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

The Degeneracy Loci for Smooth Moduli of Sheaves

Yu Zhao Beijing Institute of Technology [email protected]
Abstract.

Let SS be a smooth projective surface over \mathbb{C}. We prove that, under certain technical assumptions, the degeneracy locus of the universal sheaf over the moduli space of stable sheaves is either empty or an irreducible Cohen-Macaulay variety of the expected dimension. We also provide a criterion for when the degeneracy locus is non-empty. This result generalizes the work of Bayer, Chen, and Jiang [5] for the Hilbert scheme of points on surfaces.

The above result is a special case of a general phenomenon: for a perfect complex of Tor-amplitude [0,1][0,1], the geometry of the degeneracy locus is closely related to the geometry of the derived Grassmannian. We analyze their birational geometry and relate it to the incidence varieties studied in [14]. As a corollary, we prove a statement that was previously claimed by the author in [37].

1. Introduction

1.1. The Degeneracy Loci for Smooth Moduli of Sheaves

Given a coherent sheaf \mathcal{F} over a variety XX, the degeneracy locus D¯,l\bar{D}_{\mathcal{F},l} is defined as the closed subvariety

D¯,l:={pXrank(|k(p))l},l0,\bar{D}_{\mathcal{F},l}:=\{p\in X\mid\mathrm{rank}(\mathcal{F}|_{k(p)})\geq l\},\quad l\geq 0,

where k(p)k(p) is the residue field of pp. The degeneracy locus plays a central role in Brill-Noether theory, enumerative geometry, and algebraic combinatorics, as discussed in [2, 10].

For the universal ideal sheaf over the Hilbert scheme of nn points on a projective surface SS over \mathbb{C}, Chen, Jiang, and Bayer [5] proved that its degeneracy locus D¯,l\bar{D}_{\mathcal{F},l} is either empty or an irreducible Cohen-Macaulay variety of dimension

vdim:=2n+2+l(1l).\mathrm{vdim}:=2n+2+l(1-l).

Moreover, they showed that D¯,l\bar{D}_{\mathcal{F},l} is non-empty if and only if vdim\mathrm{vdim} is positive.

In this paper, we generalize [5] to the rank r>0r>0 moduli space of stable coherent sheaves. We fix c1NS(S)c_{1}\in\mathrm{NS}(S) and an ample line bundle HH. For any nH4(S,)n\in H^{4}(S,\mathbb{Z})\cong\mathbb{Z}, let Hs(r,c1,n)\mathcal{M}^{s}_{H}(r,c_{1},n) be the moduli space of Gieseker HH-stable coherent sheaves with the total Chern class

r+c1+n[pts],r+c_{1}+n[pts],

where [pts][pts] is the cycle of a point in SS. Let 𝒰n\mathcal{U}_{n} be the universal coherent sheaf over Hs(r,c1,n)×S\mathcal{M}^{s}_{H}(r,c_{1},n)\times S. For l0l\geq 0, the expected dimension of the degeneracy locus D¯𝒰n,l\bar{D}_{\mathcal{U}_{n},l} is defined as

vdimn,l:=dim(Hs(r,c1,n))+min{0,l(rl)}.\mathrm{vdim}_{n,l}:=\dim(\mathcal{M}_{H}^{s}(r,c_{1},n))+\min\{0,l(r-l)\}.

We prove that

Theorem 1.1.

Assuming 3.8, for any nn\in\mathbb{Z} and l0l\geq 0, the degeneracy locus D¯𝒰n,l\bar{D}_{\mathcal{U}_{n},l} is either empty or an irreducible Cohen-Macaulay variety of dimension vdimn,l\mathrm{vdim}_{n,l}.

Determining whether D¯𝒰n,l\bar{D}_{\mathcal{U}_{n},l} is empty is more challenging than in the case of Hilbert schemes of points, due to the increased complexity of the moduli space and the involvement of the remainder of ll divided by rr. We prove that

Theorem 1.2.

There exists an integer θ0\theta\geq 0 that depends on SS, HH, rr, and c1c_{1}, such that D¯𝒰n,l\bar{D}_{\mathcal{U}_{n},l}\neq\emptyset if and only if Hs(r,c1,n)\mathcal{M}_{H}^{s}(r,c_{1},n)\neq\emptyset and

vdimn,lθ+t(rt),\mathrm{vdim}_{n,l}\geq\theta+t(r-t),

where tt is the remainder of ll divided by rr.

Remark 1.3.

The precise value of θ\theta is known in many cases. When the rank r=1r=1, we have θ=2\theta=2, as shown in [5]. According to Yoshioka [33, 34], θ=2\theta=2 for K3 surfaces (for a generic ample divisor HH) and θ=4\theta=4 for abelian surfaces. When S=2S=\mathbb{P}^{2} or 1×1\mathbb{P}^{1}\times\mathbb{P}^{1}, the value of θ\theta depends on rr and c1c_{1}, as computed by Drezet and Le Potier [9].

Remark 1.4.

Similar arguments as in Theorem 1.1 and Theorem 1.2 also apply to fr(r,n)\mathcal{M}_{fr}(r,n), the moduli space of framed sheaves on 2\mathbb{P}^{2}, where we have θ=2\theta=2. Since fr(r,n)\mathcal{M}_{fr}(r,n) is also the quiver variety of the Jordan quiver, we expect similar results for more quiver varieties, which will be studied in future work.

Remark 1.5.

Besides its importance in algebraic geometry, the enumerative and birational geometry properties of 𝒰n\mathcal{U}_{n} play a central role in the representations of quantum toroidal algebras [27, 28], deformed WW-algebras [22], and categorification problems [36, 35]. These have further implications in the AGT correspondence [30] and the P=WP=W conjecture [12].

1.2. The Degeneracy Locus and the Grassmannian

One of the main ingredients in the proof of Theorem 1.1 follows from a duality relation between the degeneracy locus and the Grassmannian. Let XX be an irreducible, locally complete intersection variety over an arbitrary field 𝔽\mathbb{F}, and let

V𝜎WV\xrightarrow{\sigma}W

be a morphism of locally free sheaves over XX. Given any l0l\geq 0, we consider the ll-th degeneracy locus D¯σ,l:=D¯coker(σ),l\bar{D}_{\sigma,l}:=\bar{D}_{\mathrm{coker}(\sigma),l}, i.e., the closed subvariety consisting of points where dim(coker(σ))l\dim(\mathrm{coker}(\sigma))\leq l.

By a classical theorem in commutative algebra (e.g., Theorem 14.4 of [10]), for any l0l\geq 0, we have

(1.1) codim(D¯σ,l)max{0,l(le)},\mathrm{codim}(\bar{D}_{\sigma,l})\leq\max\{0,l(l-e)\},

where e:=rank(W)rank(V)e:=\mathrm{rank}(W)-\mathrm{rank}(V). We say that D¯σ,l\bar{D}_{\sigma,l} is of expected dimension if equality holds in 1.1; in this situation, D¯σ,l\bar{D}_{\sigma,l} is Cohen-Macaulay.

On the other hand, for any integer l0l\geq 0, we consider the Grassmannian (in Grothendieck’s notation) GrX(coker(σ),l)Gr_{X}(\mathrm{coker}(\sigma),l), which parametrizes the quotients of σ\sigma by locally free sheaves of rank ll. By Krull’s height theorem,

(1.2) dim(GrX(coker(σ),l))dim(X)l(le),\dim(Gr_{X}(\mathrm{coker}(\sigma),l))\geq\dim(X)-l(l-e),

and we say that GrX(coker(σ),l)Gr_{X}(\mathrm{coker}(\sigma),l) is of expected dimension if equality holds in 1.2 or if GrX(coker(σ),l)=Gr_{X}(\mathrm{coker}(\sigma),l)=\emptyset. Theorem 1.1 follows from Nakajima-Yoshioka’s theory on perverse coherent sheaves of blow-ups [25, 24, 26] and the following proposition.

Proposition 1.6.

The following are equivalent:

  1. (1)

    For all l0l\geq 0, the ll-th degeneracy locus D¯σ,l\bar{D}_{\sigma,l} is of expected dimension.

  2. (2)

    For all l0l\geq 0, the ll-th degeneracy locus D¯σ,l\bar{D}_{\sigma^{\vee},l} is of expected dimension.

  3. (3)

    For all l0l\geq 0, the rank ll Grassmannians GrX(coker(σ),l)Gr_{X}(\mathrm{coker}(\sigma),l) are of expected dimension.

  4. (4)

    For all l0l\geq 0, the rank ll Grassmannians GrX(coker(σ),l)Gr_{X}(\mathrm{coker}(\sigma^{\vee}),l) are of expected dimension.

Moreover, if one of these four families of varieties consists entirely of irreducible varieties, then the other three families are also irreducible.

Remark 1.7.

Under the assumptions of Proposition 1.6, more homological and birational properties of the degeneracy loci and Grassmannians can be studied. For more details, we refer to [1, Appendix A].

Remark 1.8.

While preparing this paper, the author was reminded by Qingyuan Jiang that a similar argument to Proposition 1.6 was given in his note [17].

Following Proposition 1.6, we give the following definition:

Definition 1.9.

We say that σ\sigma has good degeneracy loci if the assumptions in Proposition 1.6 are satisfied. Moreover, we say σ\sigma has perfect degeneracy loci if all the varieties in Proposition 1.6 are irreducible.

1.3. Incidence Varieties, Birational Geometry, and Derived Algebraic Geometry

To prove Theorem 1.2, we need to consider incidence varieties and study their birational geometry, adopting a higher viewpoint from derived algebraic geometry.

First, we observe that the morphism σ\sigma between two locally free sheaves can be naturally generalized to a Tor-amplitude [0,1][0,1] perfect complex. That is, we assume σ\sigma is a two-term complex of locally free sheaves, locally. Let ee denote the rank of σ\sigma. Then, for any point pXp\in X, we have the Fredholm relation

(1.3) dimh0(σ|k(p))dimh1(σ|k(p))=e,\dim\ h^{0}(\sigma|_{k(p)})-\dim\ h^{1}(\sigma|_{k(p)})=e,

where k(p)k(p) is the residue field at pp.

For any l0l\geq 0, Jiang [14] constructed a derived Grassmannian GrX(σ,l)Gr_{X}(\sigma,l), which is a derived enhancement of the Grassmannian

GrX(h0(σ),l).Gr_{X}(h^{0}(\sigma),l).

In particular, Jiang [14] proved that GrX(σ,l)Gr_{X}(\sigma,l) is quasi-smooth, which generalizes the notion of locally complete intersection varieties in derived algebraic geometry. If σ\sigma is a Tor-amplitude [0,1][0,1] perfect complex, then so is the shifted derived dual σ[1]\sigma^{\vee}[1]. When e:=rank(σ)0e:=\mathrm{rank}(\sigma)\geq 0, for any two integers d+,d0d_{+},d_{-}\geq 0, Jiang [14] defined an incidence variety

IncX(σ,d+,d),\mathrm{Inc}_{X}(\sigma,d_{+},d_{-}),

which is a fiber product of GrX(σ,d+)Gr_{X}(\sigma,d_{+}) and GrX(σ[1],d)Gr_{X}(\sigma^{\vee}[1],d_{-}) over XX, albeit with different derived structures. Specifically, we have

IncX(σ,d+,0)GrX(σ,d+),IncX(σ,0,d)GrX(σ[1],d).\mathrm{Inc}_{X}(\sigma,d_{+},0)\cong Gr_{X}(\sigma,d_{+}),\quad\mathrm{Inc}_{X}(\sigma,0,d_{-})\cong Gr_{X}(\sigma^{\vee}[1],d_{-}).

Given a quasi-smooth derived scheme YY over 𝔽\mathbb{F}, we say that YY is classical if it has a trivial derived structure, i.e., π0(Y)Y\pi_{0}(Y)\cong Y, where π0(Y)\pi_{0}(Y) is the underlying scheme. By Corollary 2.11 of [3], YY is classical if and only if it has the expected dimension, i.e.,

vdim(Y)=dim(π0(Y)),\mathrm{vdim}(Y)=\dim(\pi_{0}(Y)),

where vdim(Y)\mathrm{vdim}(Y) is the virtual dimension. Moreover, YY is always a locally complete intersection variety if YY is classical. Hence, Proposition 1.6 can be derived from the following generalization:

Theorem 1.10.

Assuming that e:=rank(σ)0e:=\mathrm{rank}(\sigma)\geq 0, the following are equivalent:

  1. (1)

    σ\sigma has good (resp. perfect) degeneracy loci;

  2. (2)

    For any integers d+,d0d_{+},d_{-}\geq 0, the incidence variety

    IncX(σ,d+,d)\mathrm{Inc}_{X}(\sigma,d_{+},d_{-})

    is classical (resp. and irreducible);

  3. (3)

    For any integer d+0d_{+}\geq 0, the derived Grassmannian

    GrX(σ,d+)Gr_{X}(\sigma,d_{+})

    is classical (resp. and irreducible);

  4. (4)

    For any integer d0d_{-}\geq 0, the derived Grassmannian

    GrX(σ[1],d)Gr_{X}(\sigma^{\vee}[1],d_{-})

    is classical (resp. and irreducible).

When the assumptions of Theorem 1.10 are satisfied, the derived Grassmannians and incidence varieties exhibit the following birational property:

Proposition 1.11 (Proposition 2.17).

Let σ\sigma be a Tor-amplitude [0,1][0,1] complex of rank e0e\geq 0 with perfect degeneracy loci. Then for d+,d0d_{+},d_{-}\geq 0 such that d+=d+ed_{+}=d_{-}+e, the morphisms

r+:IncX(σ,d+,d)GrX(σ,d+),\displaystyle r^{+}:\mathrm{Inc}_{X}(\sigma,d_{+},d_{-})\to Gr_{X}(\sigma,d_{+}),\quad r:IncX(σ,d+,d)GrX(σ[1],d),\displaystyle r^{-}:\mathrm{Inc}_{X}(\sigma,d_{+},d_{-})\to Gr_{X}(\sigma^{\vee}[1],d_{-}),
prσ,d+:GrX(σ,d+)D¯σ,d+,\displaystyle\mathrm{pr}_{\sigma,d_{+}}:Gr_{X}(\sigma,d_{+})\to\bar{D}_{\sigma,d_{+}},\quad prσ,d:GrX(σ[1],d)D¯σ,d+\displaystyle\mathrm{pr}_{\sigma,d_{-}}:Gr_{X}(\sigma^{\vee}[1],d_{-})\to\bar{D}_{\sigma,d_{+}}

are birational. Moreover, all the above morphisms are isomorphisms if

d+ddim(X)<(d++1)(d+1).d_{+}d_{-}\leq\dim(X)<(d_{+}+1)(d_{-}+1).

By Proposition 1.11, when all GrX(σ,d+)Gr_{X}(\sigma,d_{+}) and GrX(σ[1],d)Gr_{X}(\sigma^{\vee}[1],d_{-}) are smooth varieties of expected dimensions, the main theorem of [32, 16] provides new examples of the DK conjecture of Bondal-Orlov [7] and Kawamata [18].

1.4. The Virtual Fundamental Class of Incidence Varieties

Let σ\sigma be a Tor-amplitude [0,1][0,1] perfect complex over a proper smooth \mathbb{C}-variety XX such that rank(σ)0\mathrm{rank}(\sigma)\geq 0. There are two natural projection morphisms (given by the fiber diagram):

r+:IncX(σ,d+,d)GrX(σ,d+),r:IncX(σ[1],d)GrX(σ[1],d).r_{+}:\mathrm{Inc}_{X}(\sigma,d_{+},d_{-})\to Gr_{X}(\sigma,d_{+}),\quad r_{-}:\mathrm{Inc}_{X}(\sigma^{\vee}[1],d_{-})\to Gr_{X}(\sigma^{\vee}[1],d_{-}).

According to [14], the morphisms r+r_{+} and rr_{-} are quasi-smooth, and IncX(σ,d+,d)\mathrm{Inc}_{X}(\sigma,d_{+},d_{-}) is also quasi-smooth. Following [8], this quasi-smooth structure induces a perfect obstruction theory in the sense of Li-Tian [21] or Behrend-Fantechi [6], thereby inducing a virtual fundamental class

[IncX(σ,d+,d)]virCHd0(IncX(σ,d+,d)),[\mathrm{Inc}_{X}(\sigma,d_{+},d_{-})]^{\mathrm{vir}}\in CH_{d_{0}}(\mathrm{Inc}_{X}(\sigma,d_{+},d_{-})),

where

d0=dim(X)+(1d+)d+(d+1)d+dd+.d_{0}=\dim(X)+(1-d_{+})d_{+}-(d_{-}+1)d_{-}+d_{-}d_{+}.

In particular, if all GrX(σ,d+)Gr_{X}(\sigma,d_{+}) and GrX(σ[1],d)Gr_{X}(\sigma^{\vee}[1],d_{-}) are smooth varieties of expected dimensions, the virtual fundamental class [IncX(σ,d+,d)]vir[\mathrm{Inc}_{X}(\sigma,d_{+},d_{-})]^{\mathrm{vir}} can be viewed as a correspondence in

CH(GrX(σ,d+)×GrX(σ[1],d)).CH^{*}(Gr_{X}(\sigma,d_{+})\times Gr_{X}(\sigma^{\vee}[1],d_{-})).

This correspondence plays a crucial role in the author’s earlier paper [37], which studies the Clifford algebra action on the cohomology of the moduli space of rank 11 perverse coherent sheaves.

Now, by Theorem 1.10, all the incidence varieties IncX(σ,d+,d)\mathrm{Inc}_{X}(\sigma,d_{+},d_{-}) are locally complete intersection varieties of expected dimensions, thereby proving an earlier argument of the author in [37].

Proposition 1.12 (Proposition 4.6 of [37]).

We have

[IncX(σ,d+,d)]vir=[IncX(σ,d+,d)]CHd0(IncX(σ,d+,d)).[\mathrm{Inc}_{X}(\sigma,d_{+},d_{-})]^{\mathrm{vir}}=[\mathrm{Inc}_{X}(\sigma,d_{+},d_{-})]\in CH_{d_{0}}(\mathrm{Inc}_{X}(\sigma,d_{+},d_{-})).

1.5. Perverse Coherent Sheaves on Blow-ups and Derived Grassmannians

For the tautological bundles/sheaves over moduli spaces, the geometry of derived Grassmannians can often be understood more easily than that of degeneracy loci, as the derived Grassmannians themselves can be described as moduli spaces. A notable example stems from the ADHM construction, which classifies anti-self-dual instantons over 4\mathbb{R}^{4}. Given integers r>0r>0 and n0n\geq 0, we consider two vector spaces VV and WW with dimensions nn and rr, respectively. The ADHM data is defined as

Ms(V,W):={(X,Y,i,j)Hom(V,V)Hom(V,V)Hom(W,V)Hom(V,W)},M^{s}(V,W):=\{(X,Y,i,j)\in\mathrm{Hom}(V,V)\oplus\mathrm{Hom}(V,V)\oplus\mathrm{Hom}(W,V)\oplus\mathrm{Hom}(V,W)\},

subject to the equation

[X,Y]+ij=0,[X,Y]+ij=0,

and the stability condition that for any subspace VVV^{\prime}\subset V, closed under XX and YY and contained in the kernel of jj, we must have V=0V^{\prime}=0. The moduli space of anti-self-dual instantons is given as the geometric quotient

fr(r,n):=Ms(V,W)/GL(V),\mathcal{M}_{fr}(r,n):=M^{s}(V,W)/\mathrm{GL}(V),

which is a smooth variety of dimension 2rn2rn. The variety fr(r,n)\mathcal{M}_{\mathrm{fr}}(r,n) also coincides with the moduli space of framed sheaves over 2\mathbb{P}^{2} and the quiver variety for the Jordan quiver (i.e., a point with one loop).

The universal sheaf 𝒰0\mathcal{U}_{0} on fr(r,n)\mathcal{M}_{fr}(r,n) is represented by the complex

𝒱XYj𝒱𝒱𝒲YXi𝒱,\mathcal{V}\xrightarrow{X\oplus Y\oplus j}\mathcal{V}\oplus\mathcal{V}\oplus\mathcal{W}\xrightarrow{Y\oplus-X\oplus i}\mathcal{V},

which is a two-term complex, as XYjX\oplus Y\oplus j is injective at any closed point by the stability condition. The derived Grassmannians of 𝒰0\mathcal{U}_{0} and 𝒰0[1]\mathcal{U}_{0}^{\vee}[1] were described by Nakajima and Yoshioka [25] via a quiver description. They further generalized this to the moduli space of stable sheaves over a surface, which are known as perverse coherent sheaves on blow-ups [24]. This framework forms another essential component of our proof for Theorem 1.1 and Theorem 1.2.

Recently, the geometric representation theory of the moduli space of perverse coherent sheaves has gained attention, particularly because their cohomology groups form a representation of the affine super-Yangian of 𝔤𝔩(1|1)\mathfrak{gl}(1|1). For further details, we refer to [29, 11, 37].

1.6. Some Notations

The results in Section 2 hold for an arbitrary field 𝔽\mathbb{F}. In Section 3, we will focus specifically on the field of complex numbers \mathbb{C}.

1.7. Acknowledgement

I would like to thank Qingyuan Jiang for many helpful discussions.

2. Degeneracy Theory for Tor-Amplitude [0,1][0,1]-Complexes

In this section, we assume that XX is a locally complete intersection variety over a field 𝔽\mathbb{F}, and σ\sigma is a rank ee Tor-amplitude [0,1][0,1]-complex.

2.1. Degeneracy Loci of a Tor-Amplitude [0,1][0,1]-Complex

Definition 2.1 (Section 14.4 of [10]).

For any integer ll, we define the ll-th degeneracy locus as

D¯σ,l:={xdimh0(σ|k(x))l},\bar{D}_{\sigma,l}:=\{x\mid\dim\ h^{0}(\sigma|_{k(x)})\geq l\},

which is a closed subvariety of XX, where k(x)k(x) denotes the residue field of xx. We also define the locally closed subvariety

Dσ,l:=D¯σ,lD¯σ,l+1={xdimh0(σ|k(x))=l}.D_{\sigma,l}:=\bar{D}_{\sigma,l}-\bar{D}_{\sigma,l+1}=\{x\mid\dim\ h^{0}(\sigma|_{k(x)})=l\}.

By the definition of degeneracy loci, we have:

D¯σ,kXifkmin{0,e},Dσ,kifk<min{0,e},\bar{D}_{\sigma,k}\cong X\quad\mathrm{if}\quad k\leq\min\{0,e\},\quad D_{\sigma,k}\cong\emptyset\quad\mathrm{if}\quad k<\min\{0,e\},

and the duality relation:

D¯σ,k=D¯σ[1],ke.\bar{D}_{\sigma,k}=\bar{D}_{\sigma^{\vee}[1],k-e}.
Example 2.2.

Given two integers m,n>0m,n>0, consider the affine variety

𝕍m,n:=|Hom(𝔽m,𝔽n)|.\mathbb{V}_{m,n}:=|\mathrm{Hom}(\mathbb{F}^{m},\mathbb{F}^{n})|.

There is a universal map of locally free sheaves over 𝕍m,n\mathbb{V}_{m,n}:

vm,n:𝒪m𝒪n,v_{m,n}:\mathcal{O}^{m}\to\mathcal{O}^{n},

which can be regarded as a Tor-amplitude [0,1][0,1]-complex. The degeneracy loci are given by:

D¯vm,n,k={Erank(E)nk},Dvm,n,k={Erank(E)=nk},\bar{D}_{v_{m,n},k}=\{E\mid\mathrm{rank}(E)\leq n-k\},\quad D_{v_{m,n},k}=\{E\mid\mathrm{rank}(E)=n-k\},

which are Cohen-Macaulay and locally complete intersection varieties of codimension

max{0,k(k+mn)}.\max\{0,k(k+m-n)\}.

In general, we have the following property for the Cohen-Macaulay loci:

Proposition 2.3 (Theorem 14.4 (b)(c) of [10]).

For any integer kk, the codimension of the degeneracy loci satisfies the inequalities:

codim(D¯σ,k)max{0,k(ke)},codim(Dσ,k)max{0,k(ke)}.\mathrm{codim}(\bar{D}_{\sigma,k})\leq\max\{0,k(k-e)\},\quad\mathrm{codim}(D_{\sigma,k})\leq\max\{0,k(k-e)\}.

In particular, if equality holds, then D¯σ,k\bar{D}_{\sigma,k} (resp. Dσ,kD_{\sigma,k}) is Cohen-Macaulay (resp. a locally complete intersection variety).

Proof.

As the argument is local, we can assume that XX is affine and

σ{𝒪m𝑠𝒪n}.\sigma\cong\{\mathcal{O}^{m}\xrightarrow{s}\mathcal{O}^{n}\}.

Then the argument follows directly from Theorem 14.4 (b)(c) of [10]. ∎

Definition 2.4.

We say that σ\sigma has good degeneracy loci if for any k0k\geq 0,

codim(D¯σ,k)=max{0,k(ke)}.\mathrm{codim}(\bar{D}_{\sigma,k})=\max\{0,k(k-e)\}.

If σ\sigma has good degeneracy loci, we call the degeneracy loci perfect if D¯σ,k\bar{D}_{\sigma,k} is irreducible for all k0k\geq 0.

The following properties of the degeneracy loci follow from Proposition 2.3 and the Fredholm relation 1.3.

Lemma 2.5.

The complex σ\sigma has good (resp. perfect) degeneracy loci if and only if σ[1]\sigma^{\vee}[1] has good (resp. perfect) degeneracy loci.

Lemma 2.6.

If σ\sigma has good degeneracy loci, then any degeneracy locus D¯σ,k\bar{D}_{\sigma,k} is Cohen-Macaulay, and Dσ,kD_{\sigma,k} is a locally complete intersection variety.

Lemma 2.7.

The complex σ\sigma has good (resp. perfect) degeneracy loci if and only if for any k0k\geq 0,

codim(Dσ,k)=max{0,k(ke)},\mathrm{codim}(D_{\sigma,k})=\max\{0,k(k-e)\},

(and Dσ,kD_{\sigma,k} is irreducible).

2.2. Derived Grassmannians of a Tor-Amplitude [0,1][0,1]-Complex

Definition 2.8 (Definition 4.3 of [14]).

Given an integer d0d\geq 0, the derived Grassmannian GrX(σ,d)Gr_{X}(\sigma,d) is defined as the moduli space of quotients

{σ𝒱},\{\sigma\twoheadrightarrow\mathcal{V}\},

where 𝒱\mathcal{V} is a rank dd locally free sheaf. We define

prσ,d:GrX(σ,d)Xpr_{\sigma,d}:Gr_{X}(\sigma,d)\to X

as the projection morphism.

Remark 2.9.

It is notable that we always have GrX(σ,0)XGr_{X}(\sigma,0)\cong X.

The notion of quasi-smooth morphisms and schemes generalizes locally complete intersection morphisms and schemes in derived algebraic geometry:

Definition 2.10.

A morphism of derived schemes f:XYf:X\to Y is defined to be quasi-smooth if the cotangent complex Lf:=LX/YL_{f}:=L_{X/Y} has Tor-amplitude [0,1][0,1]. A derived scheme XX is defined to be quasi-smooth if the morphism from XX to Spec()\mathrm{Spec}(\mathbb{C}) is quasi-smooth. For a quasi-smooth morphism ff (resp. scheme XX), we define its virtual dimension, denoted

vdim(f),(resp. vdim(X)),vdim(f),\quad\text{(resp. }vdim(X)),

as the rank of the cotangent complex LfL_{f} (resp. LXL_{X}). The relative canonical bundle KfK_{f} (resp. KXK_{X}) is defined as the determinant of LfL_{f} (resp. LXL_{X}). (We refer to [31] or [15] for details on the determinant of a Tor-amplitude [0,1][0,1] complex.)

Theorem 2.11 ([14]).

The derived Grassmannian GrX(σ,d)Gr_{X}(\sigma,d) is a proper derived scheme over XX, and the projection map

prσ,d:GrX(σ,d)Xpr_{\sigma,d}:Gr_{X}(\sigma,d)\to X

is quasi-smooth, with relative virtual dimension d(ed)d(e-d).

We observe that if σ\sigma has Tor-amplitude [0,1][0,1], then the shifted dual σ[1]\sigma^{\vee}[1] also has Tor-amplitude [0,1][0,1].

Definition 2.12 (Incidence Correspondence, Definition 2.7 of [16]).

Given (d+,d)02(d_{+},d_{-})\in\mathbb{Z}_{\geq 0}^{2}, the incidence scheme IncX(σ,d+,d)\mathrm{Inc}_{X}(\sigma,d_{+},d_{-}) can be defined in two ways, which were proven equivalent in [14]:

  1. (1)

    Over the projection morphism pr+:GrX(σ,d+)Xpr_{+}:Gr_{X}(\sigma,d_{+})\to X, let σ+\sigma_{+} be the fiber of the universal quotient:

    θ+:pr+σ𝒱+.\theta_{+}:pr^{*}_{+}\sigma\to\mathcal{V}^{+}.

    We define

    IncX(σ,d+,d):=GrGrX(σ,d+)(σ+[1],d).\mathrm{Inc}_{X}(\sigma,d_{+},d_{-}):=Gr_{Gr_{X}(\sigma,d_{+})}(\sigma_{+}^{\vee}[1],d_{-}).
  2. (2)

    Over the projection morphism pr:GrX(σ[1],d)Xpr_{-}:Gr_{X}(\sigma^{\vee}[1],d_{-})\to X, let σ\sigma_{-} be the fiber of the universal quotient:

    θ:prσ[1]𝒱.\theta_{-}:pr^{*}_{-}\sigma^{\vee}[1]\to\mathcal{V}^{-\vee}.

    We define

    (2.1) IncX(σ,d+,d):=GrGrX(σ[1],d)(σ[1],d+).\mathrm{Inc}_{X}(\sigma,d_{+},d_{-}):=Gr_{Gr_{X}(\sigma^{\vee}[1],d_{-})}(\sigma_{-}^{\vee}[1],d_{+}).

The above definitions induce canonical projection morphisms:

rd+,d+:IncX(σ,d+,d)GrX(σ,d+),rd+,d:IncX(σ,d+,d)GrX(σ[1],d),r^{+}_{d_{+},d_{-}}:\mathrm{Inc}_{X}(\sigma,d_{+},d_{-})\to Gr_{X}(\sigma,d_{+}),\quad r^{-}_{d_{+},d_{-}}:\mathrm{Inc}_{X}(\sigma,d_{+},d_{-})\to Gr_{X}(\sigma^{\vee}[-1],d_{-}),

which we will abbreviate as r+r^{+} and rr^{-} respectively when context allows.

2.3. Degeneracy Loci and the Classicality of Incidence Varieties

In this subsection, we show the relation between the classicality of incidence varieties and the goodness of the degeneracy loci:

Theorem 2.13.

Assuming that e:=rank(σ)0e:=\mathrm{rank}(\sigma)\geq 0, the following are equivalent:

  1. (1)

    σ\sigma has a good (resp. perfect) degeneracy loci.

  2. (2)

    For any integer d+,d0d_{+},d_{-}\geq 0, the incidence variety

    IncX(σ,d+,d)\mathrm{Inc}_{X}(\sigma,d_{+},d_{-})

    is classical (resp. and irreducible).

  3. (3)

    For any integer d+0d_{+}\geq 0, the derived Grassmannian

    GrX(σ,d+)Gr_{X}(\sigma,d_{+})

    is classical (resp. and irreducible).

  4. (4)

    For any integer d0d_{-}\geq 0, the derived Grassmannian

    GrX(σ[1],d)Gr_{X}(\sigma^{\vee}[1],d_{-})

    is classical (resp. and irreducible).

Theorem 2.13 follows from Lemma 2.5 and the following two propositions:

Proposition 2.14.

Assuming e:=rank(σ)0e:=\mathrm{rank}(\sigma)\geq 0, if σ\sigma has a good (resp. perfect) degeneracy loci, the incidence variety

IncX(σ,d+,d)\mathrm{Inc}_{X}(\sigma,d_{+},d_{-})

is classical (resp. and irreducible).

Proposition 2.15.

If for any d0d\geq 0, the derived Grassmannian

GrX(σ,d)Gr_{X}(\sigma,d)

is classical (resp. and irreducible), then σ\sigma has good (resp. perfect) degeneracy loci.

The key point of Propositions 2.14 and 2.15 is the following lemma:

Lemma 2.16.

Over Dσ,lD_{\sigma,l}, the underlying map of the projection from IncX(σ,d+,d)\mathrm{Inc}_{X}(\sigma,d_{+},d_{-}) to XX is a Gr(l,d+)×Gr(le,d)Gr(l,d_{+})\times Gr(l-e,d_{-}) bundle.

Proof.

This directly follows from the definition of Dσ,lD_{\sigma,l}. ∎

Proof of Proposition 2.14.

Let σ\sigma be a good degeneracy loci of rank e0e\geq 0 and tt be the projection from the incidence variety to XX. Then the underlying scheme of the preimage over Dσ,lD_{\sigma,l} is empty if l<max{d+,e+d}l<\max\{d_{+},e+d_{-}\} and has dimension

dim(X)codim(Dσ,l)+d+(ld+)+d(led)\displaystyle\dim(X)-\mathrm{codim}(D_{\sigma,l})+d_{+}(l-d_{+})+d_{-}(l-e-d_{-})
=dim(X)l(le)+d+(ld+)+d(led)\displaystyle=\dim(X)-l(l-e)+d_{+}(l-d_{+})+d_{-}(l-e-d_{-})
=vdim(IncX(σ,d+,d))(ld+)(lde)vdim(IncX(σ,d+,d))\displaystyle=\mathrm{vdim}(\mathrm{Inc}_{X}(\sigma,d_{+},d_{-}))-(l-d_{+})(l-d_{-}-e)\leq\mathrm{vdim}(\mathrm{Inc}_{X}(\sigma,d_{+},d_{-}))

if lmax{d+,e+d}l\geq\max\{d_{+},e+d_{-}\}. Moreover, equality holds if and only if l=max{d+,e+d}l=\max\{d_{+},e+d_{-}\}. Hence, if σ\sigma has good degeneracy loci, we have

dim(π0(IncX(σ,d+,d)))=vdim(IncX(σ,d+,d)),\dim(\pi_{0}(\mathrm{Inc}_{X}(\sigma,d_{+},d_{-})))=\mathrm{vdim}(\mathrm{Inc}_{X}(\sigma,d_{+},d_{-})),

and thus it is classical. Moreover, if σ\sigma has perfect degeneracy loci, IncX(σ,d+,d)\mathrm{Inc}_{X}(\sigma,d_{+},d_{-}) has only one irreducible component of maximal dimension and is therefore also irreducible. ∎

Proof of Proposition 2.15.

For any dmax{0,e}d\geq\max\{0,e\}, the underlying scheme map of prσ,dpr_{\sigma,d} over Dσ,dD_{\sigma,d} is an isomorphism. Thus, if the underlying scheme of the derived Grassmannian

GrX(σ,d)Gr_{X}(\sigma,d)

is irreducible, so is Dσ,dD_{\sigma,d}. Moreover, if the derived Grassmannian is classical, then

dim(X)codim(Dσ,d)=dim(Dσ,d)vdim(GrX(σ,d))=dim(X)d(de),\dim(X)-\mathrm{codim}(D_{\sigma,d})=\dim(D_{\sigma,d})\leq\mathrm{vdim}(Gr_{X}(\sigma,d))=\dim(X)-d(d-e),

and hence codim(Dσ,d)=d(de)\mathrm{codim}(D_{\sigma,d})=d(d-e) by Proposition 2.3. Thus, by Lemma 2.5, σ\sigma has a good degeneracy loci and a perfect degeneracy loci if all the derived Grassmannians are irreducible. ∎

2.4. Birational Geometry of Incidence Varieties for Perfect Degeneracy Loci

In this subsection, we consider the birational geometry of incidence varieties. By Lemma 2.16, we have the following result:

Proposition 2.17 (Proposition 1.11).

Let σ\sigma be a tor-amplitude [0,1][0,1] complex of rank e0e\geq 0 with perfect degeneracy loci. Then, for integers d+,d0d_{+},d_{-}\geq 0 such that d+=d+ed_{+}=d_{-}+e, the morphisms

r+:IncX(σ,d+,d)\displaystyle r^{+}:\mathrm{Inc}_{X}(\sigma,d_{+},d_{-}) GrX(σ,d+),\displaystyle\to Gr_{X}(\sigma,d_{+}),
r:IncX(σ,d+,d)\displaystyle r^{-}:\mathrm{Inc}_{X}(\sigma,d_{+},d_{-}) GrX(σ[1],d),\displaystyle\to Gr_{X}(\sigma^{\vee}[1],d_{-}),
prσ,d+:GrX(σ,d+)\displaystyle pr_{\sigma,d_{+}}:Gr_{X}(\sigma,d_{+}) D¯σ,d+,\displaystyle\to\bar{D}_{\sigma,d_{+}},
prσ,d:GrX(σ[1],d)\displaystyle pr_{\sigma,d_{-}}:Gr_{X}(\sigma^{\vee}[1],d_{-}) D¯σ,d+\displaystyle\to\bar{D}_{\sigma,d_{+}}

are birational. Furthermore, all of the above morphisms are isomorphisms if

d+ddim(X)<(d++1)(d+1).d_{+}d_{-}\leq\dim(X)<(d_{+}+1)(d_{-}+1).
Proof.

We only need to observe that the projections

prσ,d+,prσ[1],dpr_{\sigma,d_{+}},\quad pr_{\sigma^{\vee}[1],d_{-}}

are isomorphisms over Dσ,d+D_{\sigma,d_{+}}. Moreover, when

d+ddim(X)<(d++1)(d+1),d_{+}d_{-}\leq\dim(X)<(d_{+}+1)(d_{-}+1),

the degeneracy locus D¯σ,d++1\bar{D}_{\sigma,d_{+}+1} is empty, since its dimension is negative. ∎

3. Degeneracy Loci on the Smooth Moduli Surface of Coherent Sheaves over a Surface

Let SS be a smooth projective surface over \mathbb{C}. Fix an ample line bundle HH, an integer r>0r\in\mathbb{Z}_{>0}, and a class c1H2(S,)c_{1}\in H^{2}(S,\mathbb{Z}). Given nH4(S,)n\in H^{4}(S,\mathbb{Z})\cong\mathbb{Z}, we denote by

Hss(r,c1,n)andHs(r,c1,n)\mathcal{M}_{H}^{ss}(r,c_{1},n)\quad\text{and}\quad\mathcal{M}_{H}^{s}(r,c_{1},n)

the moduli spaces of Gieseker HH-semistable (resp. stable) sheaves on SS with Chern classes (r,c1,n)(r,c_{1},n). Let

𝒰Coh(Hs(r,c1,n)×S)\mathcal{U}\in\mathrm{Coh}(\mathcal{M}_{H}^{s}(r,c_{1},n)\times S)

be the universal coherent sheaf.

In this section, we prove Theorem 1.1 and Theorem 1.2.

3.1. A General Proposition

We first state two general propositions:

Proposition 3.1.

Let {Xm}m\{X_{m}\}_{m\in\mathbb{Z}} be a family of irreducible smooth varieties over \mathbb{C} with constant integers const and r>0r>0 such that for each mm\in\mathbb{Z}, XmX_{m} is either empty or an irreducible smooth variety of dimension

const+2rm.const+2rm.

Let 𝒱:={𝒱mPerf(Xm)}\mathcal{V}:=\{\mathcal{V}_{m}\in\text{Perf}(X_{m})\} be a family of tor-amplitude [0,1][0,1] and fixed rank rr complexes such that for all m,lm,l,

GrXm+l(𝒱m+l[1],l)GrXm(𝒱m,l)Gr_{X_{m+l}}(\mathcal{V}^{\vee}_{m+l}[1],l)\cong Gr_{X_{m}}(\mathcal{V}_{m},l)

are classical irreducible varieties. Then the complex 𝒱\mathcal{V} has perfect degeneracy loci for any XmX_{m}.

Proof.

This follows from Theorem 2.13. ∎

Proposition 3.2.

Assuming the setting of Proposition 3.1, if there exists an integer θ0\theta_{0} such that XmX_{m} is empty if and only if

const+2rmθ0,const+2rm\geq\theta_{0},

then the degeneracy locus

D¯𝒱m,l\bar{D}_{\mathcal{V}_{m},l}\neq\emptyset

if and only if the expected dimension

const+2rm+l(rl)t(rt)θ0,const+2rm+l(r-l)-t(r-t)\geq\theta_{0},

where tt is the remainder of ll divided by rr.

Proof.

We make induction on ll. When l<rl<r, the left side of the inequality is the dimension of XmX_{m}, and D¯𝒱m,l\bar{D}_{\mathcal{V}_{m},l} is not empty if and only if XmX_{m} is not empty. Thus, the statement holds in this case.

When lrl\geq r, we have

D¯𝒱m,l=D¯𝒱m[1],lr,\bar{D}_{\mathcal{V}_{m},l}=\bar{D}_{\mathcal{V}_{m}^{\vee}[1],l-r},

which is birational to GrXml+r(𝒱,lr)Gr_{X_{m-l+r}}(\mathcal{V},l-r). Therefore, D¯𝒱m,l\bar{D}_{\mathcal{V}_{m},l} is not empty if and only if D¯𝒱ml+r,lr\bar{D}_{\mathcal{V}_{m-l+r},l-r} is not empty. By induction, this holds if and only if

const+2r(ml+r)+(lr)(2rl)t(rt)θ0.const+2r(m-l+r)+(l-r)(2r-l)-t(r-t)\geq\theta_{0}.

Simplifying the right-hand side, we get

const+2rm+l(rl)t(rt)θ0.const+2rm+l(r-l)-t(r-t)\geq\theta_{0}.

Corollary 3.3.

Let

vdimm,l:=dim(Xm)+max{0,l(rl)}.vdim_{m,l}:=\dim(X_{m})+\max\{0,l(r-l)\}.

Then the degeneracy locus

D¯𝒱m,l\bar{D}_{\mathcal{V}_{m},l}\neq\emptyset

if and only if XmX_{m}\neq\emptyset and

vdimm,lθ0+t(rt),vdim_{m,l}\geq\theta_{0}+t(r-t),

where tt is the remainder of ll divided by rr.

3.2. Moduli Space of Framed Coherent Sheaves on the Plane and ADHM Construction

Given integers (r,n)>0×0(r,n)\in\mathbb{Z}_{>0}\times\mathbb{Z}_{\geq 0}, we denote by

fr(r,n)\mathcal{M}_{fr}(r,n)

the moduli space of framed sheaves on the projective plane. This moduli space can also be interpreted as the quiver variety of the Jordan quiver. Specifically, we consider two vector spaces VV and WW of dimensions nn and rr, respectively. The ADHM datum for this setup is given by

Ms(V,W):={(X,Y,i,j)Hom(V,V)Hom(V,V)Hom(W,V)Hom(V,W)}M^{s}(V,W):=\{(X,Y,i,j)\in\operatorname{Hom}(V,V)\oplus\operatorname{Hom}(V,V)\oplus\operatorname{Hom}(W,V)\oplus\operatorname{Hom}(V,W)\}

subject to the equation

[X,Y]+ij=0,[X,Y]+ij=0,

along with the stability condition that for any subspace VVV^{\prime}\subset V closed under XX and YY and in the kernel of jj, we must have V=0V^{\prime}=0.

Theorem 3.4 (Chapter 2 of [23]).
fr(r,n)Ms(V,W)/GL(V),\mathcal{M}_{fr}(r,n)\cong M^{s}(V,W)/GL(V),

where the general linear group GL(V)GL(V) acts by conjugation. Moreover, fr(r,n)\mathcal{M}_{fr}(r,n) is a non-empty smooth variety of dimension 2rn2rn.

The vector spaces VV and WW induce tautological locally free sheaves, denoted by 𝒱\mathcal{V} and 𝒲\mathcal{W}. The universal sheaf 𝒰\mathcal{U} on fr(r,n)×𝔸2\mathcal{M}_{fr}(r,n)\times\mathbb{A}^{2} is represented by the two-term complex

𝒱(Xxid)(Yyid)j𝒱𝒱𝒲(Yyid)(Xxid)i𝒱,\mathcal{V}\xrightarrow{(X-x\cdot\text{id})\oplus(Y-y\cdot\text{id})\oplus j}\mathcal{V}\oplus\mathcal{V}\oplus\mathcal{W}\xrightarrow{(Y-y\cdot\text{id})\oplus-(X-x\cdot\text{id})\oplus i}\mathcal{V},

for (x,y)𝔸2(x,y)\in\mathbb{A}^{2}. This is a two-term complex because the map

(Xxid)(Yyid)j(X-x\cdot\text{id})\oplus(Y-y\cdot\text{id})\oplus j

is injective at any closed point by the stability condition. We denote by 𝒰0\mathcal{U}_{0} the restriction of 𝒰\mathcal{U} to fr(r,n)×{(0,0)}\mathcal{M}_{fr}(r,n)\times\{(0,0)\}, i.e., setting x=y=0x=y=0.

The derived Grassmannian of 𝒰\mathcal{U} is described by the perverse coherent sheaves of Nakajima-Yoshioka [25], which also admit a quiver-like description. Let V1V_{1}, V2V_{2}, and WW be vector spaces of dimensions n1n_{1}, n2n_{2}, and rr, respectively. The datum Ms(V1,V2,W)M^{s}(V_{1},V_{2},W) includes elements (B1,B2,d,i,j)(B_{1},B_{2},d,i,j) in

Hom(V1,V2)Hom(V2,V1)Hom(W,V1)Hom(V2,W),\operatorname{Hom}(V_{1},V_{2})\oplus\operatorname{Hom}(V_{2},V_{1})\oplus\operatorname{Hom}(W,V_{1})\oplus\operatorname{Hom}(V_{2},W),

satisfying the equation

B1dB2B2dB1=0,B_{1}dB_{2}-B_{2}dB_{1}=0,

and the stability condition that for any subspaces

V1V1,V2V2,0WV_{1}^{\prime}\subset V_{1},\quad V_{2}^{\prime}\subset V_{2},\quad 0\subset W

which are closed under the quiver representation, we must have V1=0V_{1}^{\prime}=0. By the stability condition, dd is injective, implying n1n2n_{1}\geq n_{2}.

Theorem 3.5 (Nakajima-Yoshioka [25]).

The moduli space

perv(n1,n2,r):=Ms(V1,V2,W)/(GL(V1)×GL(V2))\mathcal{M}_{\text{perv}}(n_{1},n_{2},r):=M^{s}(V_{1},V_{2},W)/(GL(V_{1})\times GL(V_{2}))

is either an irreducible smooth variety of dimension 2n1n2n12n22+(n1+n2)r2n_{1}n_{2}-n_{1}^{2}-n_{2}^{2}+(n_{1}+n_{2})r or an empty set.

We define the morphisms

ζ:perv(n1,n2,r)(n1,r),\displaystyle\zeta:\mathcal{M}_{\text{perv}}(n_{1},n_{2},r)\to\mathcal{M}(n_{1},r),
(B1,B2,d,i,j)(dB1,dB2,i,dj),\displaystyle(B_{1},B_{2},d,i,j)\mapsto(dB_{1},dB_{2},i,dj),

and

η:perv(n1,n2,r)(n2,r),\displaystyle\eta:\mathcal{M}_{\text{perv}}(n_{1},n_{2},r)\to\mathcal{M}(n_{2},r),
(B1,B2,d,i,j)(B1d,B2d,id,j),\displaystyle(B_{1},B_{2},d,i,j)\mapsto(B_{1}d,B_{2}d,id,j),

which are well-defined by [25].

Theorem 3.6 ([25]).

We have

perv(n1,n2,r)\displaystyle\mathcal{M}_{\text{perv}}(n_{1},n_{2},r) Grfr(n1,r)(𝒰0[1],n1n2),\displaystyle\cong\operatorname{Gr}_{\mathcal{M}_{fr}(n_{1},r)}(\mathcal{U}_{0}^{\vee}[1],n_{1}-n_{2}),
perv(n1,n2,r)\displaystyle\mathcal{M}_{\text{perv}}(n_{1},n_{2},r) Grfr(n2,r)(𝒰0,n1n2),\displaystyle\cong\operatorname{Gr}_{\mathcal{M}_{fr}(n_{2},r)}(\mathcal{U}_{0},n_{1}-n_{2}),

where ζ\zeta and η\eta are the projection morphisms.

Applying Theorem 3.6, we conclude:

Corollary 3.7.

Fixing r>0r>0, the sheaf 𝒰0nCoh(fr(r,n))\mathcal{U}_{0}\in\coprod_{n}\operatorname{Coh}(\mathcal{M}_{fr}(r,n)) satisfies the assumptions in Proposition 3.1, with

const=θ0=0.const=\theta_{0}=0.

In particular, 𝒰0nCoh(fr(r,n))\mathcal{U}_{0}\in\coprod_{n}\operatorname{Coh}(\mathcal{M}_{fr}(r,n)) and 𝒰nCoh(fr(r,n))×𝔸2\mathcal{U}\in\coprod_{n}\operatorname{Coh}(\mathcal{M}_{fr}(r,n))\times\mathbb{A}^{2} both have perfect degeneracy loci. The degeneracy locus

D¯𝒰0,landD¯𝒰,l\bar{D}_{\mathcal{U}_{0},l}\quad\text{and}\quad\bar{D}_{\mathcal{U},l}

is non-empty if and only if

2rn+l(rl)t(rt)0,2rn+l(r-l)-t(r-t)\geq 0,

where tt is the remainder when ll is divided by rr.

3.3. Moduli Space of Gieseker Stable Sheaves

The geometry of the moduli space of Gieseker stable coherent sheaves can be complex. To simplify computations, we assume the following conditions:

Assumption 3.8.

We make the following assumptions:

  1. (1)

    The total Chern class is primitive, i.e.,

    gcd(r,c1H)=1.\gcd(r,c_{1}\cdot H)=1.
  2. (2)

    Either r=1r=1, or KS𝒪SK_{S}\cong\mathcal{O}_{S}, or c1(KS)H<0c_{1}(K_{S})\cdot H<0.

  3. (3)

    There exists a semistable coherent sheaf \mathcal{F} of rank rr with first Chern character c1c_{1}, i.e.,

    nHss(r,c1,n).\coprod_{n\in\mathbb{Z}}\mathcal{M}^{ss}_{H}(r,c_{1},n)\neq\emptyset.
Remark 3.9.

3.8 holds for generic ample divisors when the surface SS is a Del Pezzo, K3, or abelian surface.

We state the following general results concerning the moduli space Hs(r,c1,n)\mathcal{M}_{H}^{s}(r,c_{1},n):

Proposition 3.10 ([13]; see also Proposition 2.10 of [28] and Lemma 3.3 of [19]).

If 3.8 holds, then

Hss(r,c1,n)=Hs(r,c1,n),\mathcal{M}_{H}^{ss}(r,c_{1},n)=\mathcal{M}_{H}^{s}(r,c_{1},n),

which is either empty or a smooth projective variety of dimension

dimHs(r,c1,n)=const+2rn,\dim\mathcal{M}_{H}^{s}(r,c_{1},n)=const+2rn,

where

const:=(1r)c12(r21)χ(𝒪S)+h1(𝒪S).const:=(1-r)c_{1}^{2}-(r^{2}-1)\chi(\mathcal{O}_{S})+h^{1}(\mathcal{O}_{S}).
Proposition 3.11 (Proposition 2.14 of [28]).

Assuming 3.8, the universal sheaf 𝒰\mathcal{U}, regarded as an object in Perf(Hs(r,c1,n)×S)\mathrm{Perf}(\mathcal{M}_{H}^{s}(r,c_{1},n)\times S), has tor-amplitude in [0,1][0,1].

Lemma 3.12 (Proposition 5.5 of [27]).

Under 3.8, for any non-splitting short exact sequence

0𝒰𝒰x0,0\to\mathcal{U}^{\prime}\to\mathcal{U}\to\mathbb{C}_{x}\to 0,

where xSx\in S and both 𝒰\mathcal{U}^{\prime} and 𝒰\mathcal{U} have rank rr and first Chern character c1c_{1}, 𝒰\mathcal{U} is stable if and only if 𝒰\mathcal{U}^{\prime} is stable.

Fixing rr and c1c_{1}, the Bogomolov inequality implies:

Hss(r,c1,N)=forN0.\mathcal{M}^{ss}_{H}(r,c_{1},N)=\emptyset\quad\text{for}\quad N\ll 0.

Thus, there exists a smallest integer N0N_{0} such that:

Hss(r,c1,N0).\mathcal{M}^{ss}_{H}(r,c_{1},N_{0})\neq\emptyset.

By Lemma 3.12, we have:

Lemma 3.13.

For every nN0n\geq N_{0},

Hss(r,c1,n).\mathcal{M}^{ss}_{H}(r,c_{1},n)\neq\emptyset.

Let θ0:=const+2rN0\theta_{0}:=const+2rN_{0}. Yoshioka has shown the following results:

Theorem 3.14 (Theorem 0.1 of [33]; Theorem 0.1 of [34]).

When SS is a K3 surface and HH is a generic ample divisor, θ0=0\theta_{0}=0. When SS is an abelian surface, θ0=2\theta_{0}=2.

Remark 3.15.

When SS is 2\mathbb{P}^{2} or 1×1\mathbb{P}^{1}\times\mathbb{P}^{1}, the constant θ0\theta_{0} was computed by Drezet-Le Potier in [9].

3.4. Perverse coherent sheaves on blow-ups

Now recall the moduli space of perverse coherent sheaves of Nakajima-Yoshioka [25, 24, 26]. Let oSo\in S be a closed point and S^:=BloS\hat{S}:=Bl_{o}S. We denote p:S^Sp:\hat{S}\to S as the projection map and C:=p1(o)1C:=p^{-1}(o)\cong\mathbb{P}^{1} as the exceptional divisor. We denote

𝒪C(m):=𝒪(mC)|C\mathcal{O}_{C}(m):=\mathcal{O}(-mC)|_{C}

which is the unique degree mm line bundle on the rational curve CC for any mm\in\mathbb{Z}.

Definition 3.16.

A stable perverse coherent sheaf EE on S^\hat{S} with respect to HH is a coherent sheaf such that

Hom(E,𝒪C(1))=0,Hom(𝒪C,E)=0Hom(E,\mathcal{O}_{C}(-1))=0,\quad Hom(\mathcal{O}_{C},E)=0

and pEp_{*}E is Gieseker stable with respect to HH.

Given the decomposition of cohomology groups:

H(S^,)=p(H(S^,))[C]H^{*}(\hat{S},\mathbb{Q})=p^{*}(H^{*}(\hat{S},\mathbb{Q}))\oplus\mathbb{Q}[C]

where [C]H1,1(S^)[C]\in H^{1,1}(\hat{S}) and [C]2=1[C]^{2}=-1, we define a morphism

p!:H(S^,)H(S,),p!(c):=p(ctd(S^))td(S)1.p_{!}:H^{*}(\hat{S},\mathbb{Q})\to H^{*}(S,\mathbb{Q}),\quad p_{!}(c):=p_{*}(c\ td(\hat{S}))td(S)^{-1}.

The morphism p!p_{!} is compatible with the algebraic KK-theory, i.e. we have the commutative diagram

K0(S^){K_{0}(\hat{S})}K0(S){K_{0}(S)}H(S^,){H^{*}(\hat{S},\mathbb{Q})}H(S,){H^{*}(S,\mathbb{Q})}p\scriptstyle{p_{*}}ch\scriptstyle{ch}ch\scriptstyle{ch}p!\scriptstyle{p_{!}}

where chch is the Chern character morphism. Moreover, we will have

p!([S^])=[S],p!([C])=12[pt],p![pt]=pt.p_{!}([\hat{S}])=[S],\quad p_{!}([C])=\frac{1}{2}[pt],\quad p_{!}[pt]=pt.

and

p!p=id:H(S,)H(S,).p_{!}p^{*}=id:H^{*}(S,\mathbb{Q})\to H^{*}(S,\mathbb{Q}).

Fixing a total Chern class c=(r,c1,n)c=(r,c_{1},n) we denote

ch(c):=r+c1+12c12nch(c):=r+c_{1}+\frac{1}{2}c_{1}^{2}-n

which is the Chern chracter of cc.

Definition 3.17.

For any total Chern class cc on SS and ll\in\mathbb{Z}, we define

M0(c,l)M^{0}(c,l)

as the moduli space of perverse coherent sheaves with the total Chern chracter

vd:=pch(c)d(ch(𝒪C(1)))=pch(c)d([C]12[pts]).v_{d}:=p^{*}ch(c)-d(ch(\mathcal{O}_{C}(-1)))=p^{*}ch(c)-d([C]-\frac{1}{2}[pts]).
Proposition 3.18 (Lemma 3.22 of [24] and Lemma 3.3 of [19]).

Assuming 3.8 and 3.8, M0(c,l)M^{0}(c,l) is either empty or a smooth variety of expected dimension

const+2rnr(l+r).const+2rn-r(l+r).
Proposition 3.19 (Proposition 3.1 of [20] and Lemma 3.4 of [24]).

For a stable coherent sheaf EE on S^\hat{S}, we have

RipE(aC)=0,a=0,1,i1,R^{i}p_{*}E(aC)=0,\quad a=0,1,\quad i\geq 1,

and moreover,

RpE(aC)=pE(aC),a=0,1,Rp_{*}E(aC)=p_{*}E(aC),\quad a=0,1,

are all stable.

By Proposition 3.19, there exist canonical morphisms

ζ:M0(c,l)MHs(c),\displaystyle\zeta:M^{0}(c,l)\to M^{s}_{H}(c), EpE,\displaystyle E\to p_{*}E,
η:M0(c,l)MHs(cl[pts]),\displaystyle\eta:M^{0}(c,l)\to M^{s}_{H}(c-l[\mathrm{pts}]), EpE(C).\displaystyle E\to p_{*}E(C).

We denote

𝒰oCoh(Hs(c)):=𝒰|Hs(c)×o.\mathcal{U}_{o}\in\mathrm{Coh}(\mathcal{M}^{s}_{H}(c)):=\mathcal{U}|_{\mathcal{M}^{s}_{H}(c)\times o}.
Theorem 3.20 (Theorem 4.1 of [24], also see Theorem 4.1 of [19]).

We have

M0(c,l)GrHs(c)(𝒰o[1],l),M0(c,l)GrHs(cl[pts])(𝒰o,l)M^{0}(c,l)\cong Gr_{\mathcal{M}^{s}_{H}(c)}(\mathcal{U}^{\vee}_{o}[1],l),\quad M^{0}(c,l)\cong Gr_{\mathcal{M}^{s}_{H}(c-l[pts])}(\mathcal{U}_{o},l)

where the projection morphisms are just ζ\zeta and η\eta respectively.

By Proposition 3.10, Lemma 3.13 and Theorem 3.20, for any oSo\in S, the family Xm:=Hs(r,c1,m)X_{m}:=\mathcal{M}_{H}^{s}(r,c_{1},m) and the coherent sheaf 𝒰o\mathcal{U}_{o} satisfy the conditions in Proposition 3.1 and Corollary 3.3, with the given constconst and θ0\theta_{0}. Hence by Proposition 3.1 and Corollary 3.3 we have

Corollary 3.21 (Theorem 1.1 and Theorem 1.2).

Let θ:=θ0+2\theta:=\theta_{0}+2, and define

vdimn,l:=dim(Hs(r,c1,n))+2+max{0,l(rl)},vdim_{n,l}:=\dim(\mathcal{M}_{H}^{s}(r,c_{1},n))+2+\max\{0,l(r-l)\},

for any nn\in\mathbb{Z} and l0l\geq 0. Then, the degeneracy locus

D¯𝒰n,l\bar{D}_{\mathcal{U}_{n},l}

is either an empty set or an irreducible Cohen-Macaulay variety of dimension vdimn,lvdim_{n,l}. Moreover, D¯𝒰n,l\bar{D}_{\mathcal{U}_{n},l}\neq\emptyset if and only if Hs(r,c1,n)\mathcal{M}_{H}^{s}(r,c_{1},n)\neq\emptyset and

vdimn,lθ+t(rt),vdim_{n,l}\geq\theta+t(r-t),

where tt is the remainder of ll divided by rr.

References

  • [1] Nicolas Addington and Ryan Takahashi. A categorical sl2sl_{2} action on some moduli spaces of sheaves. Trans. Amer. Math. Soc., 375(12):8969–9005, 2022.
  • [2] Enrico Arbarello, Maurizio Cornalba, Phillip A Griffiths, et al. Geometry of algebraic curves: volume II with a contribution by Joseph Daniel Harris. Springer, 2011.
  • [3] D. Arinkin and D. Gaitsgory. Singular support of coherent sheaves and the geometric langlands conjecture. Selecta Mathematica, 21(1):1–199, Jan 2015.
  • [4] Vladimir Baranovsky. Moduli of sheaves on surfaces and action of the oscillator algebra. Journal of Differential Geometry, 55(2):193–227, 2000.
  • [5] Arend Bayer, Huachen Chen, and Qingyuan Jiang. Brill-Noether theory of Hilbert schemes of points on surfaces. International Mathematics Research Notices, 2024(10):8403–8416, 2024.
  • [6] K. Behrend and B. Fantechi. The intrinsic normal cone. Invent. Math., 128(1):45–88, 1997.
  • [7] A. Bondal and D. Orlov. Derived categories of coherent sheaves. In Proceedings of the International Congress of Mathematicians, Vol. II (Beijing, 2002), pages 47–56. Higher Ed. Press, Beijing, 2002.
  • [8] Ionuţ Ciocan-Fontanine and Mikhail Kapranov. Virtual fundamental classes via dg–manifolds. Geometry & Topology, 13(3):1779–1804, 2009.
  • [9] J.-M. Drezet 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.
  • [10] William Fulton. Intersection theory, volume 2 of Ergebnisse der Mathematik und ihrer Grenzgebiete (3) [Results in Mathematics and Related Areas (3)]. Springer-Verlag, Berlin, 1984.
  • [11] Dmitry Galakhov, Alexei Morozov, and Nikita Tselousov. Super-Schur polynomials for affine super Yangian {Y(gl(1|1))^\widehat{Y(gl(1|1))}}. Journal of High Energy Physics, 2023(8):49, Aug 2023.
  • [12] Tamás Hausel, Anton Mellit, Alexandre Minets, and Olivier Schiffmann. P=WP=W via H2H_{2}. arXiv preprint arXiv:2209.05429, 2022.
  • [13] Daniel Huybrechts and Manfred Lehn. The geometry of moduli spaces of sheaves. Cambridge Mathematical Library. Cambridge University Press, second edition, 2010.
  • [14] Qingyuan Jiang. Derived Grassmannians and derived Schur functors. arXiv preprint arXiv:2212.10488, 2022.
  • [15] Qingyuan Jiang. Derived projectivizations of complexes. arXiv preprint arXiv:2202.11636, 2022.
  • [16] Qingyuan Jiang. Derived categories of derived Grassmannians. arXiv preprint arXiv:2307.02456, 2023.
  • [17] Qingyuan Jiang. Notes on Brill-Noether for Hilbert schemes of points, 2024. https://drive.google.com/file/d/1Y843D5xIzSdXwhZ7CGQkB5UwvL1WJ_Gm/view.
  • [18] Yujiro Kawamata. D-equivalence and K-equivalence. Journal of Differential Geometry, 61(1):147–171, 2002.
  • [19] Naoki Koseki. Categorical blow-up formula for Hilbert schemes of points. arXiv preprint arXiv:2110.08315, 2021.
  • [20] Nikolas Kuhn and Yuuji Tanaka. A blowup formula for virtual enumerative invariants on projective surfaces. arXiv preprint arXiv:2107.08155, 2021.
  • [21] Jun Li and Gang Tian. Virtual moduli cycles and Gromov-Witten invariants of algebraic varieties. J. Amer. Math. Soc., 11(1):119–174, 1998.
  • [22] Anton Mellit, Alexandre Minets, Olivier Schiffmann, and Eric Vasserot. Coherent sheaves on surfaces, COHAs and deformed W1+W_{1+\infty}-algebras. arXiv preprint arXiv:2311.13415, 2023.
  • [23] Hiraku Nakajima. Lectures on Hilbert schemes of points on surfaces, volume 18 of University Lecture Series. 1999.
  • [24] Hiraku Nakajima and Kota Yoshioka. Perverse coherent sheaves on blow-up. II. wall-crossing and Betti numbers formula. Journal of Algebraic Geometry, 20(1):47–100, 03 2010.
  • [25] Hiraku Nakajima and Kota Yoshioka. Perverse coherent sheaves on blow-up. I. a quiver description. Advanced Studies in Pure Mathematics, 61:349, 2011.
  • [26] Hiraku Nakajima and Kota Yoshioka. Perverse coherent sheaves on blowup, III: Blow-up formula from wall-crossing. Kyoto Journal of Mathematics, 51(2):263–335, 04 2011.
  • [27] Andrei Neguţ. Shuffle algebras associated to surfaces. Selecta Mathematica. New Series, 25(3):57, 2019.
  • [28] Andrei Neguţ. Hecke correspondences for smooth moduli spaces of sheaves. Publications mathématiques de l’IHÉS, 135(1):337–418, Jun 2022.
  • [29] Miroslav Rapcak, Yan Soibelman, Yaping Yang, and Gufang Zhao. Cohomological Hall algebras and perverse coherent sheaves on toric Calabi-Yau 3-folds. arXiv preprint arXiv:2007.13365, 2020.
  • [30] Olivier Schiffmann and Eric Vasserot. Cherednik algebras, W-algebras and the equivariant cohomology of the moduli space of instantons on 𝔸2\mathbb{A}^{2}. Publications Mathématiques. Institut de Hautes Études Scientifiques, 118:213–342, 2013.
  • [31] Timo Schürg, Bertrand Toën, and Gabriele Vezzosi. Derived algebraic geometry, determinants of perfect complexes, and applications to obstruction theories for maps and complexes. Journal für die reine und angewandte Mathematik (Crelles Journal), 2015(702):1–40, 2015.
  • [32] Yukinobu Toda. Derived categories of Quot schemes of locally free quotients via categorified Hall products. Math. Res. Lett., 30(1):239–265, 2023.
  • [33] Kota Yoshioka. Irreducibility of moduli spaces of vector bundles on K3 surfaces. arXiv preprint math/9907001, 1999.
  • [34] Kota Yoshioka. Moduli spaces of stable sheaves on abelian surfaces. Mathematische Annalen, 321:817–884, 2001.
  • [35] Yu Zhao. Moduli space of sheaves and categorified commutator of functors. arXiv preprint arXiv:2112.12434, 2021.
  • [36] Yu Zhao. A categorical quantum toroidal action on the hilbert schemes. Journal of the Institute of Mathematics of Jussieu, page 1–44, 2023.
  • [37] Yu Zhao. Hilbert schemes on blowing ups and the free Boson, 2024.