumps by at least kχ(𝐯)k-\chi(\mathbf{v}).

The Brill-Noether loci are clearly nested: Bk(𝐯)Bk+1(𝐯)B_{k}(\mathbf{v})\supseteq B_{k+1}(\mathbf{v}) for all kk.

Remark 2.3.

We restrict to stable sheaves in the definition because the number of global sections is not constant in SS-equivalence classes, so h0([E])h^{0}([E]) is not defined for strictly semistable points [E]M(𝐯)[E]\in M(\mathbf{v}).

Note that for a stable sheaf FF with μ(F)<0\mu(F)<0, any map 𝒪F\mathcal{O}\rightarrow F corresponding to 0sH0(F)0\neq s\in H^{0}(F) would be destabilizing, so H0(F)=0H^{0}(F)=0 and there are no Brill-Noether loci in the associated moduli space. Thus in what follows we assume the slope is nonnegative. In these cases, we expect h0(F)=max{0,χ(F)}h^{0}(F)=\max\{0,\chi(F)\}. When μ<3\mu<-3, we have h0(E)=0h^{0}(E)=0 but h2(E)=h0(E(3))h^{2}(E)=h^{0}(E^{\vee}(-3)) by Serre-duality, where μ(E(3))>0\mu(E^{\vee}(-3))>0. Thus h2h^{2}-jumping loci are Serre-dual to h0h^{0}-jumping loci when the slope is sufficiently negative. When 3<μ(E)<0-3<\mu(E)<0, Serre duality implies that h0(E)=h2(E)=0h^{0}(E)=h^{2}(E)=0, so there are no Brill-Noether loci for any cohomology.

We endow Bk(𝐯)B^{k}(\mathbf{v}) with a determinantal scheme structure, as follows (cf. [CHW17, Proposition 2.6] and [CMR10, §2]). Let /S\mathscr{E}/S be a proper flat family of stable sheaves on 2\mathbb{P}^{2} of Chern character 𝐯\mathbf{v}. We will examine the relative Brill-Noether locus BSk(𝐯)B^{k}_{S}(\mathbf{v}); when Ms(𝐯)M^{s}(\mathbf{v}) admits a universal family 𝒰\mathscr{U}, this will give the appropriate scheme structure on Bk(𝐯)B^{k}(\mathbf{v}). In general, one can work on the stack s(𝐯)\mathcal{M}^{s}(\mathbf{v}) and take the image in the coarse moduli space Ms(𝐯)M^{s}(\mathbf{v}) (see [Alp13, Theorem 4.16 and Example 8.7]).

Choose sSs\in S and let d0d\gg 0 be sufficiently large and let C2C\subseteq\mathbb{P}^{2} be a general-enough curve of degree dd, chosen so that the singularities of s\mathscr{E}_{s} do not meet CC. After replacing SS by an open subset, we can assume CC does not meet the singularities of any member of \mathscr{E}. Let p:S×2Sp:S\times\mathbb{P}^{2}\rightarrow S and q:S×22q:S\times\mathbb{P}^{2}\rightarrow\mathbb{P}^{2} be the projections. Form the exact sequence

0q𝒪(d)(q𝒪(d)|C)00\rightarrow\mathscr{E}\rightarrow\mathscr{E}\otimes q^{*}\mathcal{O}(d)\rightarrow\mathscr{E}\otimes(q^{*}\mathcal{O}(d)|_{C})\rightarrow 0

of sheaves on S×2S\times\mathbb{P}^{2}.

Choosing dd large enough so that each of the cohomology groups H1(sq𝒪(d))H^{1}(\mathscr{E}_{s}\otimes q^{*}\mathcal{O}(d)) vanish, we conclude R1p(q𝒪(d))=0R^{1}p_{*}(\mathscr{E}\otimes q^{*}\mathcal{O}(d))=0 and we obtain the exact sequence

0pp(q𝒪(d))ϕp((q𝒪(d)|C))R1p()00\rightarrow p_{*}\mathscr{E}\rightarrow p_{*}(\mathscr{E}\otimes q^{*}\mathcal{O}(d))\stackrel{{\scriptstyle\phi}}{{\rightarrow}}p_{*}(\mathscr{E}\otimes(q^{*}\mathcal{O}(d)|_{C}))\rightarrow R^{1}p_{*}(\mathscr{E})\rightarrow 0

of sheaves on SS. When dd is sufficiently large the map ϕ\phi is a map between bundles of the same rank, and the kkth determinantal variety associated to ϕ\phi is supported on those sSs\in S such that h0(s)kh^{0}(\mathscr{E}_{s})\geq k, or equivalently h1(s)kχ(𝐯)h^{1}(\mathscr{E}_{s})\geq k-\chi(\mathbf{v}). Thus scheme-theoretically we define Bk(𝐯)B^{k}(\mathbf{v}) to be this determinantal variety, and observe that its support agrees with the set described above.

The construction above does not depend on the choice of dd or CC; precisely, over an open subset UU of SS over which the source and target of ϕ\phi are trivial, the Fitting ideal Fitt0(R1p)\text{Fitt}_{0}(R^{1}p_{*}\mathcal{E}) is generated by detϕ\det\phi, viewed as a section of 𝒪U\mathcal{O}_{U}. See [Eis13, Chapter 20]. The Fitting ideal is the ideal sheaf of Bk(𝐯)B^{k}(\mathbf{v}), and gives the desired scheme structure.

Being a determinantal variety, the codimension of a component ZZ of Bk(𝐯)B^{k}(\mathbf{v}) is bounded:

codimM(𝐯)(Z)k(kχ(𝐯)).\text{codim}_{M(\mathbf{v})}(Z)\leq k(k-\chi(\mathbf{v})).

Thus since, for a general bundle EM(𝐯)E\in M(\mathbf{v}), we have

dim(M(𝐯))=ext1(E,E)=r(𝐯)2(2Δ(𝐯)1)+1\dim(M(\mathbf{v}))=\text{ext}^{1}(E,E)=r(\mathbf{v})^{2}(2\Delta(\mathbf{v})-1)+1

the dimension of a component ZZ of Bk(𝐯)B^{k}(\mathbf{v}) is bounded below:

dim(Z)r(𝐯)2(2Δ(𝐯)1)+1k(kχ(𝐯)).\dim(Z)\geq r(\mathbf{v})^{2}(2\Delta(\mathbf{v})-1)+1-k(k-\chi(\mathbf{v})).
Definition 2.4.

The expected codimension of the kkth Brill-Noether locus Bk(𝐯)M(𝐯)B^{k}(\mathbf{v})\subseteq M(\mathbf{v}) is

expcodimk(𝐯)=k(kχ(𝐯)).\text{expcodim}^{k}(\mathbf{v})=k(k-\chi(\mathbf{v})).

3. Extension parametrizations of moduli spaces

In this section we study the stability of extension sheaves. When extension sheaves are semistable, we obtain families of sheaves over loci in the associated moduli space.

Definition 3.1.

Suppose that 𝐯,𝐯,𝐯′′K(2)\mathbf{v},\mathbf{v}^{\prime},\mathbf{v}^{\prime\prime}\in K(\mathbb{P}^{2}) are stable characters with 𝐯=𝐯+𝐯′′\mathbf{v}=\mathbf{v}^{\prime}+\mathbf{v}^{\prime\prime}. Let /S\mathscr{E}^{\prime}/S^{\prime} and ′′/S′′\mathscr{E}^{\prime\prime}/S^{\prime\prime} be families of semistable sheaves of Chern characters 𝐯\mathbf{v}^{\prime} and 𝐯′′\mathbf{v}^{\prime\prime}, respectively. If the general extension EE

0sEs′′′′00\rightarrow\mathscr{E}^{\prime}_{s^{\prime}}\rightarrow E\rightarrow\mathscr{E}^{\prime\prime}_{s^{\prime\prime}}\rightarrow 0

with sSs^{\prime}\in S^{\prime} and s′′S′′s^{\prime\prime}\in S^{\prime\prime} is semi-stable, then an induced rational map

Ext1(s′′′′,s){\mathbb{P}\text{Ext}^{1}(\mathscr{E}^{\prime\prime}_{s^{\prime\prime}},\mathscr{E}^{\prime}_{s^{\prime}})}M(𝐯){M(\mathbf{v})}S×S′′{S^{\prime}\times S^{\prime\prime}}

defined on a locus where Ext1(s′′′′,s)\text{Ext}^{1}(\mathscr{E}^{\prime\prime}_{s^{\prime\prime}},\mathscr{E}^{\prime}_{s^{\prime}}) has constant rank, is called an extension parametrization associated to M(𝐯)M(\mathbf{v}).

Stability of these extensions is the central technical challenge of this paper. However, there is a class called the extremal extensions whose stability is not hard to show, which we now recall.

3.1. Extremal extensions

Stability of extension sheaves is known when the slope of the subsheaf is extremal with respect to the extension sheaf. The following definition is derived from [CH16, Definition 4.1], but we make some slight modifications for our purposes.

Definition 3.2.

A Chern character 𝐯\mathbf{v}^{\prime} on 2\mathbb{P}^{2} is called extremal for 𝐯\mathbf{v} if 𝐯\mathbf{v}^{\prime} satisfies:

  1. (D1)

    ch0(𝐯)<ch0(𝐯)\text{ch}_{0}(\mathbf{v}^{\prime})<\text{ch}_{0}(\mathbf{v});

  2. (D2)

    μ(𝐯)μ(𝐯)\mu(\mathbf{v}^{\prime})\leq\mu(\mathbf{v});

  3. (D3)

    𝐯\mathbf{v}^{\prime} is stable;

and furthermore, for any 𝐰K(2)\mathbf{w}\in K(\mathbb{P}^{2}) satisfying (D1)-(D3):

  1. (E1)

    μ(𝐰)μ(𝐯)\mu(\mathbf{w})\leq\mu(\mathbf{v}^{\prime});

  2. (E2)

    If μ(𝐰)=μ(𝐯)\mu(\mathbf{w})=\mu(\mathbf{v}^{\prime}), then Δ(𝐰)Δ(𝐯)\Delta(\mathbf{w})\geq\Delta(\mathbf{v}^{\prime});

  3. (E3)

    If μ(𝐰)=μ(𝐯)\mu(\mathbf{w})=\mu(\mathbf{v}^{\prime}) and Δ(𝐰)=Δ(𝐯)\Delta(\mathbf{w})=\Delta(\mathbf{v}^{\prime}), then ch0(𝐯)ch0(𝐰)\text{ch}_{0}(\mathbf{v}^{\prime})\leq\text{ch}_{0}(\mathbf{w}).

A triple Ξ=(𝐯,𝐯,𝐯′′)\Xi=(\mathbf{v}^{\prime},\mathbf{v},\mathbf{v}^{\prime\prime}) of Chern characters on 2\mathbb{P}^{2} is called extremal if 𝐯+𝐯′′=𝐯\mathbf{v}^{\prime}+\mathbf{v}^{\prime\prime}=\mathbf{v}, where 𝐯\mathbf{v}^{\prime} is the extremal character for 𝐯\mathbf{v}, and 𝐯′′\mathbf{v}^{\prime\prime} is stable.

It is not obvious a priori that extremal decompositions exist for any 𝐯K(2)\mathbf{v}\in K(\mathbb{P}^{2}) since 𝐯′′\mathbf{v}^{\prime\prime} may not be stable, but at least it is clear when 𝐯\mathbf{v} has large discriminant, see e.g. [CH16, Lemma 4.3], or simply observe that Δ(𝐯′′)0\Delta(\mathbf{v}^{\prime\prime})\gg 0 when Δ(𝐯)0\Delta(\mathbf{v})\gg 0.

Our main use of Definition 3.2 will be to construct extension parametrizations associated to moduli spaces. The following simple observation is crucial.

Proposition 3.3.

Suppose Ξ=(𝐯,𝐯,𝐯′′)\Xi=(\mathbf{v}^{\prime},\mathbf{v},\mathbf{v}^{\prime\prime}) is an extremal triple, and we have a nonsplit short exact sequence

0EEE′′00\rightarrow E^{\prime}\rightarrow E\rightarrow E^{\prime\prime}\rightarrow 0

with EM(𝐯)E^{\prime}\in M(\mathbf{v}^{\prime}) and E′′M(𝐯′′)E^{\prime\prime}\in M(\mathbf{v}^{\prime\prime}). Then EE is semistable.

Proof.

Suppose that EQE\rightarrow Q is a destabilizing quotient of EE, so that μ(Q)μ(E)\mu(Q)\leq\mu(E) with r(Q)<r(E)r(Q)<r(E). After perhaps passing to a further quotient of QQ, we may assume QQ is stable. Slope-closeness (E1) implies that μ(Q)μ(E)\mu(Q)\leq\mu(E^{\prime}). If μ(Q)<μ(E)\mu(Q)<\mu(E^{\prime}), the composition EEQE^{\prime}\rightarrow E\rightarrow Q is zero by stability. So there is an induced map E′′QE^{\prime\prime}\rightarrow Q. However μ(𝐯′′)μ(𝐯)>μ(Q)\mu(\mathbf{v}^{\prime\prime})\geq\mu(\mathbf{v})>\mu(Q), hence the induced map is zero. This implies that EQE\rightarrow Q vanishes.

We now have μ(Q)=μ(E)\mu(Q)=\mu(E^{\prime}). Discriminant-minimality (E2) implies that Δ(Q)Δ(E)\Delta(Q)\geq\Delta(E^{\prime}). If Δ(Q)>Δ(E)\Delta(Q)>\Delta(E^{\prime}), we get a contradiction as in the preceding paragraph.

Hence μ(Q)=μ(E)\mu(Q)=\mu(E^{\prime}) and Δ(Q)=Δ(E)\Delta(Q)=\Delta(E^{\prime}). By (E3), E,QE^{\prime},Q are stable. Since EE is a non-split extension, EQE^{\prime}\rightarrow Q is zero, inducing a map E′′QE^{\prime\prime}\rightarrow Q. Now by (D1), (D2) this map can be non-zero only if μ(𝐯′′)=μ(𝐯)=μ(𝐯)\mu(\mathbf{v}^{\prime\prime})=\mu(\mathbf{v})=\mu(\mathbf{v}^{\prime}) and Δ(𝐯′′)=Δ(𝐯)=Δ(𝐯)\Delta(\mathbf{v}^{\prime\prime})=\Delta(\mathbf{v})=\Delta(\mathbf{v}^{\prime}). Hence EE is semi-stable. ∎

Corollary 3.4.

Let 𝐯K(2)\mathbf{v}\in K(\mathbb{P}^{2}) be a stable character which admits an extremal decomposition (𝐯,𝐯,𝐯′′)(\mathbf{v}^{\prime},\mathbf{v},\mathbf{v}^{\prime\prime}), and let /S\mathscr{E}^{\prime}/S^{\prime} and ′′/S′′\mathscr{E}^{\prime\prime}/S^{\prime\prime} be families of semistable sheaves of characters 𝐯\mathbf{v}^{\prime} and 𝐯′′\mathbf{v}^{\prime\prime} respectively. Then the induced extension parametrization

Ext1(s′′′′,s){\mathbb{P}\mathrm{Ext}^{1}(\mathscr{E}^{\prime\prime}_{s^{\prime\prime}},\mathscr{E}^{\prime}_{s^{\prime}})}M(𝐯){M(\mathbf{v})}S×S′′{S^{\prime}\times S^{\prime\prime}}

defined over the locus where Ext1(s′′′′,s)\mathrm{Ext}^{1}(\mathscr{E}^{\prime\prime}_{s^{\prime\prime}},\mathscr{E}^{\prime}_{s^{\prime}}) has minimal rank, exists, where sSs^{\prime}\in S^{\prime} and s′′S′′s^{\prime\prime}\in S^{\prime\prime}.

Proof.

By Proposition 3.3, for any sSs^{\prime}\in S^{\prime} and s′′S′′s^{\prime\prime}\in S^{\prime\prime}, a nonsplit extension

0sEs′′′′00\rightarrow\mathscr{E}^{\prime}_{s^{\prime}}\rightarrow E\rightarrow\mathscr{E}^{\prime\prime}_{s^{\prime\prime}}\rightarrow 0

is semistable of Chern character 𝐯\mathbf{v}. So on any locus in S×S′′S^{\prime}\times S^{\prime\prime} where Ext1(s′′′′,s)\text{Ext}^{1}(\mathscr{E}^{\prime\prime}_{s^{\prime\prime}},\mathscr{E}^{\prime}_{s^{\prime}}) is of constant rank, there is an induced map M(𝐯)\mathbb{P}\dashrightarrow M(\mathbf{v}), where \mathbb{P} is the projective bundle over S×S′′S^{\prime}\times S^{\prime\prime} with fiber Ext1(s′′′′,s)\mathbb{P}\text{Ext}^{1}(\mathscr{E}^{\prime\prime}_{s^{\prime\prime}},\mathscr{E}^{\prime}_{s^{\prime}}) over (s,s′′)S×S′′(s^{\prime},s^{\prime\prime})\in S^{\prime}\times S^{\prime\prime}. ∎

Remark 3.5.

When the subsheaf is not extremal for the extension sheaf, in general additional conditions are required to prove generic stability, see for instance Lemma 4.5 and Lemma 5.9.

4. Emptiness & non-emptiness of Brill-Noether loci

For a stable Chern character 𝐯K(2)\mathbf{v}\in K(\mathbb{P}^{2}) there are only finitely many values of k0k\geq 0 such that Bk(𝐯)M(𝐯)B^{k}(\mathbf{v})\subseteq M(\mathbf{v}) is nonempty. That is, there are only finitely many values of h0(E)h^{0}(E) for a stable sheaf EM(𝐯)E\in M(\mathbf{v}). We first give a bound on this quantity. Later on in this section we will analyze which such values are achieved.

Theorem 4.1.

If EE is a slope-semistable sheaf on 2\mathbb{P}^{2} with Chern character 𝐯\mathbf{v} such that ch1(𝐯)0\text{ch}_{1}(\mathbf{v})\geq 0, then

h0(E)max{p(ch1(𝐯)),ch0(𝐯)},h^{0}(E)\leq\max\{p(\text{ch}_{1}(\mathbf{v})),\text{ch}_{0}(\mathbf{v})\},

where p(x)=p𝒪2(x)=x2+3x+22p(x)=p_{\mathcal{O}_{\mathbb{P}^{2}}}(x)=\frac{x^{2}+3x+2}{2}.

In the proof we will consider partial evaluation maps V𝒪EV\otimes\mathcal{O}\rightarrow E for a subspace VH0(E)V\subseteq H^{0}(E). We fix EE and VV for the following lemma.

Lemma 4.2.

Let FEF\subseteq E be the image sheaf of the evaluation map V𝒪EV\otimes\mathcal{O}\rightarrow E, and let 0=F0F1Fm=F0=F_{0}\subseteq F_{1}\subseteq\cdots\subseteq F_{m}=F be the Harder-Narasimhan filtration of FF, with graded pieces gri=Fi/Fi1\text{gr}_{i}=F_{i}/F_{i-1}.

If r(F)<r(E)r(F)<r(E), then c1(gri)<c1(E)c_{1}(\text{gr}_{i})<c_{1}(E) for each ii.

Proof.

If not, then since r(gri)r(F)<r(E)r(\text{gr}_{i})\leq r(F)<r(E),

μ(E)=c1(E)r(E)c1(gri)r(E)<c1(gri)r(gri)=μ(gri).\mu(E)=\frac{c_{1}(E)}{r(E)}\leq\frac{c_{1}(\text{gr}_{i})}{r(E)}<\frac{c_{1}(\text{gr}_{i})}{r(\text{gr}_{i})}=\mu(\text{gr}_{i}).

(Note that we need c1(E)0c_{1}(E)\geq 0 for VV to be nonempty.) The inclusions gr1=F1FE\text{gr}_{1}=F_{1}\hookrightarrow F\hookrightarrow E and semistability of EE imply

μ(gr1)μ(E)<μ(gri)\mu(\text{gr}_{1})\leq\mu(E)<\mu(\text{gr}_{i})

contradicting that the slopes of the gri\text{gr}_{i} are non-increasing. We conclude that c1(gri)<c1(E)c_{1}(\text{gr}_{i})<c_{1}(E) for each ii. ∎

We observe for use in the cases below that the function p(x)/xp(x)/x is increasing for integers x2x\geq 2, with p(1)=p(2)/2p(1)=p(2)/2.

Proof of Theorem 4.1.

First, since we have EEE\subseteq E^{\vee\vee}, we have h0(E)h0(E)h^{0}(E)\leq h^{0}(E^{\vee\vee}). Because EE^{\vee\vee} is slope-semistable with r(E)=r(E)r(E^{\vee\vee})=r(E) and c1(E)=c1(E)c_{1}(E^{\vee\vee})=c_{1}(E), it suffices to prove the theorem when EE is locally free.

Case I. In this case we assume r(E)=ch0(𝐯)p(ch1(𝐯))r(E)=\text{ch}_{0}(\mathbf{v})\geq p(\text{ch}_{1}(\mathbf{v})), and proceed by induction on r(E)r(E). In this case when r(E)=1r(E)=1 we have c1(E)0c_{1}(E)\leq 0, and when c1(E)<0c_{1}(E)<0 slope-semistability implies h0(E)=0h^{0}(E)=0, as desired. When c1(E)=0c_{1}(E)=0, it follows that EE is the ideal sheaf of a finite length subscheme Z2Z\subseteq\mathbb{P}^{2}. Thus h0(E)1h^{0}(E)\leq 1, and equality holds if and only if ZZ is empty.

We first show that any r(E)r(E)-dimensional subspace VH0(E)V\subseteq H^{0}(E) spans a full rank subsheaf of EE. That is, the evaluation map V𝒪EV\otimes\mathcal{O}\rightarrow E is full rank. Suppose not, so that r(F)<r(E)r(F)<r(E). To show this we first claim that each gri\text{gr}_{i} has r(gri)p(c1(gri))r(\text{gr}_{i})\geq p(c_{1}(\text{gr}_{i})), so by induction on the rank we have h0(gri)r(gri)h^{0}(\text{gr}_{i})\leq r(\text{gr}_{i}). It follows that

h0(F)h0(gri)r(gri)=r(F)<r(E).h^{0}(F)\leq\sum h^{0}(\text{gr}_{i})\leq\sum r(\text{gr}_{i})=r(F)<r(E).

Since by definition h0(F)r(E)h^{0}(F)\geq r(E), we conclude from this contradiction that any partial evaluation map of any r(E)r(E) independent sections in this case is of full rank.

To check the claim, assume otherwise that r(gri)<p(c1(gri))r(\text{gr}_{i})<p(c_{1}(\text{gr}_{i})). It is enough to produce a contradiction for those gri\text{gr}_{i} with c1(gri)>0c_{1}(\text{gr}_{i})>0, since when c1(gri)0c_{1}(\text{gr}_{i})\leq 0 we have h0(gri)r(gri)h^{0}(\text{gr}_{i})\leq r(\text{gr}_{i}). Then

μ(E)=c1(E)r(E)c1(E)p(c1(E))c1(gri)p(c1(gri))<c1(gri)r(gri)=μ(gri),\mu(E)=\frac{c_{1}(E)}{r(E)}\leq\frac{c_{1}(E)}{p(c_{1}(E))}\leq\frac{c_{1}(\text{gr}_{i})}{p(c_{1}(\text{gr}_{i}))}<\frac{c_{1}(\text{gr}_{i})}{r(\text{gr}_{i})}=\mu(\text{gr}_{i}),

where we have used Lemma 4.2 and that x/p(x)x/p(x) is non-increasing when x1x\geq 1. The slopes μ(gri)\mu(\text{gr}_{i}) in the Harder-Narasimhan filtration are decreasing and EE is semistable, so we have

μ(gri)μ(gr1)μ(E)\mu(\text{gr}_{i})\leq\mu(\text{gr}_{1})\leq\mu(E)

so this contradiction proves the claim. We conclude that any r(E)r(E) sections span a full rank subsheaf of EE.

Therefore an r(E)r(E)-dimensional subspace VH0(E)V\subseteq H^{0}(E) determines a nonzero section of 𝒪(c1(E))\mathcal{O}(c_{1}(E)) via the natural map

r(E)H0(E)H0(r(E)E)=H0(𝒪(c1(E))),\wedge^{r(E)}H^{0}(E)\rightarrow H^{0}(\wedge^{r(E)}E)=H^{0}(\mathcal{O}(c_{1}(E))),

where we have rE=𝒪(c1(E))\wedge^{r}E=\mathcal{O}(c_{1}(E)) because EE is locally free. In this way we obtain a morphism

f:G(r(E),H0(E))H0(𝒪(c1(E))).f:G(r(E),H^{0}(E))\rightarrow\mathbb{P}H^{0}(\mathcal{O}(c_{1}(E))).

When h0(E)r(E)+1h^{0}(E)\geq r(E)+1, it is easy to see that ff is nonconsant. Indeed, any r(E)r(E)-dimensional subspace VH0(E)V\subseteq H^{0}(E) sits in a short exact sequence

0V𝒪EQ00\rightarrow V\otimes\mathcal{O}\rightarrow E\rightarrow Q\rightarrow 0

because the evaluation map is of full rank. Choose sH0(E)Vs\in H^{0}(E)\smallsetminus V, and consider a general point pSupp(Q)p\in\text{Supp}(Q). Then 0spH0(Q|p)0\neq s_{p}\in H^{0}(Q|_{p}) does not lie in V|pE|pV|_{p}\subseteq E|_{p}. Choosing a generating set s1,,sr(E)s_{1},...,s_{r(E)} for the span sp,V|pEp\langle s_{p},V|_{p}\rangle\subseteq E_{p}; then the sis_{i} span a rank r(E)r(E) subspace VH0(E)V^{\prime}\subseteq H^{0}(E) such that the induced section r(E)VH0(r(E)E)\wedge^{r(E)}V^{\prime}\in H^{0}(\wedge^{r(E)}E) is independent from r(E)V\wedge^{r(E)}V. For instance, the multiplicity of r(E)V\wedge^{r(E)}V^{\prime} at pp is lower than that of r(E)V\wedge^{r(E)}V.

However if h0(E)r(E)+1h^{0}(E)\geq r(E)+1, we have

dimG(r(E),H0(E))r(E)p(c1(E))\dim G(r(E),H^{0}(E))\geq r(E)\geq p(c_{1}(E))

since EE has r(E)p(c1(E))r(E)\geq p(c_{1}(E)), and dimH0(𝒪(c1(E)))=p(c1(E))1\dim\mathbb{P}H^{0}(\mathcal{O}(c_{1}(E)))=p(c_{1}(E))-1. Since the Picard rank of a Grassmannian is 1, there is no non-constant map from a Grassmannian to a lower-dimensional projective variety. We conclude that h0(E)r(E)h^{0}(E)\leq r(E), as desired.

Case II. We are now in the situation that r(E)<p(c1(E))=c1(E)2+3c1(E)+22r(E)<p(c_{1}(E))=\frac{c_{1}(E)^{2}+3c_{1}(E)+2}{2}, and we want to prove h0(E)p(c1(E))h^{0}(E)\leq p(c_{1}(E)). We will use induction on c1(E)c_{1}(E). If c1(E)=0c_{1}(E)=0, then p(c1(E))=1p(c_{1}(E))=1 and there is nothing to prove.

We first show that any p(c1(E))p(c_{1}(E))-dimensional subspace VH0(E)V\subseteq H^{0}(E) spans a full rank subsheaf of EE. We proceed by contradiction, so we let FF be the image of the evaluation map V𝒪EV\otimes\mathcal{O}\rightarrow E and assume r(F)<r(E)r(F)<r(E). Since p(x)/xp(x)/x is non-decreasing, when c1(gri)1c_{1}(\text{gr}_{i})\geq 1 we have

p(c1(gri))=c1(gri)p(c1(gri))c1(gri)c1(gri)p(c1(E))c1(E)p(c1(E))r(gri)r(E)p(c_{1}(\text{gr}_{i}))=c_{1}(\text{gr}_{i})\frac{p(c_{1}(\text{gr}_{i}))}{c_{1}(\text{gr}_{i})}\leq c_{1}(\text{gr}_{i})\frac{p(c_{1}(E))}{c_{1}(E)}\leq p(c_{1}(E))\frac{r(\text{gr}_{i})}{r(E)}

by Lemma 4.2. The gri\text{gr}_{i} may fall into either Case I or Case II. When gri\text{gr}_{i} is in Case II, we obtain

h0(gri)p(c1(gri))p(c1(E))r(gri)r(E)h^{0}(\text{gr}_{i})\leq p(c_{1}(\text{gr}_{i}))\leq p(c_{1}(E))\frac{r(\text{gr}_{i})}{r(E)}

by induction on c1c_{1}. When c1(gri)<1c_{1}(\text{gr}_{i})<1, we have

h0(gri)r(gri)=r(E)r(gri)r(E)p(c1(E))r(gri)r(E)h^{0}(\text{gr}_{i})\leq r(\text{gr}_{i})=r(E)\frac{r(\text{gr}_{i})}{r(E)}\leq p(c_{1}(E))\frac{r(\text{gr}_{i})}{r(E)}

by the assumption that EE is in Case II. For those gri\text{gr}_{i} in Case I with r(gri)p(c1(gri))r(\text{gr}_{i})\geq p(c_{1}(\text{gr}_{i})), we also have

h0(gri)r(gri)=r(E)r(gri)r(E)<p(c1(E))r(gri)r(E).h^{0}(\text{gr}_{i})\leq r(\text{gr}_{i})=r(E)\frac{r(\text{gr}_{i})}{r(E)}<p(c_{1}(E))\frac{r(\text{gr}_{i})}{r(E)}.

We now have

(2) h0(F)h0(gri)p(c1(E))r(gri)r(E)=p(c1(E))r(F)r(E)<p(c1(E)).h^{0}(F)\leq\sum h^{0}(\text{gr}_{i})\leq\sum p(c_{1}(E))\frac{r(\text{gr}_{i})}{r(E)}=p(c_{1}(E))\frac{r(F)}{r(E)}<p(c_{1}(E)).

By definition we have h0(F)p(c1(E))h^{0}(F)\geq p(c_{1}(E)), and this contradiction proves that any p(c1(E))p(c_{1}(E))-dimensional subspace VV spans a full rank subsheaf of EE.

We can additionally assume that the dependency locus of VV is zero-dimensional. If not, consider the image FEF\subseteq E of the evaluation map V𝒪EV\otimes\mathcal{O}\rightarrow E and its Harder-Narasimhan filtration. Then c1(F)<c1(E)c_{1}(F)<c_{1}(E). By (2), we know that h0(F)p(c1(E))h^{0}(F)\leq p(c_{1}(E)) (only the final inequality in (2) is no longer strict), and equality holds only if for every ii, h0(gri)=p(c1(E))r(gri)r(E)h^{0}(\text{gr}_{i})=p(c_{1}(E))\frac{r(\text{gr}_{i})}{r(E)} for every graded piece. This is true only if c1(gri)>0c_{1}(\text{gr}_{i})>0; r(gri)<p(c1(gri))r(\text{gr}_{i})<p(c_{1}(\text{gr}_{i})); c1(gri)=1c_{1}(\text{gr}_{i})=1; and c1(E)=2c_{1}(E)=2. In this case, c1(F)<c1(E)=2c_{1}(F)<c_{1}(E)=2, c1(F)1c_{1}(F)\leq 1. If FF is unstable, then there is at least one Harder-Narasimhan factor gri\text{gr}_{i} whose degree is non-positive, which is not the case. If FF is stable, it satisfies r(F)<p(c1(F))r(F)<p(c_{1}(F)), then by induction, h0(F)p(c1(F))<p(c1(E))h^{0}(F)\leq p(c_{1}(F))<p(c_{1}(E)). In any case, we have h0(F)<p(c1(E))h^{0}(F)<p(c_{1}(E)), a contradiction.

We now return to the bound on h0(E)h^{0}(E). We first prove h0(E)p(c1(E))h^{0}(E)\leq p(c_{1}(E)) when 0μ(E)<10\leq\mu(E)<1. We have just shown that any p(c1(E))p(c_{1}(E)) sections of EE span a full rank subsheaf, and their dependency locus is zero-dimensional. Let L2L\subseteq\mathbb{P}^{2} be a general line. Consider

0E(1)EE|L00\rightarrow E(-1)\rightarrow E\rightarrow E|_{L}\rightarrow 0

and write E|L=i=1r(E)𝒪L(di)E|_{L}=\bigoplus_{i=1}^{r(E)}\mathcal{O}_{L}(d_{i}). Since μ(E)<1\mu(E)<1, there is at least one di0d_{i}\leq 0, say d1d_{1}. Taking H0H^{0} we obtain π:H0(E)i=1r(E)H0(𝒪L(di))\pi:H^{0}(E)\rightarrow\bigoplus_{i=1}^{r(E)}H^{0}(\mathcal{O}_{L}(d_{i})). Let Wi=1r(E)H0(𝒪L(di))W\subseteq\bigoplus_{i=1}^{r(E)}H^{0}(\mathcal{O}_{L}(d_{i})) be the image of π\pi. If Wi=2r(E)H0(𝒪L(di))W\subseteq\bigoplus_{i=2}^{r(E)}H^{0}(\mathcal{O}_{L}(d_{i})), then sections of EE are dependent along LL, hence h0(E)<p(c1)h^{0}(E)<p(c_{1}). If Wi=2r(E)H0(𝒪L(di))W\not\subseteq\bigoplus_{i=2}^{r(E)}H^{0}(\mathcal{O}_{L}(d_{i})), then d1=0d_{1}=0, and since π1(Wi=2r(E)H0(𝒪L(di)))\pi^{-1}(W\cap\bigoplus_{i=2}^{r(E)}H^{0}(\mathcal{O}_{L}(d_{i}))) has codimension 1, there are h0(E)1h^{0}(E)-1 sections dependent along LL, hence h0(E)1<p(c1)h^{0}(E)-1<p(c_{1}). That is, h0(E)p(c1(E))h^{0}(E)\leq p(c_{1}(E)), as desired.

Now if nμ<n+1n\leq\mu<n+1, we induct on nn. We have just shown the statement is true for n=0n=0; now assume n1n\geq 1. Again we use that any p(c1(E))p(c_{1}(E)) sections span a full rank subsheaf, and the dependency locus is supported in dimension zero. Consider again the restriction to a line:

0E(1)EE|L0.0\rightarrow E(-1)\rightarrow E\rightarrow E|_{L}\rightarrow 0.

Taking H0H^{0} we get h0(E)h0(E(1))+h0(E|L)h^{0}(E)\leq h^{0}(E(-1))+h^{0}(E|_{L}). By induction, h0(E(1))p(c1(E)r(E))h^{0}(E(-1))\leq p(c_{1}(E)-r(E)). Consider E|L=i=1r(E)𝒪L(di)E|_{L}=\bigoplus_{i=1}^{r(E)}\mathcal{O}_{L}(d_{i}). If some di<0d_{i}<0, say d1d_{1}, then the span of sections in WW is contained in i2𝒪L(di)\bigoplus_{i\geq 2}\mathcal{O}_{L}(d_{i}), and it follows that h0(E)<p(c1(E))h^{0}(E)<p(c_{1}(E)). Hence we assume di0d_{i}\geq 0 for all ii. Then h0(E|L)=i=1r(E)h0(𝒪L(di+1))=c1(E)+r(E)h^{0}(E|_{L})=\sum_{i=1}^{r(E)}h^{0}(\mathcal{O}_{L}(d_{i}+1))=c_{1}(E)+r(E). Now since n1n\geq 1, c1(E)r(E)1c_{1}(E)\geq r(E)\geq 1 and we obtain

r(E)(2c1(E)r(E)+1)2c1(E)\displaystyle r(E)(2c_{1}(E)-r(E)+1)-2c_{1}(E) =(r(E)1)2c1(E)r(E)2+r(E)\displaystyle=(r(E)-1)2c_{1}(E)-r(E)^{2}+r(E)
(r(E)1)2r(E)r(E)2+r(E)\displaystyle\geq(r(E)-1)2r(E)-r(E)^{2}+r(E)
=r(E)(r(E)1)0\displaystyle=r(E)(r(E)-1)\geq 0

hence r(E)(2c1(E)r(E)+1)2c1(E)r(E)(2c_{1}(E)-r(E)+1)\geq 2c_{1}(E), and

p(c1(E))p(c1(E)r(E))\displaystyle p(c_{1}(E))-p(c_{1}(E)-r(E)) =2c1(E)r(E)r(E)2+3r(E)2\displaystyle=\frac{2c_{1}(E)r(E)-r(E)^{2}+3r(E)}{2}
=r(E)+r(E)(2c1(E)r(E)+1)2\displaystyle=r(E)+\frac{r(E)(2c_{1}(E)-r(E)+1)}{2}
r(E)+c1(E).\displaystyle\geq r(E)+c_{1}(E).

In particular we have

h0(E)h0(E(1))+h0(E|L)p(c1(E)r(E))+(r(E)+c1(E))p(c1(E)).h^{0}(E)\leq h^{0}(E(-1))+h^{0}(E|_{L})\leq p(c_{1}(E)-r(E))+(r(E)+c_{1}(E))\leq p(c_{1}(E)).

This finishes the proof. ∎

We will later need the following corollary, which is an immediate consequence of Theorem 4.1.

Corollary 4.3.

If EE is a semistable sheaf on 2\mathbb{P}^{2} with c1(E)=1c_{1}(E)=1 and r(E)3r(E)\geq 3, then h0(E)r(E)h^{0}(E)\leq r(E).

Remark 4.4.

We can describe the behavior of EE with c1(E)=1c_{1}(E)=1 also when the rank is small. If the rank is 1, then EE is a twist of an ideal sheaf EIZ(1)E\simeq I_{Z}(1). Then h0(E)0h^{0}(E)\neq 0 if and only if ZZ is colinear, in which case

  • h0(E)=1h^{0}(E)=1 if (Z)2\ell(Z)\geq 2,

  • h0(E)=2h^{0}(E)=2 if (Z)=1\ell(Z)=1, and

  • h0(E)=3h^{0}(E)=3 if (Z)=0\ell(Z)=0.

If the rank is 2, suppose h0(E)3h^{0}(E)\geq 3 and consider the partial evaluation map V𝒪EV\otimes\mathcal{O}\rightarrow E for VH0(E)V\subseteq H^{0}(E) a rank 3 subspace. Let FF be the image of V𝒪V\otimes\mathcal{O} in EE. If the rank of FF is one, then it falls into one of the above cases. Since r(V)=3r(V)=3, we have h0(F)3h^{0}(F)\geq 3, so as above F=𝒪F=\mathcal{O}, but in this case the kernel K=ker(V𝒪F)K=\ker(V\otimes\mathcal{O}\rightarrow F) has h0(K)0h^{0}(K)\neq 0, which contradicts the definition of FF. It follows that r(F)=2r(F)=2, and in fact is semistable of slope 1/21/2.

This produces a short exact sequence

0KV𝒪F00\rightarrow K\rightarrow V\otimes\mathcal{O}\rightarrow F\rightarrow 0

with r(K)=1r(K)=1, c1(K)=1c_{1}(K)=-1. It follows that KIZ(1)K\simeq I_{Z}(-1) for some codimension two subscheme Z2Z\subseteq\mathbb{P}^{2}. In particular, ch2(K)=12(Z)\text{ch}_{2}(K)=\frac{1}{2}-\ell(Z). Thus

Δ(F)=12(12)2(Z)1/22=38(Z)2.\Delta(F)=\frac{1}{2}\left(\frac{1}{2}\right)^{2}-\frac{\ell(Z)-1/2}{2}=\frac{3}{8}-\frac{\ell(Z)}{2}.

Semistability additionally implies that Δ(F)=38+k2\Delta(F)=\frac{3}{8}+\frac{k}{2} for some k0k\geq 0 (see Section 2.2), from which we find k=(Z)0k=-\ell(Z)\leq 0, so k=0k=0 and FE=T2(1)F\simeq E=T_{\mathbb{P}^{2}}(-1). In all other cases, we again have h0(E)r(E)h^{0}(E)\leq r(E).

4.1. The Brill-Noether loci Br(𝐯)B^{r}(\mathbf{v})

In this section we show that Bk(𝐯)B^{k}(\mathbf{v})\neq\emptyset whenever kch0(𝐯)k\leq\text{ch}_{0}(\mathbf{v}). We can say a great deal about the loci of sheaves EE with h0(E)r=r(E)h^{0}(E)\geq r=r(E), i.e., the Brill-Noether loci Br(𝐯)B^{r}(\mathbf{v}). We will need the following statements to study the Brill-Noether loci when ch0(𝐯)p(ch1(𝐯))\text{ch}_{0}(\mathbf{v})\geq p(\text{ch}_{1}(\mathbf{v})).

Lemma 4.5.

Let 𝐯K(2)\mathbf{v}\in K(\mathbb{P}^{2}) be a stable Chern character and write r=ch0(𝐯)r=\text{ch}_{0}(\mathbf{v}) and c1=ch1(𝐯)c_{1}=\text{ch}_{1}(\mathbf{v}). Assume that rp(ch1(𝐯))r\geq p(\text{ch}_{1}(\mathbf{v})).

  1. (1)

    For every EBr(𝐯)E\in B^{r}(\mathbf{v}), evaluation map on global sections evE\text{ev}_{E} sits in an exact sequence

    0𝒪rEQE0,0\rightarrow\mathcal{O}^{r}\rightarrow E\rightarrow Q_{E}\rightarrow 0,

    where QEQ_{E} is a pure sheaf supported on a 1-dimensional subscheme of 2\mathbb{P}^{2}.

  2. (2)

    Let EBr(𝐯)E\in B^{r}(\mathbf{v}) be a general sheaf. If QEQ_{E} is supported on a smooth curve CC of degree c1(E)c_{1}(E), then QEQ_{E} is the pushforward of a line bundle on CC. Writing QE𝒪C(D)Q_{E}\simeq\mathcal{O}_{C}(D), we have

    degCDc12+3c12r1.\deg_{C}D\leq\frac{c_{1}^{2}+3c_{1}}{2}-r\leq-1.
  3. (3)

    Conversely, let 𝒪C(D)\mathcal{O}_{C}(D) be the pushforward of a line bundle on a smooth curve C2C\subseteq\mathbb{P}^{2}, where degCD12(deg(C)2+3deg(C))r1\deg_{C}D\leq\frac{1}{2}(\deg(C)^{2}+3\deg(C))-r\leq-1. Then a general extension sheaf

    0𝒪rE𝒪C(D)00\rightarrow\mathcal{O}^{r}\rightarrow E\rightarrow\mathcal{O}_{C}(D)\rightarrow 0

    is stable.

Proof.

(1) First, the proof of Theorem 4.1, Case I, implies that the evaluation map has full rank. It follows that QEQ_{E} is a torsion sheaf. Consider a negative twist

0𝒪(m)rE(m)QE(m)00\rightarrow\mathcal{O}(-m)^{r}\rightarrow E(-m)\rightarrow Q_{E}(-m)\rightarrow 0

of the above exact sequence, where m0m\gg 0. If QEQ_{E} is has torsion TQET\subseteq Q_{E} supported in codimension two, then T(m)TT(-m)\simeq T, so in particular T(m)T(-m) and QE(m)Q_{E}(-m) have global sections. The associated cohomology exact sequence

H0(E(m))H0(QE(m))H1(𝒪(m)r)H^{0}(E(-m))\rightarrow H^{0}(Q_{E}(-m))\rightarrow H^{1}(\mathcal{O}(-m)^{r})

has outer terms which vanish and an inner term which does not. This contradiction proves that QEQ_{E} is supported in pure codimension one.

(2) There is a filtration 0=FkQEFk1QEF1QEQE0=F_{k}Q_{E}\subseteq F_{k-1}Q_{E}\subseteq\cdots\subseteq F_{1}Q_{E}\subseteq Q_{E}, where each graded piece is the pushforward of a sheaf on CC [Dré08]. If FiQE0F_{i}Q_{E}\neq 0 then its support is CC, otherwise it would be a torsion subsheaf of QEQ_{E} supported on zero dimensional scheme, contradicting the purity of QEQ_{E} from (1). Now if F1QE0F_{1}Q_{E}\neq 0, then c1(F1QE)c1c_{1}(F_{1}Q_{E})\geq c_{1}, but QE/F1QEQ_{E}/F_{1}Q_{E} has support CC, whose first Chern class is c1\geq c_{1}, hence c1(QE)2c1c_{1}(Q_{E})\geq 2c_{1}, a contradiction.

For the degree, note that since rp(c1)r\geq p(c_{1}), we have

μ(E)c1p(c1)1/3\mu(E)\leq\frac{c_{1}}{p(c_{1})}\leq 1/3

since x/p(x)1/3x/p(x)\leq 1/3 for integers xx. Thus EE lies directly above the 𝒪{}^{\perp}\mathcal{O}-branch of the Drézet-Le Potier curve, which gives χ(E,𝒪)0\chi(E,\mathcal{O})\leq 0. Thus

χ(𝒪C(D),𝒪2)=χ(E,𝒪2)χ(𝒪2r,𝒪2)r.\chi(\mathcal{O}_{C}(D),\mathcal{O}_{\mathbb{P}^{2}})=\chi(E,\mathcal{O}_{\mathbb{P}^{2}})-\chi(\mathcal{O}_{\mathbb{P}^{2}}^{r},\mathcal{O}_{\mathbb{P}^{2}})\leq-r.

On the other hand,

χ(𝒪C(D),𝒪2)=χ(𝒪C,𝒪C(D3H))=degC(D3H)+1g(C)\chi(\mathcal{O}_{C}(D),\mathcal{O}_{\mathbb{P}^{2}})=\chi(\mathcal{O}_{C},\mathcal{O}_{C}(D-3H))=\deg_{C}(D-3H)+1-g(C)

so

degCD3c1+1c123c1+22r\deg_{C}D-3c_{1}+1-\frac{c_{1}^{2}-3c_{1}+2}{2}\leq-r

i.e., degCD12(c12+3c1)r\deg_{C}D\leq\frac{1}{2}(c_{1}^{2}+3c_{1})-r. The last inequality follows from the assumption rp(c1)r\geq p(c_{1}).

(3) Let FF be a possible destabilizing semistable quotient sheaf of EE. Since FF is torsion-free, there is no nonzero map from 𝒪C(D)F\mathcal{O}_{C}(D)\rightarrow F, hence the composition 𝒪rEF\mathcal{O}^{r}\rightarrow E\rightarrow F is nonzero. Since 𝒪rE\mathcal{O}^{r}\rightarrow E is of full rank, the composition 𝒪rEF\mathcal{O}^{r}\rightarrow E\rightarrow F is of full rank as well. The sheaves 𝒪r\mathcal{O}^{r} and FF are both semistable, which gives μ(F)0\mu(F)\geq 0. On the other hand, since FF has smaller rank than EE, c1(F)<c1(E)c_{1}(F)<c_{1}(E), hence the assumption on deg(D)\deg(D) implies FF also has r(F)p(c1(F))r(F)\geq p(c_{1}(F)). Therefore h0(F)r(F)h^{0}(F)\leq r(F), hence 𝒪rF\mathcal{O}^{r}\rightarrow F gives only r(F)r(F) sections, i.e., it factors through 𝒪r(F)F\mathcal{O}^{r(F)}\rightarrow F. By part (1), we have a short exact sequence

0𝒪r(F)FQF00\rightarrow\mathcal{O}^{r(F)}\rightarrow F\rightarrow Q_{F}\rightarrow 0

where QFQ_{F} is supported on a curve CC^{\prime}, and degCc1(F)<c1=degC\deg C^{\prime}\leq c_{1}(F)<c_{1}=\deg C. Now we have a diagram

0{0}𝒪r{\mathcal{O}^{r}}E{E}𝒪C(D){\mathcal{O}_{C}(D)}0{0}0{0}𝒪r(F){\mathcal{O}^{r(F)}}F{F}QF{Q_{F}}0{0}

where the vertical maps are surjective. The map 𝒪C(D)QF\mathcal{O}_{C}(D)\rightarrow Q_{F} induces an inclusion Supp QFC\text{Supp }Q_{F}\subseteq C. Supp QF\text{Supp }Q_{F} has strictly smaller degree than CC. Since CC is irreducible, this is impossible. We conclude QF=0Q_{F}=0, so FF is a slope 0 semistable sheaf with r(F)r(F) sections. Hence F=𝒪r(F)F=\mathcal{O}^{r(F)}, and the exact sequence defining EE is partially split.

We count dimensions to show such extensions form a closed locus in the associated extension space. If degD12(deg(C)2+3deg(C))r\deg D\leq\frac{1}{2}(\deg(C)^{2}+3\deg(C))-r, then ext1(𝒪C(D),𝒪2)r\text{ext}^{1}(\mathcal{O}_{C}(D),\mathcal{O}_{\mathbb{P}^{2}})\geq r. If E=𝒪2qEE=\mathcal{O}_{\mathbb{P}^{2}}^{q}\oplus E^{\prime} where EExt1(𝒪C(D),𝒪2rq)E^{\prime}\in\text{Ext}^{1}(\mathcal{O}_{C}(D),\mathcal{O}_{\mathbb{P}^{2}}^{r-q}), then the dimension of the locus of such EExt1(𝒪C(D),𝒪2r)E\in\text{Ext}^{1}(\mathcal{O}_{C}(D),\mathcal{O}_{\mathbb{P}^{2}}^{r}) is no larger than

dimG(q,r)+ext1(𝒪C(D),𝒪rq)=q(rq)+(rq)e,\dim G(q,r)+\text{ext}^{1}(\mathcal{O}_{C}(D),\mathcal{O}^{r-q})=q(r-q)+(r-q)e,

and ext1(𝒪C(D),𝒪2r)=re\text{ext}^{1}(\mathcal{O}_{C}(D),\mathcal{O}_{\mathbb{P}^{2}}^{r})=re. Now q(rq)+(rq)ere=(req)qq(r-q)+(r-q)e-re=(r-e-q)q, so we see that when ere\geq r and q>0q>0,

dimG(q,r)+ext1(𝒪C(D),𝒪2rq)<ext1(𝒪C(D),𝒪2r).\dim G(q,r)+\text{ext}^{1}(\mathcal{O}_{C}(D),\mathcal{O}_{\mathbb{P}^{2}}^{r-q})<\text{ext}^{1}(\mathcal{O}_{C}(D),\mathcal{O}_{\mathbb{P}^{2}}^{r}).

The lemma follows. ∎

Theorem 4.6.

Suppose that 𝐯K(2)\mathbf{v}\in K(\mathbb{P}^{2}) is stable and ch0(𝐯)p(ch1(𝐯))\text{ch}_{0}(\mathbf{v})\geq p(\text{ch}_{1}(\mathbf{v})), ch1(𝐯)>0\text{ch}_{1}(\mathbf{v})>0. Then Br(𝐯)B^{r}(\mathbf{v}) is nonempty and contains a component of the expected dimension.

Proof.

By Lemma 4.5 (1), we get a morphism ψ:Br(𝐯)H0(𝒪2(ch1(𝐯)))=V\psi:B^{r}(\mathbf{v})\rightarrow\mathbb{P}H^{0}(\mathcal{O}_{\mathbb{P}^{2}}(\text{ch}_{1}(\mathbf{v})))=\mathbb{P}V, by sending EM(𝐯)E\in M(\mathbf{v}) with rr global sections s1,,srH0(E)s_{1},...,s_{r}\in H^{0}(E) to s1srVs_{1}\wedge\cdots\wedge s_{r}\in\mathbb{P}V, and by Lemma 4.5 (3), ψ\psi is surjective. Denoting by U=VΓU=\mathbb{P}V\smallsetminus\Gamma the locus of smooth curves, we see that ψ\psi factors through the relative Picard scheme Pic𝒞/Ud\text{Pic}^{d}_{\mathcal{C}/U} over the universal curve 𝒞U\mathcal{C}\rightarrow U, where d=deg(D)d=\deg(D). The induced map ϕ:Br(𝐯)Pic𝒞/Ud\phi:B^{r}(\mathbf{v})\rightarrow\text{Pic}^{d}_{\mathcal{C}/U} is surjective, and since EE uniquely determines QE=𝒪C(D)Q_{E}=\mathcal{O}_{C}(D) its fibers are quotients of open subsets of the extension spaces Ext1(𝒪C(D),𝒪2r)\text{Ext}^{1}(\mathcal{O}_{C}(D),\mathcal{O}_{\mathbb{P}^{2}}^{r}). It follows in particular that ψ1(U)Br(𝐯)\psi^{-1}(U)\subseteq B^{r}(\mathbf{v}) is irreducible and open, so its closure forms a nonempty irreducible component.

We now compute its dimension. Write r=ch0(𝐯)r=\text{ch}_{0}(\mathbf{v}), c1=ch1(𝐯)c_{1}=\text{ch}_{1}(\mathbf{v}), μ=c1/r\mu=c_{1}/r, and Δ(𝐯)=Δ0+k/r\Delta(\mathbf{v})=\Delta_{0}+k/r where Δ0\Delta_{0} is the minimal discriminant of a stable sheaf with μ=μ(𝐯)\mu=\mu(\mathbf{v}) and r=ch0(𝐯)r=\text{ch}_{0}(\mathbf{v}). Note in particular that 𝐯\mathbf{v} lies directly above the 𝒪{}^{\perp}\mathcal{O}-branch of the Drézet-Le Potier curve, which is given by

Δ=p(μ)=12μ232μ+1.\Delta=p(-\mu)=\frac{1}{2}\mu^{2}-\frac{3}{2}\mu+1.

In particular, the characters (μ,δ(μ))(\mu,\delta(\mu)) are integral, so δ(μ)=Δ0\delta(\mu)=\Delta_{0} for these slopes. Then M(𝐯)M(\mathbf{v}) is irreducible of dimension

dimM(𝐯)=r2(2(Δ0+k/r)1)+1=1+c123c1r+r2+2kr.\dim M(\mathbf{v})=r^{2}(2(\Delta_{0}+k/r)-1)+1=1+c_{1}^{2}-3c_{1}r+r^{2}+2kr.

The expected codimension of the Brill-Noether locus Bm(𝐯)B^{m}(\mathbf{v}) is m(mχ(𝐯))m(m-\chi(\mathbf{v})). Now

χ(𝐯)=r(p(μ)Δ0k/r)=3c1k,\chi(\mathbf{v})=r\left(p\left(\mu\right)-\Delta_{0}-k/r\right)=3c_{1}-k,

so the expected codimension of Br(𝐯)B^{r}(\mathbf{v}) is r(r3c1+k)r(r-3c_{1}+k), and its expected dimension is

expdimBr(𝐯)=(1+c123c1r+r2+2kr)r(r3c1+k)=1+c12+kr.\text{expdim}B^{r}(\mathbf{v})=(1+c_{1}^{2}-3c_{1}r+r^{2}+2kr)-r(r-3c_{1}+k)=1+c_{1}^{2}+kr.

The dimension of the locus of interest is equal to dimϕ1Picd(𝒞/U)\dim\phi^{-1}\text{Pic}^{d}(\mathcal{C}/U). The general EExt1(𝒪C(D),𝒪r)E\in\text{Ext}^{1}(\mathcal{O}_{C}(D),\mathcal{O}^{r}) admits precisely one map 𝒪rE\mathcal{O}^{r}\rightarrow E, and automorphisms of 𝒪r\mathcal{O}^{r} determine distinct extension classes but isomorphic extension sheaves. Thus

dimϕ1Pic𝒞/Ud=dimU+g(C)+ext1(𝒪C(D),𝒪r)r2+1.\dim\phi^{-1}\text{Pic}^{d}_{\mathcal{C}/U}=\dim U+g(C)+\text{ext}^{1}(\mathcal{O}_{C}(D),\mathcal{O}^{r})-r^{2}+1.

We have ext2(𝒪C(D),𝒪)=h0(𝒪C(D3H))=0\text{ext}^{2}(\mathcal{O}_{C}(D),\mathcal{O})=h^{0}(\mathcal{O}_{C}(D-3H))=0, and clearly hom(𝒪C(D),𝒪)=0\hom(\mathcal{O}_{C}(D),\mathcal{O})=0, so ext1(𝒪C(D),𝒪)=χ(𝒪C(D),𝒪)\text{ext}^{1}(\mathcal{O}_{C}(D),\mathcal{O})=-\chi(\mathcal{O}_{C}(D),\mathcal{O}). From the exact sequence

0𝒪rE𝒪C(D)00\rightarrow\mathcal{O}^{r}\rightarrow E\rightarrow\mathcal{O}_{C}(D)\rightarrow 0

we obtain

χ(𝒪C(D),𝒪)=χ(E,𝒪)χ(𝒪r,𝒪)=kr.\chi(\mathcal{O}_{C}(D),\mathcal{O})=\chi(E,\mathcal{O})-\chi(\mathcal{O}^{r},\mathcal{O})=-k-r.

Hence

dimϕ1(𝒪C(D))=r(k+r)r2=kr.\dim\phi^{-1}(\mathcal{O}_{C}(D))=r(k+r)-r^{2}=kr.

As dimU=dimH0(𝒪2(c1))=12(c12+3c1)\dim U=\dim\mathbb{P}H^{0}(\mathcal{O}_{\mathbb{P}^{2}}(c_{1}))=\frac{1}{2}(c_{1}^{2}+3c_{1}), we have

dimPic𝒞/Ud=dimU+g(C)=(c12+3c1)/2+(c123c1+2)/2=c12+1,\dim\text{Pic}^{d}_{\mathcal{C}/U}=\dim U+g(C)=(c_{1}^{2}+3c_{1})/2+(c_{1}^{2}-3c_{1}+2)/2=c_{1}^{2}+1,

hence

dimψ1(U)=c12+1+kr,\dim\psi^{-1}(U)=c_{1}^{2}+1+kr,

which equals the expected dimension. ∎

When we have instead ch0(𝐯)<p(ch1(𝐯))\text{ch}_{0}(\mathbf{v})<p(\text{ch}_{1}(\mathbf{v})), we prove a weaker statement: that the Brill-Noether locus Br(𝐯)B^{r}(\mathbf{v}) is nonempty.

Note that the condition χ(𝐯)=ch0(𝐯)\chi(\mathbf{v})=\text{ch}_{0}(\mathbf{v}) can be expressed via Riemann-Roch as the polynomial condition

Δ(𝐯)=p(μ(𝐯))1=12μ(𝐯)2+32μ(𝐯).\Delta(\mathbf{v})=p(\mu(\mathbf{v}))-1=\frac{1}{2}\mu(\mathbf{v})^{2}+\frac{3}{2}\mu(\mathbf{v}).

This defines a parabola ξr\xi_{r} in the (μ,Δ)(\mu,\Delta)-plane. (ξr\xi_{r} does not depend on the rank.)

In the region of the (μ,Δ)(\mu,\Delta)-plane with μ>0\mu>0, the parabola ξr\xi_{r} meets the Drézet-Le Potier curve only at the 𝒪{}^{\perp}\mathcal{O} branch, at the point (μ,Δ)=(1/3,5/9)(\mu,\Delta)=(1/3,5/9). The Chern characters 𝐯K(2)\mathbf{v}\in K(\mathbb{P}^{2}) lying on ξr\xi_{r} satisfy χ(𝐯)=ch0(𝐯)\chi(\mathbf{v})=\text{ch}_{0}(\mathbf{v}), and a general sheaf EM(𝐯)E\in M(\mathbf{v}) has h0(E)=r(E)h^{0}(E)=r(E).

Theorem 4.7.

Let 𝐯K(2)\mathbf{v}\in K(\mathbb{P}^{2}) be a stable Chern character with μ(𝐯)>0\mu(\mathbf{v})>0. The Brill-Noether locus Br(𝐯)B^{r}(\mathbf{v}) is nonempty.

Proof.

When r=1r=1, c11c_{1}\geq 1, we may take IZ(c1)I_{Z}(c_{1}) where ZZ lies on a curve of degree c1c_{1}. When c1=1c_{1}=1, by Theorem 4.6 the theorem is true for r3r\geq 3. Hence in the following we assume r>1r>1, c1>1c_{1}>1 or c1=1,r=2c_{1}=1,r=2. We separate into two cases.

Case I. In the first case we assume 0<μ(𝐯)1/30<\mu(\mathbf{v})\leq 1/3; in particular, 𝐯\mathbf{v} lies above the 𝒪{}^{\perp}\mathcal{O}-branch of the Drézet-Le Potier curve. We induct on the rank for a stronger statement: there exists EM(𝐯)E\in M(\mathbf{v}) such that h0(E)=ch0(E)h^{0}(E)=\text{ch}_{0}(E). When the rank is 1 or 2 the statement is vacuous. When r=3r=3, c1=1c_{1}=1, by Theorem 4.6 the theorem is true.

Let 𝐯\mathbf{v}^{\prime} be the associated extremal character, and 𝐯′′=𝐯𝐯\mathbf{v}^{\prime\prime}=\mathbf{v}-\mathbf{v}^{\prime} the quotient character. Then 0μ(𝐯)μ(𝐯)1/30\leq\mu(\mathbf{v}^{\prime})\leq\mu(\mathbf{v})\leq 1/3. We have μ(𝐯)=0\mu(\mathbf{v}^{\prime})=0 if and only if c1=1c_{1}=1, where the only possiblity for the theorem to be false is r=2r=2, which is not considered in this case, hence 0<μ(𝐯)1/30<\mu(\mathbf{v}^{\prime})\leq 1/3.

We also have 0<μ(𝐯′′)1/30<\mu(\mathbf{v}^{\prime\prime})\leq 1/3. To see this, first consider the case where μ(𝐯)=μ(𝐯)\mu(\mathbf{v}^{\prime})=\mu(\mathbf{v}), which occurs if and only if ch0(𝐯)\text{ch}_{0}(\mathbf{v}) and ch1(𝐯)\text{ch}_{1}(\mathbf{v}) are not coprime. In this case, μ(𝐯′′)=μ(𝐯)1/3\mu(\mathbf{v}^{\prime\prime})=\mu(\mathbf{v})\leq 1/3. Otherwise, ch0(𝐯)\text{ch}_{0}(\mathbf{v}) and ch1(𝐯)\text{ch}_{1}(\mathbf{v}) are coprime. If μ(𝐯)=1/3\mu(\mathbf{v})=1/3, then ch0(𝐯)=3\text{ch}_{0}(\mathbf{v})=3 and ch1(𝐯)=1\text{ch}_{1}(\mathbf{v})=1, which has already been addressed. In the remaining cases, ch0(𝐯)4\text{ch}_{0}(\mathbf{v})\geq 4. In this case μ(𝐯′′)\mu(\mathbf{v}^{\prime\prime}) is the right Farey neighbor of μ(𝐯)\mu(\mathbf{v}), which is no greater than 1/31/3.

An immediate consequence is that the extremal decomposition exists in this case. Denote the Drézet-Le Potier curve by Δ=δ(μ)\Delta=\delta(\mu). If μ(𝐯)<μ(𝐯)\mu(\mathbf{v}^{\prime})<\mu(\mathbf{v}), then Δ(𝐯)δ(μ(𝐯))\Delta(\mathbf{v}^{\prime})\leq\delta(\mu(\mathbf{v}^{\prime})) since (r(𝐯),μ(𝐯),δ(μ(𝐯)))(r(\mathbf{v}^{\prime}),\mu(\mathbf{v}^{\prime}),\delta(\mu(\mathbf{v}^{\prime}))) is integral (𝐯\mathbf{v}^{\prime} is either on the Drézet-Le Potier curve or semi-exceptional). Then χ(𝐯,𝒪)0\chi(\mathbf{v}^{\prime},\mathcal{O})\geq 0 (with equality only if 𝐯=ch(𝒪k)\mathbf{v}^{\prime}=\text{ch}(\mathcal{O}^{k})), and χ(𝐯,𝒪)0\chi(\mathbf{v},\mathcal{O})\leq 0, hence χ(𝐯′′,𝒪)=χ(𝐯,𝒪)χ(𝐯,𝒪)0\chi(\mathbf{v}^{\prime\prime},\mathcal{O})=\chi(\mathbf{v},\mathcal{O})-\chi(\mathbf{v}^{\prime},\mathcal{O})\leq 0, so 𝐯′′\mathbf{v}^{\prime\prime} lies on or above the Drézet-Le Potier curve since 0<μ(𝐯′′)1/30<\mu(\mathbf{v}^{\prime\prime})\leq 1/3. If μ(𝐯)=μ(𝐯)\mu(\mathbf{v}^{\prime})=\mu(\mathbf{v}), we may write 𝐯=(mr0,μ,δ(μ)+kmr)\mathbf{v}=(mr_{0},\mu,\delta(\mu)+\frac{k}{mr}) where ch1(𝐯)=md0\text{ch}_{1}(\mathbf{v})=md_{0}, μ=d/r0\mu=d/r_{0}, and r0,d0r_{0},d_{0} are coprime. Let 𝐯0:=(r0,μ,δ(μ)+kr)\mathbf{v}^{\prime}_{0}:=(r_{0},\mu,\delta(\mu)+\frac{k}{r}), 𝐯0′′:=((m1)r0,μ,δ(μ))\mathbf{v}^{\prime\prime}_{0}:=((m-1)r_{0},\mu,\delta(\mu)), one may check that 𝐯0\mathbf{v}^{\prime}_{0} and 𝐯0′′\mathbf{v}^{\prime\prime}_{0} are indeed integral. Then 𝐯=𝐯0+𝐯0′′\mathbf{v}=\mathbf{v}^{\prime}_{0}+\mathbf{v}^{\prime\prime}_{0}. Now clearly Δ(𝐯0)Δ(𝐯)\Delta(\mathbf{v}^{\prime}_{0})\geq\Delta(\mathbf{v}), hence by definition of extremal character, Δ(𝐯)Δ(𝐯0)\Delta(\mathbf{v}^{\prime})\leq\Delta(\mathbf{v}^{\prime}_{0}), hence Δ(𝐯′′)Δ(𝐯0′′)=δ(μ)\Delta(\mathbf{v}^{\prime\prime})\geq\Delta(\mathbf{v}^{\prime\prime}_{0})=\delta(\mu), so 𝐯′′\mathbf{v}^{\prime\prime} also lies on or above the Drézet-Le Potier curve.

By assumption r2r\geq 2, both 𝐯\mathbf{v}^{\prime} and 𝐯′′\mathbf{v}^{\prime\prime} have smaller rank than 𝐯\mathbf{v}, thus by induction we can assume that there are EM(𝐯)E^{\prime}\in M(\mathbf{v}^{\prime}) and E′′M(𝐯′′)E^{\prime\prime}\in M(\mathbf{v}^{\prime\prime}) satisfying h0(E)=r(E)h^{0}(E^{\prime})=r(E^{\prime}) and h0(E′′)=r(E′′)h^{0}(E^{\prime\prime})=r(E^{\prime\prime}), and by Proposition 3.3 any nontrivial extension sheaf

0EEE′′00\rightarrow E^{\prime}\rightarrow E\rightarrow E^{\prime\prime}\rightarrow 0

is semistable. To show h0(E)=r(E)=r(E)+r(E′′)h^{0}(E)=r(E)=r(E^{\prime})+r(E^{\prime\prime}), it is enough to show that we can find such an extension such that the associated connecting homomorphism H0(E′′)H1(E)H^{0}(E^{\prime\prime})\rightarrow H^{1}(E^{\prime}) vanishes.

Indeed, the association eδee\mapsto\delta_{e} determines a linear map

Ext1(E′′,E)Hom(H0(E′′),H1(E)).\text{Ext}^{1}(E^{\prime\prime},E^{\prime})\rightarrow\text{Hom}(H^{0}(E^{\prime\prime}),H^{1}(E^{\prime})).

We need to show that it has nonzero kernel; in fact, the dimension of the source exceeds the dimension of the target. By stability and Serre duality, ext2(E′′,E)=hom(E,E′′(3))=0\text{ext}^{2}(E^{\prime\prime},E^{\prime})=\hom(E^{\prime},E^{\prime\prime}(-3))=0, so

ext1(E′′,E)χ(E′′,E)=rr′′(p(μμ′′)ΔΔ′′),\text{ext}^{1}(E^{\prime\prime},E^{\prime})\geq-\chi(E^{\prime\prime},E^{\prime})=-r^{\prime}r^{\prime\prime}(p(\mu^{\prime}-\mu^{\prime\prime})-\Delta^{\prime}-\Delta^{\prime\prime}),

and we know already that h0(E′′)=r′′h^{0}(E^{\prime\prime})=r^{\prime\prime}, and

h1(E)=h0(E)χ(E)=rr(p(μ)Δ).h^{1}(E^{\prime})=h^{0}(E^{\prime})-\chi(E^{\prime})=r^{\prime}-r^{\prime}(p(\mu^{\prime})-\Delta^{\prime}).

Since 𝐯′′\mathbf{v}^{\prime\prime} lies above the 𝒪{}^{\perp}\mathcal{O}-branch of the Drézet-Le Potier curve, χ(E′′,𝒪)0\chi(E^{\prime\prime},\mathcal{O})\leq 0. It follows in particular from Riemann-Roch that Δ′′p(μ′′)\Delta^{\prime\prime}\geq p(-\mu^{\prime\prime}). We then have:

ext1(E′′,E)hom(H0(E′′),H1(E))\displaystyle\text{ext}^{1}(E^{\prime\prime},E)-\hom(H^{0}(E^{\prime\prime}),H^{1}(E^{\prime})) rr′′(p(μμ′′)ΔΔ′′)\displaystyle\geq-r^{\prime}r^{\prime\prime}(p(\mu^{\prime}-\mu^{\prime\prime})-\Delta^{\prime}-\Delta^{\prime\prime})
r′′(rr(p(μ)Δ))\displaystyle\qquad-r^{\prime\prime}(r^{\prime}-r^{\prime}(p(\mu^{\prime})-\Delta^{\prime}))
=rr′′(p(μμ′′)p(μ)Δ′′+1)\displaystyle=-r^{\prime}r^{\prime\prime}(p(\mu^{\prime}-\mu^{\prime\prime})-p(\mu^{\prime})-\Delta^{\prime\prime}+1)
rr′′(p(μμ′′)p(μ)p(μ′′)+1)\displaystyle\geq-r^{\prime}r^{\prime\prime}(p(\mu^{\prime}-\mu^{\prime\prime})-p(\mu^{\prime})-p(-\mu^{\prime\prime})+1)
=rr′′μμ′′=c1c1′′>0\displaystyle=r^{\prime}r^{\prime\prime}\mu^{\prime}\mu^{\prime\prime}=c_{1}^{\prime}c_{1}^{\prime\prime}>0

as desired. We conclude there is EM(𝐯)E\in M(\mathbf{v}) with

h0(E)=h0(E)+h0(E′′)=r(E)+r(E′′)=r(E),h^{0}(E)=h^{0}(E^{\prime})+h^{0}(E^{\prime\prime})=r(E^{\prime})+r(E^{\prime\prime})=r(E),

i.e., Br(𝐯)B^{r}(\mathbf{v})\neq\emptyset.

Case II. In the second case, we assume μ(𝐯)>1/3\mu(\mathbf{v})>1/3. Let 𝐯K(2)\mathbf{v}\in K(\mathbb{P}^{2}) be a stable Chern character of slope μ=μ(𝐯)\mu=\mu(\mathbf{v}) and rank r=ch0(𝐯)r=\text{ch}_{0}(\mathbf{v}), and let Δ0\Delta_{0} be the minimum discriminant of a stable character of slope μ\mu and rank rr. Then we may write Δ(𝐯)=Δ(𝐯0)+k/r\Delta(\mathbf{v})=\Delta(\mathbf{v}_{0})+k/r for some k0k\geq 0. For every rational number μ=c1/r\mu^{\prime}=c_{1}^{\prime}/r^{\prime}, the character 𝐯=(r,μ,ξr(μ))\mathbf{v}^{\prime}=(r^{\prime},\mu^{\prime},\xi_{r}(\mu)) is integral since χ(𝐯)=r\chi(\mathbf{v}^{\prime})=r^{\prime} is an integer.

Refer to caption
Figure 2. The regions verifying ξr(μ)δ(μ)>1/r\xi_{r}(\mu)-\delta(\mu)>1/r.

We claim that the minimal-discriminant character 𝐯0=(μ(𝐯),Δ0)\mathbf{v}_{0}=(\mu(\mathbf{v}),\Delta_{0}) lies below the parabola ξr\xi_{r} in the (μ,Δ)(\mu,\Delta)-plane, so χ(𝐯0)>r\chi(\mathbf{v}_{0})>r. To check this we will show that ξr(μ)δ(μ)>1/r\xi_{r}(\mu)-\delta(\mu)>1/r, so that any character with discriminant larger than ξr(μ)\xi_{r}(\mu) is obtained from one below ξr\xi_{r} by an elementary modification, so is not minimal. We divide the (μ,Δ)(\mu,\Delta)-plane into smaller regions, see Figure 2: first, it is enough to check this for 1/3<μ<1/21/3<\mu<1/2, since the Drézet-Le Potier curve is symmetric along the line μ=1/2\mu=1/2. For slopes μ\mu lying directly above the 𝒪{}^{\perp}\mathcal{O}-branch of the Drézet-Le Potier curve, one can easily check that the points (μ,δ(μ))(\mu,\delta(\mu)) are integral, so the claim is true here. This branch terminates at μ=(35)/2\mu=(3-\sqrt{5})/2 (see [CHW17, §2.3]); the branch T2(1)T_{\mathbb{P}^{2}}(-1)^{\perp} to the left of the line μ=1/2\mu=1/2 terminates at μ=1/2(38)/2\mu=1/2-(3-\sqrt{8})/2. In this region we have the coarse bounds

ξr(μ)δ(μ)\displaystyle\xi_{r}(\mu)-\delta(\mu) ξr(μ)(peak of DLP curve above E2/5)\displaystyle\geq\xi_{r}(\mu)-(\text{peak of DLP curve above }E_{2/5})
=ξr(μ)1325ξr(352)1325.12589\displaystyle=\xi_{r}(\mu)-\frac{13}{25}\geq\xi_{r}\left(\frac{3-\sqrt{5}}{2}\right)-\frac{13}{25}\approx.12589

Thus if ξr(μ)δ(μ)1/r\xi_{r}(\mu)-\delta(\mu)\leq 1/r then r<9r<9, and the few slopes μ=c1/r\mu=c_{1}/r with r<9r<9 in this region can be easily checked to have minimal discriminants lying below the parabola ξr\xi_{r}.

Now for slopes μ\mu lying directly above the T2(1)T_{\mathbb{P}^{2}}(-1)^{\perp}-branch, we have

ξr(μ)δ(μ)=ξr(μ)(12(μ12)2+32(μ12)+1)=μ2\xi_{r}(\mu)-\delta(\mu)=\xi_{r}(\mu)-\left(\frac{1}{2}\left(\mu-\frac{1}{2}\right)^{2}+\frac{3}{2}\left(\mu-\frac{1}{2}\right)+1\right)=\frac{\mu}{2}

and μ/21/r\mu/2\leq 1/r only when c12c_{1}\leq 2. All statements are clear when c1=1c_{1}=1, and when r5r\geq 5, μ=2/r\mu=2/r is to the left of this region, and for smaller ranks the minimal discriminant is easily computed to be below ξr(μ)\xi_{r}(\mu).

By generic slope-stability ([DLP85, Corollaire 4.12]), a general member of M(𝐯0)M(\mathbf{v}_{0}) is slope-stable. Choose such a E0M(𝐯0)E_{0}\in M(\mathbf{v}_{0}) with rr independent sections s1,,srH0(E0)s_{1},...,s_{r}\in H^{0}(E_{0}); choose a codimension two subscheme ZZ of length kk. Then the elementary modification EE defined by the short exact sequence

0EE0𝒪Z00\rightarrow E\rightarrow E_{0}\rightarrow\mathcal{O}_{Z}\rightarrow 0

has Chern character 𝐯\mathbf{v}. Since E0E_{0} is slope-stable, so is EE (see Section 2.2). Furthermore we can choose the map E0𝒪ZE_{0}\rightarrow\mathcal{O}_{Z} such that the map H0(E0)H0(𝒪Z)H^{0}(E_{0})\rightarrow H^{0}(\mathcal{O}_{Z}) vanishes, since the dependency locus of s1,,srs_{1},\cdots,s_{r} has dimension at least 1. So h0(E)rh^{0}(E)\geq r for all EE constructed in this way. ∎

5. Irreducibility & reducibility of Brill-Noether loci

In this section we address the reducibility of the Brill-Noether loci Bk(𝐯)B^{k}(\mathbf{v}). We show that when ch1(𝐯)=1\text{ch}_{1}(\mathbf{v})=1, all Brill-Noether loci are irreducible and of the expected dimension. When ch1(𝐯)>1\text{ch}_{1}(\mathbf{v})>1, we give examples where the Brill-Noether loci are reducible, and describe their components.

Theorem 5.1.

Suppose 𝐯K(2)\mathbf{v}\in K(\mathbb{P}^{2}) is a stable Chern character with ch1(𝐯)=1\text{ch}_{1}(\mathbf{v})=1. Then all of the Brill-Noether loci Bk(𝐯)B^{k}(\mathbf{v}) are irreducible and of the expected dimension.

Each Brill-Noether locus Bk(𝐯)B^{k}(\mathbf{v}) is the quotient of a projective bundle over some M(𝐰)M(\mathbf{w}) by a free action of a projective linear group, where ch1(𝐰)=1\text{ch}_{1}(\mathbf{w})=1 as well.

Proof.

Choose EM(𝐯)E\in M(\mathbf{v}) and set r=r(E)r=r(E). Consider the image sheaf 0FE0\neq F\subseteq E of the evaluation map on global sections for EE. The proof of Theorem 4.1, Case I implies that if r3r\geq 3, then h0(E)rh^{0}(E)\leq r, and that the evaluation map in this case is a sheaf-theoretic embedding. (In rank 2, see Remark 4.4. The only sheaf whose evaluation map on global sections is not an embedding is T2(1)T_{\mathbb{P}^{2}}(-1), and here the statements are vacuous.) Thus every EM(𝐯)E\in M(\mathbf{v}) is given by an extension

(3) 0𝒪rsEE00\rightarrow\mathcal{O}^{r-s}\rightarrow E\rightarrow E^{\prime}\rightarrow 0

where h0(E)=rsh^{0}(E)=r-s, s0s\geq 0, and EE^{\prime} has c1(E)=c1(E)=1c_{1}(E^{\prime})=c_{1}(E)=1 and r(E)=sr(E^{\prime})=s. In fact when s0s\geq 0 all such EE^{\prime} are semistable, and when s>0s>0, torsion-free. To check torsion-freeness, consider the torsion part TT of EE^{\prime}, which sits in an exact sequence

0TEE/T0.0\rightarrow T\rightarrow E^{\prime}\rightarrow E^{\prime}/T\rightarrow 0.

Then arguing as in the proof of Lemma 4.5 (1), we see that TT is supported in dimension 1. Then c1(E/T)=c1(E)c1(T)0c_{1}(E^{\prime}/T)=c_{1}(E^{\prime})-c_{1}(T)\leq 0, and the quotient EE/TE\rightarrow E^{\prime}/T destabilizes EE. Now for stability, if s1s\geq 1 then c1(E)c_{1}(E^{\prime}) and r(E)=sr(E^{\prime})=s are coprime, any destablizing quotient QQ of EE^{\prime} would have μ(Q)<μ(E)\mu(Q)<\mu(E^{\prime}), but since c1(E)=1c_{1}(E^{\prime})=1, we have μ(Q)0\mu(Q)\leq 0, so QQ would be a destablizing quotient of EE. When s=0s=0, by Lemma 4.5 (2), EE^{\prime} is of the form 𝒪L(a)\mathcal{O}_{L}(a) for some aa\in\mathbb{Z} and L2L\subseteq\mathbb{P}^{2} a line, hence stable. Set 𝐰=ch(E)=ch(E)(rs)ch(𝒪)\mathbf{w}=\text{ch}(E^{\prime})=\text{ch}(E)-(r-s)\text{ch}(\mathcal{O}).

We form extension parametrizations ϕrs:rsM(𝐯)\phi_{r-s}:\mathbb{P}_{r-s}\dashrightarrow M(\mathbf{v}) whose image lies in Brsχ(𝐯)(𝐯)B_{r-s-\chi(\mathbf{v})}(\mathbf{v}) as follows. The expected extensions (3) are determined by classes in Ext1(E,𝒪rs)\text{Ext}^{1}(E^{\prime},\mathcal{O}^{r-s}). Since Hom(E,𝒪rs)=Ext2(E,𝒪rs)=0\text{Hom}(E^{\prime},\mathcal{O}^{r-s})=\text{Ext}^{2}(E^{\prime},\mathcal{O}^{r-s})=0, ext1(E,𝒪rs)=χ(E,𝒪rs)\text{ext}^{1}(E^{\prime},\mathcal{O}^{r-s})=-\chi(E^{\prime},\mathcal{O}^{r-s}) is constant, so the extension parametrization is defined over the whole moduli space M(𝐰)M(\mathbf{w}). rs\mathbb{P}_{r-s} maps rationally to Brsχ(𝐯)(𝐯)M(𝐯)B_{r-s-\chi(\mathbf{v})}(\mathbf{v})\subseteq M(\mathbf{v}).

When s>0s>0, Proposition 3.3 implies that for any FM(𝐰)F^{\prime}\in M(\mathbf{w}) any nonzero extension 0[F]Ext1(F,𝒪h0(E))0\neq[F]\in\text{Ext}^{1}(F^{\prime},\mathcal{O}^{h^{0}(E)}) is stable. When s=0s=0, Lemma 4.5 (3) implies the same statement. In particular, each of the extension parametrizations ϕk:kM(𝐯)\phi_{k}:\mathbb{P}_{k}\rightarrow M(\mathbf{v}) is a morphism. Since each EE^{\prime} as above is stable, it follows that ϕk\phi_{k} surjects onto Bk(𝐯)B^{k}(\mathbf{v}). The moduli spaces M(𝐯)M(\mathbf{v}^{\prime}) are irreducible, so each k\mathbb{P}_{k} is as well. It follows that so is each Bk(𝐯)B^{k}(\mathbf{v}). We now show that they are of the expected dimension.

We will need to know the other numerical invariants of EE and EE^{\prime}. They are determined as follows. Set ch2(E)=a1/2=ch2(E)\text{ch}_{2}(E)=a-1/2=\text{ch}_{2}(E^{\prime}) (see Remark 5.2). This gives

Δ(E)=12r2ar+12r,χ(E)=r(p(1/r)Δ(E))=r+a+1.\Delta(E)=\frac{1}{2r^{2}}-\frac{a}{r}+\frac{1}{2r},\quad\chi(E)=r(p(1/r)-\Delta(E))=r+a+1.

Similarly,

Δ(E)=12s2as+12s.\Delta(E^{\prime})=\frac{1}{2s^{2}}-\frac{a}{s}+\frac{1}{2s}.

The dimension of \mathbb{P} is

dimϕ()=dimM(𝐰)+ext1(E,𝒪rs)1.\dim\phi(\mathbb{P})=\dim M(\mathbf{w})+\text{ext}^{1}(E^{\prime},\mathcal{O}^{r-s})-1.

The latter quantity is, by Serre duality, Theorem 2.1, and Riemann-Roch,

ext1(E,𝒪rs)\displaystyle\text{ext}^{1}(E^{\prime},\mathcal{O}^{r-s}) =(rs)h1(E(3))\displaystyle=(r-s)\cdot h^{1}(E^{\prime}(-3))
=(rs)χ(E(3))\displaystyle=-(r-s)\cdot\chi(E^{\prime}(-3))
=(rs)s(P(1/s3)Δ(E))\displaystyle=-(r-s)\cdot s(P(1/s-3)-\Delta(E^{\prime}))
=(rs)(s+a2).\displaystyle=-(r-s)\cdot(s+a-2).

The dimensions of the moduli spaces are

dimM(𝐯)=r2(2Δ(E)1)+1=r2+(12a)r+2\dim M(\mathbf{v})=r^{2}(2\Delta(E)-1)+1=-r^{2}+(1-2a)r+2

and

dimM(𝐰)=s2(2Δ(E)1)+1=s2+(12a)s+2.\dim M(\mathbf{w})=s^{2}(2\Delta(E^{\prime})-1)+1=-s^{2}+(1-2a)s+2.

The fibers of ϕ\phi are of dimension Aut(𝒪rs)=(rs)21\text{Aut}(\mathcal{O}^{r-s})=(r-s)^{2}-1, so the dimension of the image of ϕ\phi is

dimϕ()\displaystyle\dim\phi(\mathbb{P}) =s2(2Δ(E)1)+1(rs)(s+a2)(rs)2+1\displaystyle=s^{2}(2\Delta(E^{\prime})-1)+1-(r-s)\cdot(s+a-2)-(r-s)^{2}+1
=s2+(12a)s+2(rs)(s+a2)1(rs)2+1\displaystyle=-s^{2}+(1-2a)s+2-(r-s)\cdot(s+a-2)-1-(r-s)^{2}+1
=r2+(sa+2)rs2(a+1)s+2.\displaystyle=-r^{2}+(s-a+2)r-s^{2}-(a+1)s+2.

This is the actual dimension of the Brill-Noether locus; we now check that the expected dimension of the Brill-Noether loci equals the actual dimension.

Set m=rsχ(𝐯)m=r-s-\chi(\mathbf{v}); the expected codimension of the Brill-Noether locus Bm(𝐯)B_{m}(\mathbf{v}) is

expcodimm(𝐯)\displaystyle\text{expcodim}_{m}(\mathbf{v}) =(rs)m\displaystyle=(r-s)\cdot m
=(rs)(rsχ(𝐯))\displaystyle=-(r-s)\cdot(r-s-\chi(\mathbf{v}))
=(sr)(sa1)\displaystyle=(s-r)\cdot(s-a-1)
=s2+(1r+a)sarr.\displaystyle=s^{2}+(1-r+a)s-ar-r.

The expected dimension is then

expdimm(𝐯)\displaystyle\text{expdim}_{m}(\mathbf{v}) =dimM(𝐯)(s2+(1r+a)sarr)\displaystyle=\dim M(\mathbf{v})-(s^{2}+(1-r+a)s-ar-r)
=r2+(12a)r+2s2(1r+a)s+ar+r\displaystyle=-r^{2}+(1-2a)r+2-s^{2}-(1-r+a)s+ar+r
=r2+(sa+2)rs2(a+1)s+2\displaystyle=-r^{2}+(s-a+2)r-s^{2}-(a+1)s+2

as required.

For the latter statement, we consider the map kBk(𝐯)\mathbb{P}_{k}\rightarrow B^{k}(\mathbf{v}). By Lemma 6.3 in [CH18], all fibers of this map are GLk\mathbb{P}GL_{k}. ∎

Remark 5.2.

When h0(E)=rh^{0}(E)=r, the associated extension is of the form

0𝒪rE𝒪L(a)00\rightarrow\mathcal{O}^{r}\rightarrow E\rightarrow\mathcal{O}_{L}(a)\rightarrow 0

for L2L\subseteq\mathbb{P}^{2} a line and a0a\leq 0. In this case ch2(E)=a1/2\text{ch}_{2}(E)=a-1/2. We regard aa as an invariant of EE for other values of h0(E)h^{0}(E) via this formula.

5.1. Reducible Brill-Noether loci

When ch1(𝐯)>1\text{ch}_{1}(\mathbf{v})>1, the Brill-Noether loci can be reducible.

When ch0(𝐯)=1\text{ch}_{0}(\mathbf{v})=1, we consider twists of ideal sheaves IZ(d)I_{Z}(d), where d>0d>0 and Z2Z\subseteq\mathbb{P}^{2} is a subscheme of finite length. These moduli spaces are isomorphic to the Hilbert scheme 2[n]\mathbb{P}^{2[n]}, but their cohomological properties depend on the twist. We often write 2[n](d)\mathbb{P}^{2[n]}(d) for these moduli spaces.

Proposition 5.3.

Let 𝐯=(1,d,d2/2n)\mathbf{v}=(1,d,d^{2}/2-n) be the Chern character of a twist IZ(d)I_{Z}(d) of an ideal sheaf of a scheme of length nn. If n>d2n>d^{2}, the Brill-Noether loci B2(𝐯)B^{2}(\mathbf{v}) are reducible, and have at least d1d-1 components.

Proof.

Choose ZZ such that h0(IZ(d))2h^{0}(I_{Z}(d))\geq 2, and let C1,C22C_{1},C_{2}\subseteq\mathbb{P}^{2} be curves of degree dd containing ZZ. If C1C2C_{1}\cap C_{2} were zero-dimensional it would have length d2<(Z)d^{2}<\ell(Z), so in particular it could not contain ZZ. We conclude that C1C2C_{1}\cap C_{2} has one-dimensional components. The general such ZZ lies in B2B3B^{2}\smallsetminus B^{3} so the pencil spanned by C1C_{1} and C2C_{2} is determined by ZZ. In this case we call CZC_{Z} the union of the one-dimensional components of C1C2C_{1}\cap C_{2}; the general such ZZ will have CZC_{Z} irreducible.

For any 0<e<d0<e<d, we examine the following loci in B2B^{2}:

Be2:={Z:CZ is irreducible with deg(CZ)=e}¯.B^{2}_{e}:=\overline{\{Z:C_{Z}\text{ is irreducible with }\deg(C_{Z})=e\}}.

Clearly B2B^{2} is the union of the Be2B^{2}_{e} as ee ranges between 1 and d1d-1; we claim that each lies in a distinct irreducible component of B2B^{2}.

Indeed, the association of a general element IZ(d)B2I_{Z}(d)\in B^{2} to the pencil of reducible degree dd curves containing it induces a rational map ϕ:B2G(2,p(d))\phi:B^{2}\dashrightarrow G(2,p(d)), and the pencils themselves are lines on a Segre variety Σdk,kH0(𝒪2(d))\Sigma_{d-k,k}\subseteq\mathbb{P}H^{0}(\mathcal{O}_{\mathbb{P}^{2}}(d)) lying in a fiber of one of its projections Σdk,kH0(𝒪2(dk)),H0(𝒪2(k))\Sigma_{d-k,k}\rightarrow\mathbb{P}H^{0}(\mathcal{O}_{\mathbb{P}^{2}}(d-k)),\mathbb{P}H^{0}(\mathcal{O}_{\mathbb{P}^{2}}(k)). One can check that each such line (given by fixing a smooth curve of degree dkd-k and a pencil of curves of degree kk) lies in the image of ϕ\phi. Denote by Pdk,kP_{d-k,k} the loci of such lines in G(2,p(d))G(2,p(d)).

As kk varies between 1 and d1d-1, there are no containments among the Segre varieties Σdk,k\Sigma_{d-k,k} in H0(𝒪2(d))\mathbb{P}H^{0}(\mathcal{O}_{\mathbb{P}^{2}}(d)), and the line classes in the fibers of the projections form irreducible components of the image of ϕ\phi. Each of the loci Be2B^{2}_{e} dominates Pe,deP_{e,d-e}, so the claim follows. ∎

L1L_{1}L2L_{2}RRZZC1C_{1}C2C_{2}LLZZ
Figure 3. General members of the loci B21B_{2}^{1} and B22B_{2}^{2} on 2[10](3)\mathbb{P}^{2[10]}(3).

We will use reducibility in rank 1 to form reducible Brill-Noether loci in higher rank via the Serre construction.

Theorem 5.4 ([HL10], 5.1.1).

Let XX be a smooth surface, L,MPic(X)L,M\in\text{Pic}(X) be line bundles on XX, and ZXZ\subseteq X a local complete intersection subscheme of codimension 2. Then there is a locally free extension

0LEMIZ00\rightarrow L\rightarrow E\rightarrow M\otimes I_{Z}\rightarrow 0

(and hence the general such extension is locally free), if and only if ZZ satisfies the Cayley-Bacharach property with respect to LMKXL^{\vee}MK_{X}: for any ZZZ^{\prime}\subseteq Z of length (Z)1\ell(Z)-1, the map

H0(LMKXIZ)H0(LMKXIZ)H^{0}(L^{\vee}MK_{X}\otimes I_{Z})\rightarrow H^{0}(L^{\vee}MK_{X}\otimes I_{Z^{\prime}})

is surjective.

Assume that Z2Z\subseteq\mathbb{P}^{2} satisfies the Cayley-Bacharach property for 𝒪2(ba3)\mathcal{O}_{\mathbb{P}^{2}}(b-a-3). Then the general extension

0𝒪(a)EIZ(b)00\rightarrow\mathcal{O}(a)\rightarrow E\rightarrow I_{Z}(b)\rightarrow 0

is locally free. To check that the general such EE is slope stable, it is enough to check that there are no maps 𝒪2(d)E\mathcal{O}_{\mathbb{P}^{2}}(d)\rightarrow E with d>μ(E)=(a+b)/2d>\mu(E)=(a+b)/2. Indeed, any destabilizing subbundle SES\subseteq E has rank 1, and since EE is locally free, this inclusion induces SEES^{\vee\vee}\subseteq E^{\vee\vee}\simeq E, and SS^{\vee\vee} is locally free of rank 1, i.e., a line bundle.

Example 5.5.

The extensions we consider are in the above form, with (a,b)=(0,3)(a,b)=(0,3) and (Z)=10\ell(Z)=10:

0𝒪EIZ(3)0.0\rightarrow\mathcal{O}\rightarrow E\rightarrow I_{Z}(3)\rightarrow 0.

Here χ(E)=h0(E)=1\chi(E)=h^{0}(E)=1 for generic choice of ZZ. Set 𝐯=ch(E)\mathbf{v}=\text{ch}(E). Note that the Cayley-Bacharach condition is trivially satisfied, as the map in question is

H0(IZ)=00=H0(IZ).H^{0}(I_{Z})=0\rightarrow 0=H^{0}(I_{Z^{\prime}}).

Stability is also easy to check. We need to rule out the existence of maps 𝒪(d)E\mathcal{O}(d)\rightarrow E where d2d\geq 2. Any such map induces a nonzero map to from 𝒪(d)\mathcal{O}(d) to either 𝒪\mathcal{O} or IZ(3)I_{Z}(3), but either such map vanishes by stability when we choose ZZ noncollinear. Thus the general extension EE is both locally free and slope-stable. We construct multiple components in B3(𝐯)B^{3}(\mathbf{v}) by considering the components of B2(IZ(d))B^{2}(I_{Z}(d)).

The Brill-Noether locus B2B_{2} on the twisted Hilbert scheme 2[10](3)\mathbb{P}^{2[10]}(3) has (at least) two components: B12B^{2}_{1} and B22B^{2}_{2}. The general member ZB12Z\in B^{2}_{1} is contained in two cubics XX and YY each of which is the union of a fixed line LL with a conic C1C_{1} or C2C_{2}. The general ZB22Z\in B^{2}_{2} is contained in two cubics UU and VV each of which is the union of a fixed conic RR with a line L1L_{1} or L2L_{2}. See Figure 3.

We consider the following families of stable sheaves EE: construct a projective bundle P1P_{1} over B12(IZ(d))B^{2}_{1}(I_{Z}(d)) with fiber Ext1(IZ(3),𝒪)\mathbb{P}\text{Ext}^{1}(I_{Z}(3),\mathcal{O}) for ZB12Z\in B^{2}_{1}. It is generically a family of stable sheaves, so admits a map P1M(ch(E))P_{1}\dashrightarrow M(\text{ch}(E)); construct P2M(𝐯)P_{2}\dashrightarrow M(\mathbf{v}) similarly. One can check that the image closures B1B_{1} and B2B_{2} of P1P_{1} and P2P_{2} have dimensions 23, above the expected dimension 22 for a component of B2(𝐯)B^{2}(\mathbf{v}).

To conclude that B3(𝐯)B^{3}(\mathbf{v}) is reducible, we need to verify that B1B_{1} and B2B_{2} are not nested, and there is no component of B3(𝐯)B_{3}(\mathbf{v}) containing both. The first statement is straightforward, following from the definitions of the loci. For the second, suppose VB3(𝐯)V\subseteq B^{3}(\mathbf{v}) is irreducible and contains both B1B_{1} and B2B_{2}. Then the general EVE\in V has three global sections, and since the general points of B1B_{1} and B2B_{2} have global sections 𝒪F\mathcal{O}\rightarrow F with torsion-free cokernel, so too does the general point of VV. Thus EE appears in an extension

0𝒪EIZ(3)00\rightarrow\mathcal{O}\rightarrow E\rightarrow I_{Z}(3)\rightarrow 0

with (Z)=10\ell(Z)=10. Since EB3(𝐯)E\in B^{3}(\mathbf{v}), h0(IZ(3))2h^{0}(I_{Z}(3))\geq 2 so ZZ lies in B12B^{2}_{1} or B22B^{2}_{2}. However a general point of VV specializing to a general point of B1B_{1} or B2B_{2} determines a specialization of B12B^{2}_{1} to B22B_{2}^{2}, or vice versa. To see this, let 𝒱/V\mathcal{V}/V be a universal family, perhaps after shrinking VV. Then we claim there is a distinguished section 𝒪2×V𝒱\mathcal{O}_{\mathbb{P}^{2}\times V}\rightarrow\mathcal{V} whose cokernel \mathcal{I} is generically a family of ideal sheaves IZ(3)I_{Z}(3), fitting in an exact sequence

0𝒪2×V𝒱00\rightarrow\mathcal{O}_{\mathbb{P}^{2}\times V}\rightarrow\mathcal{V}\rightarrow\mathcal{I}\rightarrow 0

which on fibers recovers the sequences defining the loci B1B_{1} and B2B_{2}. In particular, it will follow that a specialization of a general point of VV to a general point of B1B_{1} and to a general point of B2B_{2} determines a specialization of B12B^{2}_{1} to B22B^{2}_{2}, contradicting Proposition 5.3. To see this, let EB1E\in B_{1} be given, and choose a section tH0(E)t\in H^{0}(E) fitting in an exact sequence

0𝒪tEIZt(3)00\rightarrow\mathcal{O}\stackrel{{\scriptstyle t}}{{\rightarrow}}E\rightarrow I_{Z_{t}}(3)\rightarrow 0

for Zt2Z_{t}\subseteq\mathbb{P}^{2} a zero-dimensional subscheme of length 10 with IZt(3)B122[10](3)I_{Z_{t}}(3)\in B^{2}_{1}\subseteq\mathbb{P}^{2[10]}(3). Then choose general s1,s2H0(E)s_{1},s_{2}\in H^{0}(E) independent from tt. The section s1s_{1} determines another sequence

0𝒪s1EIZs1(3)00\rightarrow\mathcal{O}\stackrel{{\scriptstyle s_{1}}}{{\rightarrow}}E\rightarrow I_{Z_{s_{1}}}(3)\rightarrow 0

where Zs12Z_{s_{1}}\subseteq\mathbb{P}^{2} is a zero-dimensional subscheme, necessarily of length 10. We claim that Zs1B12Z_{s_{1}}\in B^{2}_{1} as well. To see this, consider the dependency loci X(s1,t)X(s_{1},t) and X(s1,s2)X(s_{1},s_{2}) where, respectively, s1s_{1} and tt are dependent and where s1s_{1} and s2s_{2} are dependent. Each is a degree 3 curve in 2\mathbb{P}^{2}, and Zs1X(s1,t)X(s1,s2)Z_{s_{1}}\subseteq X(s_{1},t)\cap X(s_{1},s_{2}). However since (Zs1)=10\ell(Z_{s_{1}})=10, the intersection cannot be transverse, and the dependency loci share a component. Since ZtB12Z_{t}\in B^{2}_{1}, there is a line in the shared loci containing Zs1Z_{s_{1}}. The claim is now proved, which completes the example.

In fact in higher rank many Brill-Noether loci are reducible, as we now show. The context for the following result is established in Section 4. Recall that the parabola ξr\xi_{r} is the locus of characters 𝐯K(2)\mathbf{v}\in K(\mathbb{P}^{2}) satisfying χ(𝐯)=ch0(𝐯)\chi(\mathbf{v})=\text{ch}_{0}(\mathbf{v}), thought of in the (μ,Δ)(\mu,\Delta)-plane, and we set ξr(a)\xi_{r}(a) to be the intersection of ξr\xi_{r} with the vertical line μ=a\mu=a.

Remark 5.6.

In [CH18, Definition 3.1] a somewhat different definition of the extremal character to 𝐯\mathbf{v} is given. We will need to use this other definition alongside ours in what follows; we set 𝐯CH\mathbf{v}^{\prime}_{CH} to be the extremal character in the sense of Coskun-Huizenga in [CH18]. The character 𝐯CH\mathbf{v}^{\prime}_{CH} is defined as in Definition 3.2 but replacing axioms (D1) and (D2) as follows:

  1. (D1’)

    ch0(𝐯CH)ch0(𝐯)\text{ch}_{0}(\mathbf{v}^{\prime}_{CH})\leq\text{ch}_{0}(\mathbf{v}) and if ch0(𝐯CH)=ch0(𝐯)\text{ch}_{0}(\mathbf{v}^{\prime}_{CH})=\text{ch}_{0}(\mathbf{v}) then ch1(𝐯)ch1(𝐯CH)>0\text{ch}_{1}(\mathbf{v})-\text{ch}_{1}(\mathbf{v}^{\prime}_{CH})>0;

  2. (D2’)

    μ(𝐯CH)<μ(𝐯)\mu(\mathbf{v}^{\prime}_{CH})<\mu(\mathbf{v}).

It will follow from Lemma 5.7 that μ(𝐯CH)μ(𝐯)\mu(\mathbf{v}^{\prime}_{CH})\leq\mu(\mathbf{v}^{\prime}). The variant 𝐯CH\mathbf{v}_{CH}^{\prime} has the property that when ch0(𝐯)\text{ch}_{0}(\mathbf{v}) and ch1(𝐯)\text{ch}_{1}(\mathbf{v}) are not coprime, we still have μ(𝐯CH)<μ(𝐯)\mu(\mathbf{v}_{CH}^{\prime})<\mu(\mathbf{v}).

Lemma 5.7.

If gcd(r,c1)>1\gcd(r,c_{1})>1 and c1r\frac{c_{1}}{r}\not\in\mathbb{Z}, then the extremal character 𝐯CH\mathbf{v}_{CH}^{\prime} has r(𝐯CH)<r(𝐯)r(\mathbf{v}_{CH}^{\prime})<r(\mathbf{v}).

Proof.

If not, then μ(𝐯)=c11r\mu(\mathbf{v}^{\prime})=\frac{c_{1}-1}{r}. We claim that there exists an integer nn such that c11r<nr1<c1r\frac{c_{1}-1}{r}<\frac{n}{r-1}<\frac{c_{1}}{r}. This is equivalent to the inequalities (c11)(r1)r<n<(c1)(r1)r\frac{(c_{1}-1)(r-1)}{r}<n<\frac{(c_{1})(r-1)}{r}. The sequence of rr consecutive integers

(c11)(r1),(c11)(r1)+1,,c1(r1)(c_{1}-1)(r-1),\quad(c_{1}-1)(r-1)+1,\quad...,\quad c_{1}(r-1)

contains exactly one entry divisible by rr. By assumption, c1r\frac{c_{1}}{r}\notin\mathbb{Z}, so rr does not divide c1(r1)c_{1}(r-1). It suffices to prove that rr does not divide (c11)(r1)(c_{1}-1)(r-1). If r|(c11)(r1)r|(c_{1}-1)(r-1), then r|(c11)r|(c_{1}-1). Set m=gcd(c1,r)m=\gcd(c_{1},r). Then mm divides r,c11r,c_{1}-1, and c1c_{1}, which contradicts the assumption that m>1m>1. ∎

Theorem 5.8.

Let 𝐯K(2)\mathbf{v}\in K(\mathbb{P}^{2}) be a stable Chern character with μ(𝐯)>0\mu(\mathbf{v})>0. Let μCH\mu_{CH}^{\prime} denote the slope of the extremal character 𝐯CH\mathbf{v}_{CH}^{\prime} for 𝐯\mathbf{v}. If μCH>1/3\mu_{CH}^{\prime}>1/3, Δ(𝐯)0\Delta(\mathbf{v})\gg 0 and μ(𝐯)\mu(\mathbf{v})\notin\mathbb{Z}, then Br(𝐯)B^{r}(\mathbf{v}) is reducible.

  1. (1)

    If Δ(𝐯)ξr(μ(𝐯))\Delta(\mathbf{v})\geq\xi_{r}(\mu(\mathbf{v})) and μCH1/3\mu_{CH}^{\prime}\geq 1/3, then Br(𝐯)B^{r}(\mathbf{v}) contains a component of the expected dimension.

  2. (2)

    If Δ(𝐯)0\Delta(\mathbf{v})\gg 0 and μCH>1/3\mu_{CH}^{\prime}>1/3, then Br(𝐯)B^{r}(\mathbf{v}) contains a component of dimension larger than the expected dimension.

Note that the assumptions in the theorem are implied by the assumptions in Theorem 1.1 (4).

We first construct the families of sheaves necessary for the proof and study their properties. Our first family will consist of sheaves EE appearing in extensions determined by r(E)r(E) global sections of EE, and our second family will consist of sheaves FF appearing in extension classes determined by the extremal subsheaves EE^{\prime}.

First, let Pic𝒞/Ud\text{Pic}_{\mathcal{C}/U}^{d} be the relative Picard scheme over the universal curve 𝒞U\mathcal{C}\rightarrow U, where UH0𝒪2(c1)U\subseteq\mathbb{P}H^{0}\mathcal{O}_{\mathbb{P}^{2}}(c_{1}) is the space of smooth plane curves of degree c1c_{1}. We set g=12(c11)(c12)g=\frac{1}{2}(c_{1}-1)(c_{1}-2). Consider the projective bundle YY whose fiber over 𝒪C(D)\mathcal{O}_{C}(D) is Ext1(𝒪C(D),𝒪r)\mathbb{P}\text{Ext}^{1}(\mathcal{O}_{C}(D),\mathcal{O}^{r}) defined over the open subset of Pic𝒞/Ud\text{Pic}_{\mathcal{C}/U}^{d} where the dimension of this group is constant. We form a family of sheaves \mathcal{E} on 2\mathbb{P}^{2} parametrized by YY by setting EyE_{y}, yYy\in Y, to be an extension sheaf

0𝒪rEy𝒪C(D)00\rightarrow\mathcal{O}^{r}\rightarrow E_{y}\rightarrow\mathcal{O}_{C}(D)\rightarrow 0

determined by y=(𝒪C(D),e)y=(\mathcal{O}_{C}(D),e) for 𝒪C(D)Pic𝒞/Ud\mathcal{O}_{C}(D)\in\text{Pic}^{d}_{\mathcal{C}/U} and eExt1(𝒪C(D),𝒪r)e\in\mathbb{P}\text{Ext}^{1}(\mathcal{O}_{C}(D),\mathcal{O}^{r}).

Recall that a sheaf EE on 2\mathbb{P}^{2} is called prioritary if Ext2(E,E(1))=0\text{Ext}^{2}(E,E(-1))=0.

Lemma 5.9.

Let \mathcal{E} be the family of sheaves just constructed. Then for d=g1d=g-1,

  1. (1)

    \mathcal{E} is complete;

  2. (2)

    the general member of \mathcal{E} is prioritary.

When c1/r1/3c_{1}/r\geq 1/3, the general member is stable.

Proof.

(1) First, we have a natural identification

T𝒪C(D)Pic𝒞/UdExt1(𝒪C(D),𝒪C(D)).T_{\mathcal{O}_{C}(D)}\text{Pic}^{d}_{\mathcal{C}/U}\simeq\text{Ext}^{1}(\mathcal{O}_{C}(D),\mathcal{O}_{C}(D)).

Let TyT_{y} be the tangent space to YY at yYy\in Y, and let E=EyE=E_{y} be the associated sheaf. The Kodaira-Spencer map κ:TyExt1(E,E)\kappa:T_{y}\rightarrow\text{Ext}^{1}(E,E) fits into the following diagram:

0{0}Ext1(𝒪C(D),𝒪r){\text{Ext}^{1}(\mathcal{O}_{C}(D),\mathcal{O}^{r})}Ty{T_{y}}Ext1(𝒪C(D),𝒪C(D)){\text{Ext}^{1}(\mathcal{O}_{C}(D),\mathcal{O}_{C}(D))}0{0}Ext1(E,𝒪r){\text{Ext}^{1}(E,\mathcal{O}^{r})}Ext1(E,E){\text{Ext}^{1}(E,E)}Ext1(E,𝒪C(D)){\text{Ext}^{1}(E,\mathcal{O}_{C}(D))}κ\scriptstyle{\kappa}

The left-hand vertical map is obtained by applying Hom(,𝒪r)\text{Hom}(-,\mathcal{O}^{r}) to the sequence

0𝒪rE𝒪C(D)0.0\rightarrow\mathcal{O}^{r}\rightarrow E\rightarrow\mathcal{O}_{C}(D)\rightarrow 0.

Since Ext1(𝒪r,𝒪r)=0\text{Ext}^{1}(\mathcal{O}^{r},\mathcal{O}^{r})=0, it is surjective. The right-hand vertical map is obtained by applying Hom(,𝒪C(D))\text{Hom}(-,\mathcal{O}_{C}(D)) to the same sequence. The next term in the sequence is

Ext1(𝒪r,𝒪C(D))H1(C,𝒪C(D))r\text{Ext}^{1}(\mathcal{O}^{r},\mathcal{O}_{C}(D))\simeq H^{1}(C,\mathcal{O}_{C}(D))^{r}

whose dimension is

ext1(𝒪r,𝒪C(D))=rh1(C,𝒪C(D))=rh0(C,𝒪C(KCD).\text{ext}^{1}(\mathcal{O}^{r},\mathcal{O}_{C}(D))=r\cdot h^{1}(C,\mathcal{O}_{C}(D))=r\cdot h^{0}(C,\mathcal{O}_{C}(K_{C}-D).

We have

deg(KCD)=2g2d=g1.\deg(K_{C}-D)=2g-2-d=g-1.

For general DD, Riemann-Roch on CC then gives h0(𝒪C(KCD))=0h^{0}(\mathcal{O}_{C}(K_{C}-D))=0, so the right-hand vertical map in the diagram is also surjective. The snake lemma implies that κ\kappa is surjective as well, i.e., \mathcal{E} is complete.

(2) Apply Hom(E,)\text{Hom}(E,-) to the sequence

0𝒪(1)rE(1)𝒪C(DH)00\rightarrow\mathcal{O}(-1)^{r}\rightarrow E(-1)\rightarrow\mathcal{O}_{C}(D-H)\rightarrow 0

to get a sequence

(4) Ext2(E,𝒪(1)r)Ext2(E,E(1))Ext2(E,𝒪C(DH))0.\text{Ext}^{2}(E,\mathcal{O}(-1)^{r})\rightarrow\text{Ext}^{2}(E,E(-1))\rightarrow\text{Ext}^{2}(E,\mathcal{O}_{C}(D-H))\rightarrow 0.

To show general EE is prioritary, it is enough to show that the outer two terms generically vanish.

The first term is Serre dual to Hom(𝒪(1),E(3))rH0(E(2))r\text{Hom}(\mathcal{O}(-1),E(-3))^{r}\simeq H^{0}(E(-2))^{r}. This space sits in an exact sequence

0H0(𝒪(2)r)H0(E(2))H0(𝒪C(D2H))0.0\rightarrow H^{0}(\mathcal{O}(-2)^{r})\rightarrow H^{0}(E(-2))\rightarrow H^{0}(\mathcal{O}_{C}(D-2H))\rightarrow 0.

The first term vanishes, and as above for general DD, h0(𝒪C(D2H))h0(𝒪C(D))=0h^{0}(\mathcal{O}_{C}(D-2H))\leq h^{0}(\mathcal{O}_{C}(D))=0. We conclude the first term in the sequence (4) vanishes.

The last term in sequence (4) is Serre dual to Hom(𝒪C(DH),E(3))\text{Hom}(\mathcal{O}_{C}(D-H),E(-3)). To show this vanishes, it is enough to show that generic EE appearing in \mathcal{E} is torsion-free. If TET\subseteq E is the torsion part of EE, write E=E/TE^{\prime}=E/T. The induced map T𝒪C(D)T\rightarrow\mathcal{O}_{C}(D) is nonzero, so we obtain a diagram

0{0}0{0}K{K}T{T}𝒪C(D){\mathcal{O}_{C}(D)}Q{Q}𝒪r{\mathcal{O}^{r}}E{E}𝒪C(D){\mathcal{O}_{C}(D)}0{0}E{E^{\prime}}E{E^{\prime}}0{0}0{0}0{0}0{0}

The snake lemma produces an exact sequence

0K𝒪rEQ0.0\rightarrow K\rightarrow\mathcal{O}^{r}\rightarrow E^{\prime}\rightarrow Q\rightarrow 0.

Since 𝒪r\mathcal{O}^{r} is torsion-free, K=0K=0. Arguing in similar fashion to the proof of Lemma 4.5 (1), we see that the inclusion T𝒪C(D)T\hookrightarrow\mathcal{O}_{C}(D) cannot have a nonzero cokernel, which would be supported in codimension two. Thus T𝒪C(D)T\simeq\mathcal{O}_{C}(D), but then the map TET\rightarrow E splits the sequence defining EE. This contradiction implies that EE is torsion-free, as required. We conclude that the general member of the family \mathcal{E} is prioritary.

Now 𝐯(E)\mathbf{v}(E) lies on the curve ξr\xi_{r}. When μ(E)1/3\mu(E)\geq 1/3, 𝐯(E)\mathbf{v}(E) lies on or above the Drézet-Le Potier curve. To conclude that the general member of \mathcal{E} is also stable, one proceeds in entirely the same way as in [LP97, §16.2]. Specifically, the stack of prioritary sheaves is irreducible and its general member is stable. The family \mathcal{E} induces a map to the stack of prioritary sheaves, and completeness implies the general member of \mathcal{E} is a general member of the stack of prioritary sheaves, so it is stable. ∎

Corollary 5.10.

A general member of \mathcal{E} is slope-stable for dg1d\leq g-1.

Proof.

It suffices to exhibit one slope-stable member. By the lemma, for d=g1d=g-1 a general member of \mathcal{E} is a general member of M(r,g1r,ξr(g1r))M(r,\frac{g-1}{r},\xi_{r}(\frac{g-1}{r})). Because the general member is slope-stable ([DLP85, Corollaire 4.12]), any elementary modification of a general member of \mathcal{E} is again slope-stable. ∎

It follows that the family \mathcal{E} determines a map YM(𝐯)Y\dashrightarrow M(\mathbf{v}), and we set Z1M(𝐯)Z_{1}\subseteq M(\mathbf{v}) to be the closure of its image.

For the second family, we first assume gcd(r,c1)=1\gcd(r,c_{1})=1. Let 𝐯\mathbf{v}^{\prime} be the extremal character associated to 𝐯\mathbf{v} in the sense of Definition 3.2. Because μ(𝐯CH)μ(𝐯)\mu(\mathbf{v}^{\prime}_{CH})\leq\mu(\mathbf{v}^{\prime}), we have μ(𝐯)>1/3\mu(\mathbf{v}^{\prime})>1/3. Setting 𝐯′′=𝐯𝐯\mathbf{v}^{\prime\prime}=\mathbf{v}-\mathbf{v}^{\prime} and letting Δ(𝐯)0\Delta(\mathbf{v})\gg 0, we may choose stable E′′M(𝐯′′)E^{\prime\prime}\in M(\mathbf{v}^{\prime\prime}) and semistable EM(𝐯)E^{\prime}\in M(\mathbf{v}^{\prime}) and form extensions

0EEE′′0.0\rightarrow E^{\prime}\rightarrow E\rightarrow E^{\prime\prime}\rightarrow 0.

Then ch(E)=𝐯\text{ch}(E)=\mathbf{v} and EE is generically stable (see Proposition 3.3). When r(E)r(E) and c1(E)c_{1}(E) are coprime, μ(𝐯)<μ(𝐯)\mu(\mathbf{v}^{\prime})<\mu(\mathbf{v}), and 𝐯\mathbf{v}^{\prime} has minimal discriminant. If r(E)r(E) and c1(E)c_{1}(E) are not coprime, our construction will differ slightly. Because μ(𝐯)>1/3\mu(\mathbf{v}^{\prime})>1/3, we have Δ(𝐯)<ξr(μ(𝐯))\Delta(\mathbf{v}^{\prime})<\xi_{r}(\mu(\mathbf{v}^{\prime})), i.e., χ(E)>r\chi(E^{\prime})>r^{\prime} (see the proof of Theorem 4.7). Write χ(E)=r+ϵ\chi(E^{\prime})=r^{\prime}+\epsilon^{\prime}, and let ϵ:=min{ϵ,r′′}\epsilon:=\min\{\epsilon^{\prime},r^{\prime\prime}\}. By Definition 3.2, r′′>0r^{\prime\prime}>0, hence ϵ>0\epsilon>0. We want to examine the Brill-Noether locus Br′′ϵ(𝐯′′)B^{r^{\prime\prime}-\epsilon}(\mathbf{v}^{\prime\prime}), i.e., the locus of sheaves E′′M(𝐯′′)E^{\prime\prime}\in M(\mathbf{v}^{\prime\prime}) with h0(E′′)r′′ϵh^{0}(E^{\prime\prime})\geq r^{\prime\prime}-\epsilon. Our second family will consist of extension sheaves EE as above, where EE^{\prime} has general cohomology h0(E)=r+ϵh^{0}(E^{\prime})=r^{\prime}+\epsilon^{\prime} and E′′E^{\prime\prime} has potentially special cohomology h0(E′′)=r′′ϵh^{0}(E^{\prime\prime})=r^{\prime\prime}-\epsilon. If we show that the appropriate Brill-Noether locus is nonempty, we may construct a projective bundle \mathbb{P} over M(𝐯)×Br′′ϵ(𝐯′′)M(\mathbf{v}^{\prime})\times B^{r^{\prime\prime}-\epsilon}(\mathbf{v}^{\prime\prime}) supporting the family of sheaves EE as above. Then the family determines a map M(𝐯)\mathbb{P}\dashrightarrow M(\mathbf{v}); let Z2Z_{2} be the closure of the image.

Proof of Theorem 5.8.

We have as above rational maps YZ1Y\dashrightarrow Z_{1} and Z2\mathbb{P}\dashrightarrow Z_{2} into M(𝐯)M(\mathbf{v}). We will show that Z2Z_{2} is well-defined (i.e., the Brill-Noether locus Br′′ϵ(𝐯′′)B^{r^{\prime\prime}-\epsilon}(\mathbf{v}^{\prime\prime}) is nonempty), that Z1Z_{1} is an irreducible component of Br(𝐯)B^{r}(\mathbf{v}), and dim(Z2)>dim(Z1)\dim(Z_{2})>\dim(Z_{1}).

We now consider the first family of sheaves. By construction, the map YZ1Y\dashrightarrow Z_{1} has fiber Aut(𝒪r)\text{Aut}(\mathcal{O}^{r}). Recall that we set Δ=Δ0+k/r\Delta=\Delta_{0}+k/r, where Δ0\Delta_{0} is minimal and k0k\geq 0. We have

dimZ1\displaystyle\dim Z_{1} =dimYdimAut(𝒪r)\displaystyle=\dim Y-\dim\text{Aut}(\mathcal{O}^{r})
=dimPic𝒞/Ud+rext1(𝒪C(D),𝒪)r2\displaystyle=\dim\text{Pic}^{d}_{\mathcal{C}/U}+r\cdot\text{ext}^{1}(\mathcal{O}_{C}(D),\mathcal{O})-r^{2}
=c12+1+r(3c1+k)r2\displaystyle=c_{1}^{2}+1+r(3c_{1}+k)-r^{2}
=rk+(const).\displaystyle=rk+(\text{const}).

Let VBr(𝐯)V\subseteq B^{r}(\mathbf{v}) be an irreducible subvariety containing Z1Z_{1}, and let EZ1E\in Z_{1} be general. Because h0(E)=rh^{0}(E)=r and the evalutation map evE\text{ev}_{E} for EE has full rank, we can shrink VV to assume that each member of VV does as well. In particular, the map ψ\psi constructed in the proof of Theorem 4.6 carries EE to a smooth curve. It follows that the general point of VV is as well, and the inverse image ψ|V1(U)V\psi|_{V}^{-1}(U)\subseteq V is an open dense subset; in particular, VZ1V\subseteq Z_{1}. We conclude finally that Z1Z_{1} is an irreducible component of Br(𝐯)B^{r}(\mathbf{v}).

We now consider the second family of sheaves. We first assume that gcd(r(E),c1(E))=1\gcd(r(E),c_{1}(E))=1. By Definition 3.2, we have r′′>0r^{\prime\prime}>0. We conclude from Theorem 4.7 that the Brill-Noether locus Br′′(𝐯′′)B^{r^{\prime\prime}}(\mathbf{v}^{\prime\prime}) is nonempty, so in particular the larger Brill-Noether locus Br′′ϵ(𝐯′′)B^{r^{\prime\prime}-\epsilon}(\mathbf{v}^{\prime\prime}) is as well. If r=1r^{\prime}=1, then EE^{\prime} is a line bundle since it has minimal discriminant, and H1(E)=0H^{1}(E^{\prime})=0. If r2r^{\prime}\geq 2, by Theorem 2.1, H1(E)=0H^{1}(E)=0 since by construction EE^{\prime} is chosen to be general. So for generic EE as in the construction of Z2Z_{2} we have

h0(E)(r+ϵ)+(r′′ϵ)r(E),h^{0}(E)\geq(r^{\prime}+\epsilon^{\prime})+(r^{\prime\prime}-\epsilon)\geq r(E),

as desired.

We compute the dimension of Z2Z_{2}. As usual, write Δ(𝐯′′)=Δ0+k/r′′\Delta(\mathbf{v}^{\prime\prime})=\Delta_{0}+k/r^{\prime\prime}, where Δ0\Delta_{0} is the minimal discriminant of a stable bundle of slope μ(𝐯′′)\mu(\mathbf{v}^{\prime\prime}) and rank r′′r^{\prime\prime}. The expected dimension of Br′′ϵ(𝐯′′)B^{r^{\prime\prime}-\epsilon}(\mathbf{v}^{\prime\prime}) is

expdimBr′′ϵ(𝐯′′)\displaystyle\text{expdim}B^{r^{\prime\prime}-\epsilon}(\mathbf{v}^{\prime\prime}) =dimM(𝐯′′)expcodimBr′′ϵ(𝐯′′)\displaystyle=\dim M(\mathbf{v}^{\prime\prime})-\text{expcodim}B^{r^{\prime\prime}-\epsilon}(\mathbf{v}^{\prime\prime})
=r′′2(2Δ(𝐯′′)1)+1(r′′ϵ)(r′′ϵχ(𝐯′′))\displaystyle=r^{\prime\prime 2}(2\Delta(\mathbf{v}^{\prime\prime})-1)+1-(r^{\prime\prime}-\epsilon)(r^{\prime\prime}-\epsilon-\chi(\mathbf{v}^{\prime\prime}))
=2r′′k(r′′ϵ)k+(const1)\displaystyle=2r^{\prime\prime}k-(r^{\prime\prime}-\epsilon)k+(\text{const}_{1})

where (const1) is independent of kk. The dimension of \mathbb{P} is given by

(5) dim=dimBr′′ϵ(𝐯′′)+dimM(𝐯)+ext1(E′′,E)12r′′k(r′′ϵ)k+(const1)+r2(2Δ(𝐯)1)=(r+ϵ)k+(const2).\begin{split}\dim\mathbb{P}&=\dim B^{r^{\prime\prime}-\epsilon}(\mathbf{v}^{\prime\prime})+\dim M(\mathbf{v}^{\prime})+\text{ext}^{1}(E^{\prime\prime},E^{\prime})-1\\ &\geq 2r^{\prime\prime}k-(r^{\prime\prime}-\epsilon)k+(\text{const}_{1})+r^{\prime 2}(2\Delta(\mathbf{v}^{\prime})-1)\\ &=(r+\epsilon)k+(\text{const}_{2}).\end{split}

where again (const2) is independent of kk.

We claim that the fiber dimension of the map Z2\mathbb{P}\rightarrow Z_{2} is bounded independent of kk; for EZ2E\in Z_{2} let E\mathbb{P}_{E} denote the fiber over EE. To show this we consider the forgetful maps EQuot(E,𝐯′′)\mathbb{P}_{E}\rightarrow\text{Quot}(E,\mathbf{v}^{\prime\prime}) to the Quot scheme that assigns to an extension class 0EEE′′00\rightarrow E^{\prime}\rightarrow E\rightarrow E^{\prime\prime}\rightarrow 0 the quotient EE′′E\rightarrow E^{\prime\prime}. The fiber of the forgetful map has dimension bounded by

dimAut(E)hom(E,E)s2\dim\text{Aut}(E^{\prime})\leq\hom(E^{\prime},E^{\prime})\leq s^{2}

where the Jordan-Hölder filtration of EE^{\prime} has length ss. In particular it is independent of kk, so it is enough to bound the dimension of the Quot scheme. There is a canonical identification T[EE′′]Quot(E,𝐯′′)Hom(E,E′′)T_{[E\rightarrow E^{\prime\prime}]}\text{Quot}(E,\mathbf{v}^{\prime\prime})\simeq\text{Hom}(E^{\prime},E^{\prime\prime}) and its dimension bounds the dimension of Quot(E,𝐯′′)\text{Quot}(E,\mathbf{v}^{\prime\prime}). Now consider a general line L2L\subseteq\mathbb{P}^{2} in the locally free loci for EE^{\prime} and E′′E^{\prime\prime} and the restriction

0E′′(1)E′′E′′|L00\rightarrow E^{\prime\prime}(-1)\rightarrow E^{\prime\prime}\rightarrow E^{\prime\prime}|_{L}\rightarrow 0

of E′′E^{\prime\prime} to LL. Then Hom(E,E′′(1))=0\text{Hom}(E^{\prime},E^{\prime\prime}(-1))=0, and by the Grauert-Mülich Theorem ([HL10, Theorem 3.2.1]), hom(E,E′′|L)=h0(EE′′|L)\hom(E^{\prime},E^{\prime\prime}|_{L})=h^{0}(E^{\prime\vee}\otimes E^{\prime\prime}|_{L}) is bounded independent of Δ(E′′)\Delta(E^{\prime\prime}), hence independent of kk, and the claim follows.

From the dimension count above, we see that

(6) dimZ2dimdimE=(r+ϵ)k+(const).\dim Z_{2}\geq\dim\mathbb{P}-\dim\mathbb{P}_{E}=(r+\epsilon)k+(\text{const}).

When the rank and degree fail to be coprime, let 𝐯CH\mathbf{v}_{CH}^{\prime} be the extremal character of 𝐯\mathbf{v}. Since Δ(𝐯)0\Delta(\mathbf{v})\gg 0, an extremal decomposition of 𝐯\mathbf{v} exists. Denote the quotient character by 𝐯CH′′\mathbf{v}_{CH}^{\prime\prime}. Now by Lemma 5.7, r(𝐯CH′′)>0r(\mathbf{v}_{CH}^{\prime\prime})>0. Since μ(𝐯CH)>1/3\mu(\mathbf{v}_{CH}^{\prime})>1/3, μ(𝐯CH′′)>1/3\mu(\mathbf{v}_{CH}^{\prime\prime})>1/3. Since 𝐯CH\mathbf{v}_{CH}^{\prime} has minimal discriminant, Δ(𝐯CH)0\Delta(\mathbf{v}_{CH}^{\prime})\gg 0, so Corollary 5.10 implies that the general member of Z1(𝐯CH′′)Z_{1}(\mathbf{v}_{CH}^{\prime\prime}) is slope-stable. The character 𝐯CH\mathbf{v}_{CH}^{\prime} is below ξr\xi_{r} (see the proof of Theorem 4.7), so we may write χ(𝐯CH)=r+ϵ\chi(\mathbf{v}_{CH}^{\prime})=r^{\prime}+\epsilon^{\prime}. Set ϵ=min{ϵ,r′′}>0\epsilon=\min\{\epsilon^{\prime},r^{\prime\prime}\}>0. We have Z1(𝐯CH′′)Br′′(𝐯CH′′)Br′′ϵ(𝐯CH′′)Z_{1}(\mathbf{v}_{CH}^{\prime\prime})\subseteq B^{r^{\prime\prime}}(\mathbf{v}_{CH}^{\prime\prime})\subseteq B^{r^{\prime\prime}-\epsilon}(\mathbf{v}_{CH}^{\prime\prime}). Let B′′B^{\prime\prime} be the component of Br′′ϵ(𝐯CH′′)B^{r^{\prime\prime}-\epsilon}(\mathbf{v}_{CH}^{\prime\prime}) that contains Z1(𝐯CH′′)Z_{1}(\mathbf{v}_{CH}^{\prime\prime}). Then a general member of B′′B^{\prime\prime} is slope-stable, and since Br′′ϵB^{r^{\prime\prime}-\epsilon} is a determinantal variety, dimB′′(r′′+ϵ)k+const′′\dim B^{\prime\prime}\geq(r^{\prime\prime}+\epsilon)k+\text{const}^{\prime\prime}, where const′′ is independent of kk and Δ(𝐯)=Δmin+k/r\Delta(\mathbf{v})=\Delta_{\min}+k/r. By precisely the same argument as in the proof of [CH18, Theorem 6.4], a general extension of the form

0EEE′′00\rightarrow E^{\prime}\rightarrow E\rightarrow E^{\prime\prime}\rightarrow 0

is stable, where E′′B′′E^{\prime\prime}\in B^{\prime\prime} and EM(𝐯)E^{\prime}\in M(\mathbf{v}^{\prime}) are general members. In particular, h0([E])h^{0}([E]) is well-defined for such [E]M(𝐯)[E]\in M(\mathbf{v}) (since there is no other sheaf that is SS-equivalent to EE). If r(𝐯CH)=1r(\mathbf{v}_{CH}^{\prime})=1, since 𝐯CH\mathbf{v}_{CH}^{\prime} has minimal discriminant, EE^{\prime} is a line bundle and h1(E)=0h^{1}(E^{\prime})=0. If r(𝐯CH)>1r(\mathbf{v}_{CH}^{\prime})>1, since EE^{\prime} is general in M(𝐯)M(\mathbf{v}^{\prime}), by Theorem 2.1, h0(E)=r+ϵh^{0}(E^{\prime})=r^{\prime}+\epsilon^{\prime} and h1(E)=0h^{1}(E^{\prime})=0. In either case,

h0(E)=(r+ϵ)+(r′′ϵ)(r+ϵ)+(r′′ϵ)=r+r′′=r.h^{0}(E)=(r^{\prime}+\epsilon^{\prime})+(r^{\prime\prime}-\epsilon)\geq(r^{\prime}+\epsilon)+(r^{\prime\prime}-\epsilon)=r^{\prime}+r^{\prime\prime}=r.

We obtain a rational map Br(𝐯)M(𝐯)\mathbb{P}\dasharrow B^{r}(\mathbf{v})\subseteq M(\mathbf{v}) where \mathbb{P} is the projective bundle over M(𝐯)×B′′M(\mathbf{v}^{\prime})\times B^{\prime\prime} whose fiber over (E,E′′)(E^{\prime},E^{\prime\prime}) is Ext1(E′′,E)\mathbb{P}\text{Ext}^{1}(E^{\prime\prime},E^{\prime}). Denote the image of this rational map by Z2Z_{2}. Then the dimension counts (5) and (6) apply, and we see that

dimZ2=(r+ϵ)k+(const)>rk+(const)=dimZ1\dim Z_{2}=(r+\epsilon)k+(\text{const})>rk+(\text{const})=\dim Z_{1}

when k0k\gg 0. This completes the proof. ∎

Example 5.11.

Let 𝐯kK(2)\mathbf{v}_{k}\in K(\mathbb{P}^{2}) be the Chern character

ch(𝐯)=(ch0(𝐯),ch1(𝐯),ch2(𝐯))=(3,2,1k)\text{ch}(\mathbf{v})=(\text{ch}_{0}(\mathbf{v}),\text{ch}_{1}(\mathbf{v}),\text{ch}_{2}(\mathbf{v}))=(3,2,-1-k)

so that

μ(𝐯)=23,Δ(𝐯)=59+k3.\mu(\mathbf{v})=\frac{2}{3},\quad\Delta(\mathbf{v})=\frac{5}{9}+\frac{k}{3}.

Then 𝐯k\mathbf{v}_{k} is stable for all k0k\geq 0. As in Theorem 5.8, there are two components in the Brill-Noether loci B3(𝐯k)B^{3}(\mathbf{v}_{k}), whose general members are given as extensions

0𝒪3E𝒪C(D)00\rightarrow\mathcal{O}^{3}\rightarrow E\rightarrow\mathcal{O}_{C}(D)\rightarrow 0

with C2C\subseteq\mathbb{P}^{2} a smooth conic and DD of degree 1k1-k, and

0T2(1)EIZ(1)00\rightarrow T_{\mathbb{P}^{2}}(-1)\rightarrow E\rightarrow I_{Z}(1)\rightarrow 0

where Z2Z\subseteq\mathbb{P}^{2} is a codimension two subscheme of length k+1k+1. For both extension types the general extension bundle EE is stable, as in Theorem 5.8, and the locus Z1Z_{1} whose points correspond to the first extension type form an irreducible component of B3(𝐯k)B^{3}(\mathbf{v}_{k}). Its dimension is

dimZ1=dimPic𝒞/U1k+ext1(𝒪C(D),𝒪3)1+dimGL3=3k+8.\dim Z_{1}=\dim\text{Pic}^{1-k}_{\mathcal{C}/U}+\text{ext}^{1}(\mathcal{O}_{C}(D),\mathcal{O}^{3})-1+\dim GL_{3}=3k+8.

The locus Z2Z_{2} corresponding to the second extension type has dimension

dimZ2=dim2[k+1]+ext1(IZ(1),T2(1))1+(const)=4k+(const)\dim Z_{2}=\dim\mathbb{P}^{2[k+1]}+\text{ext}^{1}(I_{Z}(1),T_{\mathbb{P}^{2}}(-1))-1+(\text{const})=4k+(\text{const})

Thus Z2Z_{2} lies in an irreducible component distinct from Z1Z_{1}, and we conclude the Brill-Noether locus B3(𝐯k)B^{3}(\mathbf{v}_{k}) is reducible for k0k\gg 0.

Funding

This work was supported by the National Science Foundation [grant no. 1246844] to B.G.

Acknowledgements

We are happy to thank Izzet Coskun for very many helpful conversations and his support while this work was carried out. We thank Geoffrey Smith for helpful comments on a preliminary draft of this paper. We thank the referees for many valuable suggestions on an earlier draft.

The figures in this paper were generated on Mathematica and graffitikz. graffitikz is an open-source 2D vector shape editor that supports TikZ output, created by Wenyu Jin. It can be found online at https://github.com/wyjin/graffitikz.

References

  • [ACGH85] E. Arbarello, M. Cornalba, P. Griffiths, and J.D. Harris. Geometry of Algebraic Curves: Volume I. Grundlehren der mathematischen Wissenschaften. Springer New York, 1985.
  • [Alp13] J. Alper. Good moduli spaces for Artin stacks. Annales de l’Institut Fourier, 63(6):2349–2402, 2013.
  • [CH16] I. Coskun and J. Huizenga. The ample cone of the moduli spaces of sheaves on the plane. Algebraic Geometry, 3(1):106–136, 2016.
  • [CH18] I. Coskun and J. Huizenga. The nef cone of the moduli spaces of sheaves and strong Bogomolov inequalities. Israel Journal of Mathematics, 226(1):205–236, 2018.
  • [CH20] I. Coskun and J. Huizenga. Brill–Noether theorems and globally generated vector bundles on Hirzebruch surfaces. Nagoya Mathematical Journal, 238:1–36, 2020.
  • [CH21] I. Coskun and J. Huizenga. Existence of semistable sheaves on Hirzebruch surfaces. Advances in Mathematics, 381, 2021.
  • [CHW17] I. Coskun, J. Huizenga, and M. Woolf. The effective cone of the moduli space of sheaves on the plane. Journal of the European Mathematical Society, 19(5):1421–1467, 2017.
  • [CMR10] L. Costa and R.M. Miró-Roig. Brill-Noether theory for moduli spaces of sheaves on algebraic varieties. Forum Math., 22(3):411–432, 2010.
  • [CNY21] I. Coskun, H. Nuer, and K. Yoshioka. The cohomology of the general stable sheaf on a K3 surface. arXiv: 2106.13047, 2021.
  • [DLP85] J.-M. Drézet and J. Le Potier. Fibrés stables et fibrés exceptionnels sur 2\mathbb{P}_{2}. Annales scientifiques de l’École Normale Supérieure, 4e série, 18(2):193–243, 1985.
  • [Dré86] J.-M. Drézet. Fibrés exceptionnels et suite spectrale de Beilinson généralisée sur 2()\mathbb{P}_{2}(\mathbb{C}). Mathematische Annalen, 275:25–48, 1986.
  • [Dré08] J.-M. Drézet. Moduli spaces of coherent sheaves on multiple curves. In P. Pragacz, editor, Algebraic cycles, sheaves, shtukas, and moduli, Impanga lecture notes, pages 33–43. Birkhäuser, Basel, 2008.
  • [Eis13] D. Eisenbud. Commutative Algebra: with a View Toward Algebraic Geometry. Graduate Texts in Mathematics. Springer New York, 2013.
  • [GH98] L. Göttsche and A. Hirschowitz. Weak Brill-Noether for vector bundles on the projective plane. In Algebraic geometry (Catania, 1993/Barcelona, 1994), Lecture Notes in Pure and Appl. Math., pages 63–74, New York, 1998. Dekker.
  • [HL10] D. Huybrechts and M. Lehn. The Geometry of Moduli Spaces of Sheaves. Cambridge University Press, 2 edition, 2010.
  • [LP97] J. Le Potier. Lectures on Vector Bundles, volume 54 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, 1997.
  • [New21] Peter Newstead. Higher rank Brill-Noether theory and coherent systems: Open questions, 2021.
  • [RLTLZ21] L. Roa-Leguizamón, H. Torres-López, and A. G. Zamora. On the Segre invariant for rank two vector bundles on 2\mathbb{P}^{2}. Advances in Geometry, 21(4):565–576, Jul 2021.