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

Moduli spaces of modules over even Clifford algebras and Prym varieties

Jia Choon Lee Peking University, Beijing International Center for Mathematical Research, Jingchunyuan Courtyard #78, 5 Yiheyuan Road, Haidian District, Beijing 100871, China [email protected]
Abstract.

A conic fibration has an associated sheaf of even Clifford algebras on the base. In this paper, we study the relation between the moduli spaces of modules over the sheaf of even Clifford algebras and the Prym variety associated to the conic fibration. In particular, we construct a rational map from the moduli space of modules over the sheaf of even Clifford algebras to the special subvarieties in the Prym variety, and check that the rational map is birational in some cases. As an application, we get an explicit correspondence between instanton bundles of minimal charge on cubic threefolds and twisted Higgs bundles on curves.

1. Introduction

It is well-known that a smooth cubic threefold is irrational since the famous work of Clemens and Griffiths [11]. They observed that if a threefold is rational, then its intermediate Jacobian must be isomorphic to a product of Jacobians of curves. The problem is then reduced to comparing the intermediate Jacobians of cubic threefolds with the Jacobians of curves as principally polarized abelian varieties by studying the singularity loci of their theta divisors.

Despite the success in dimension three, the rationality problem for cubic fourfolds is still not well understood. A categorical approach to the problem is to study the bounded derived category Db(Y4)D^{b}(Y_{4}) of a given smooth cubic fourfold Y4Y_{4} and it always admits a semiorthogonal decomposition:

(1) Db(Y4)𝒦u(Y4),𝒪Y4,𝒪Y4(1),𝒪Y4(2)\textrm{D}^{b}(Y_{4})\cong\langle\mathcal{K}u(Y_{4}),\mathcal{O}_{Y_{4}},\mathcal{O}_{Y_{4}}(1),\mathcal{O}_{Y_{4}}(2)\rangle

where 𝒦u(Y4):=𝒪Y4,𝒪Y4(1),𝒪Y4(2)\mathcal{K}u(Y_{4}):=\langle\mathcal{O}_{Y_{4}},\mathcal{O}_{Y_{4}}(1),\mathcal{O}_{Y_{4}}(2)\rangle^{\perp} is now known as the Kuznetsov component. It is conjectured by Kuznetsov [15] that a smooth cubic fourfold is rational if and only if the Kuznetsov component 𝒦u(Y4)\mathcal{K}u(Y_{4}) is equivalent to the category of a K3 surface. While the conjecture has been checked to hold in some cases, the general conjecture remains unsolved.

Since 𝒦u(Y4)\mathcal{K}u(Y_{4}) is expected to capture the geometry of Y4Y_{4}, an attempt to extract information out of the triangulated category 𝒦u(Y4)\mathcal{K}u(Y_{4}) is to construct Bridgeland stability conditions on 𝒦u(Y4)\mathcal{K}u(Y_{4}) and consider their moduli spaces of stable objects [4][3]. In dimension three, one can define in the same way 𝒦u(Y3):=𝒪Y3,𝒪Y3(1)\mathcal{K}u(Y_{3}):=\langle\mathcal{O}_{Y_{3}},\mathcal{O}_{Y_{3}}(1)\rangle^{\perp} for a smooth cubic threefold Y3Y_{3} and it can be shown that 𝒦u(Y3)\mathcal{K}u(Y_{3}) reconstructs the Fano surface of lines of Y3Y_{3} as a moduli space of stable objects with suitable stability conditions. The reconstruction of the Fano surface of lines then determines the intermediate Jacobian J(Y3)J(Y_{3}) [8]. Alternatively, it is observed that instanton bundles of minimal charge on Y3Y_{3} are objects in 𝒦u(Y3)\mathcal{K}u(Y_{3}). Then by the work of Markushevich-Tikhomirov and others [19][13][12][6], it is shown that the moduli space of instanton bundles of minimal charge on Y3Y_{3} is birational to the intermediate Jacobian J(Y3)J(Y_{3}). So the Kuznetsov components 𝒦u(Y3)\mathcal{K}u(Y_{3}) and 𝒦u(Y4)\mathcal{K}u(Y_{4}) can be thought of as the categorical counterparts of the intermediate Jacobian whose success in the rationality problem of cubic threefolds fits well into the philosophy of Kuznetsov’s conjecture in n=4.n=4.

On the other hand, we know that both Y3Y_{3} and Y4Y_{4} admit a conic bundle structure. More precisely, for n=3,4n=3,4, let l0Ynl_{0}\subset Y_{n} be a line that is not contained in a plane in Yn.Y_{n}. Then the blow-up Y~n:=Bll0(Yn)Bll0(n+1)\widetilde{Y}_{n}:=Bl_{l_{0}}(Y_{n})\subset Bl_{l_{0}}(\mathbb{P}^{n+1}) of YnY_{n} along l0l_{0} projects to a projective space n1\mathbb{P}^{n-1}, denoted by p:Y~nn1p:\widetilde{Y}_{n}\to\mathbb{P}^{n-1}. The map pp is a conic bundle whose discriminant locus Δn\Delta_{n} is a degree 5 hypersurface. The idea of realizing a cubic hypersuface birationally as a conic bundle can be used to study its rationality. Following the idea of Mumford, it is shown that the intermediate Jacobian J(Y3)J(Y~3)J(Y_{3})\cong J(\widetilde{Y}_{3}) is isomorphic to Prym(Δ~3,Δ3)\textrm{Prym}(\widetilde{\Delta}_{3},\Delta_{3}) where Δ~3\widetilde{\Delta}_{3} is the double cover parametrizing the irreducible components of the degenerate conics over Δ3\Delta_{3}. By analyzing the difference between Prym varieties and the Jacobian of curves as principally polarized abelian varieties, it is again shown that a smooth cubic threefold is irrational.

The conic bundle structure of a cubic hypersurface also provides us information at the level of derived category. A quadratic form on a vector space defines the Clifford algebra which decomposes into the even and odd parts. We can apply the construction of Clifford algebra relatively for the conic bundles Y~n\widetilde{Y}_{n} which is viewed as a family of conics over n1\mathbb{P}^{n-1}, and obtain a sheaf of even Clifford algebras 0\mathcal{B}_{0} on n1\mathbb{P}^{n-1}. The bounded derived category Db(n1,0)\textrm{D}^{b}(\mathbb{P}^{n-1},\mathcal{B}_{0}) of 0\mathcal{B}_{0}-modules appears as a component of the semiorthogonal decomposition of the conic bundle Y~n\widetilde{Y}_{n} [14]:

(2) Db(Y~n)=Db(n1,0),pDb(n1).\textrm{D}^{b}(\widetilde{Y}_{n})=\langle\textrm{D}^{b}(\mathbb{P}^{n-1},\mathcal{B}_{0}),p^{*}\textrm{D}^{b}(\mathbb{P}^{n-1})\rangle.

In the case n=3,4n=3,4, by comparing the semiorthogonal decompositions in (1), (2) and the one for blowing up, one can show that there are embedding functors

(3) Ξn:𝒦u(Yn)Db(n1,0)\Xi_{n}:\mathcal{K}u(Y_{n})\hookrightarrow\textrm{D}^{b}(\mathbb{P}^{n-1},\mathcal{B}_{0})

for n=3,4n=3,4 (see [8], [3]). The functors Ξn\Xi_{n} are useful in the study of 𝒦u(Yn)\mathcal{K}u(Y_{n}). For example, when n=3,4n=3,4, the construction of Bridgeland stability conditions on 𝒦u(Yn)\mathcal{K}u(Y_{n}) carried out in [8][3] uses the embedding functors Ξn\Xi_{n} as one of the key steps. Also, in the work of Lahoz-Macrì-Stellari [17], the functor Ξ3\Xi_{3} is used to provide a birational map between the moduli space of instanton bundles of minimal charge and the moduli space of 0\mathcal{B}_{0}-modules. The moduli space of 0\mathcal{B}_{0}-modules was also first considered in [17].

Motivated by the relations found in the case of cubic threefolds as described above: Prym/intermediate Jacobian, intermediate Jacobian/moduli space of instanton bundles, and instanton bundles/0\mathcal{B}_{0}-modules, it is natural to search for a relation between 0\mathcal{B}_{0}-modules and the Prym varieties. In this paper, we will focus on three dimensional conic fibrations over 2\mathbb{P}^{2} (not necessarily obtained from a cubic threefold) and study the relation between the moduli spaces of 0\mathcal{B}_{0}-modules and the Prym varieties.

Let p:X2p:X\to\mathbb{P}^{2} be a three dimensional conic fibration over 2\mathbb{P}^{2} i.e. a flat quadric fibration of relative dimension 1 over 2\mathbb{P}^{2} with simple degeneration i.e. the degenerate conics cannot be double lines and the discriminant locus Δ\Delta in 2\mathbb{P}^{2} is smooth. Let π:Δ~Δ\pi:\widetilde{\Delta}\to\Delta be the double cover parametrizing the irreducible components of degenerate conics over Δ\Delta. We consider the moduli space 𝔐d,e\mathfrak{M}_{d,e} of semistable 0\mathcal{B}_{0}-modules whose underlying 𝒪2\mathcal{O}_{\mathbb{P}^{2}}-modules have fixed Chern character (0,2d,e)(0,2d,e). This means that the 0\mathcal{B}_{0}-modules are supported on plane curves. The moduli space 𝔐d,e\mathfrak{M}_{d,e} comes with a morphism Υ:𝔐d,e|𝒪2(d)|\Upsilon:\mathfrak{M}_{d,e}\to|\mathcal{O}_{\mathbb{P}^{2}}(d)| defined by sending a 0\mathcal{B}_{0}-module to its schematic (Fitting) support.

On the Prym side, by the work of Welters [23] and Beauville [5], each linear system on Δ\Delta defines a so-called special subvariety in the Prym variety Prym(Δ~,Δ)\textrm{Prym}(\widetilde{\Delta},\Delta) of the étale double cover π:Δ~Δ\pi:\widetilde{\Delta}\to\Delta. We apply the construction to the linear system |Ld|=|𝒪2(d)|Δ||L_{d}|=\lvert\mathcal{O}_{\mathbb{P}^{2}}(d)|_{\Delta}\rvert. For each kk, there is an induced morphism π(k):Δ~(k)Δ(k)\pi^{(k)}:\widetilde{\Delta}^{(k)}\to\Delta^{(k)} between the kk-th symmetric products of Δ~\widetilde{\Delta} and Δ\Delta. As the linear system |Ld||L_{d}| can be considered as a subvariety in Δ(k)\Delta^{(k)} for k=deg(Δ)dk=\textrm{deg}(\Delta)\cdot d, we define the variety of divisors lying over |Ld||L_{d}| as Wd=(π(k))1(|Ld|)Δ~(k)W_{d}=(\pi^{(k)})^{-1}(|L_{d}|)\subset\widetilde{\Delta}^{(k)}. Then the image of WdW_{d} under the Abel-Jacobi map α~:Δ~(k)JkΔ~\widetilde{\alpha}:\widetilde{\Delta}^{(k)}\to J^{k}\widetilde{\Delta} lies in the two components of (the translate of) Prym varieties, which are called the special subvarieties. The variety WdW_{d} consists of two components Wd=Wd0Wd1W_{d}=W_{d}^{0}\cup W_{d}^{1}, each of which maps to |Ld||L_{d}|. For d<deg(Δ)d<\textrm{deg}(\Delta), we have |𝒪2(d)||Ld||\mathcal{O}_{\mathbb{P}^{2}}(d)|\cong|L_{d}|, and we denote by Ud|𝒪2(d)|U_{d}\subset|\mathcal{O}_{\mathbb{P}^{2}}(d)| the open subset of smooth degree dd curves intersecting Δ\Delta transversally.

The main construction in this paper is to construct a morphism over UdU_{d}

(4) Φ:𝔐d,e|UdWd|Ud\Phi:\mathfrak{M}_{d,e}|_{U_{d}}\to W_{d}|_{U_{d}}

for d=1,2,i=0,1d=1,2,i=0,1 and ee\in\mathbb{Z}. Here 𝔐d,e|Ud𝔐d,e\mathfrak{M}_{d,e}|_{U_{d}}\subset\mathfrak{M}_{d,e} is the open subset of 0\mathcal{B}_{0}-modules supported on curves in UdU_{d} and Wd|UdWdW_{d}|_{U_{d}}\subset W_{d} is the open subset consisting of the degree kk divisors in Δ~\widetilde{\Delta} whose images in Δ\Delta represent degree kk reduced divisors in LdL_{d}. Moreover, the morphism Φ\Phi is used to prove the following (see Theorem 4.7):

Theorem 1.1.

Let d=1,2d=1,2 and ee\in\mathbb{Z}.

  1. (1)

    The moduli space 𝔐d,e\mathfrak{M}_{d,e} is birational to one of the two connected components WdiW_{d}^{i} of Wd.W_{d}.

  2. (2)

    If 𝔐d,e\mathfrak{M}_{d,e} is birational to WdiW_{d}^{i}, then 𝔐d,e+1\mathfrak{M}_{d,e+1} is birational to Wd1i.W_{d}^{1-i}. In particular, the birational type of 𝔐d,e\mathfrak{M}_{d,e} only depends on dd and (e(e mod 2)2).

The idea behind the construction of the morphism Φ\Phi is based on a study of the representations of degenerate even Clifford algebras. A 0\mathcal{B}_{0}-module in 𝔐d,e|Ud\mathfrak{M}_{d,e}|_{U_{d}} is supported on a plane curve CC which intersects the discriminant curve Δ\Delta in finitely many points. Then the 0\mathcal{B}_{0}-module restricted to each of these points pCΔp\in C\cap\Delta gives rise to a representation of a degenerate even Clifford algebra, which is in turn shown to be equivalent to a representation of the path algebra associated the quiver

+{+}{-}α\scriptstyle{\alpha}β\scriptstyle{\beta}

with relations αβ=βα=0.\alpha\beta=\beta\alpha=0. Then an analysis of the representations coming from 0\mathcal{B}_{0}-modules reveals that there are natural candidates determining canonically the required lift of the point pCΔp\in C\cap\Delta to Δ~\widetilde{\Delta}, hence an element in Wd|UdW_{d}|_{U_{d}}.

By composing with the Abel-Jacobi map α~:Δ~(k)JkΔ~\widetilde{\alpha}:\widetilde{\Delta}^{(k)}\to J^{k}\widetilde{\Delta} that maps to the varieties JkΔ~J^{k}\widetilde{\Delta} of degree kk invertible sheaves on Σ~\widetilde{\Sigma}, we obtain a rational map (choose a base point in JkΔ~J^{k}\widetilde{\Delta} to identify with J0Δ~J^{0}\widetilde{\Delta})

(5) α~Φ:𝔐d,ePrym(Δ~,Δ)\widetilde{\alpha}\circ\Phi:\mathfrak{M}_{d,e}\dashrightarrow\textrm{Prym}(\widetilde{\Delta},\Delta)

whose image is an open subset of a special subvariety.

Next, we apply the result above to the case of cubic threefolds. In [16] and [17], it is observed that instanton bundles of minimal charge are objects in 𝒦u(Y3)\mathcal{K}u(Y_{3}). The authors use the functor Ξ3:𝒦u(Y3)Db(2,0)\Xi_{3}:\mathcal{K}u(Y_{3})\hookrightarrow\textrm{D}^{b}(\mathbb{P}^{2},\mathcal{B}_{0}) to deduce a birational map between the moduli space 𝔐Y3\mathfrak{M}_{Y_{3}} of instanton bundles of minimal charge on Y3Y_{3} and the moduli space 𝔐2,4\mathfrak{M}_{2,-4} of 0\mathcal{B}_{0}-modules. In this case, the rational map 𝔐2,4Prym(Δ~,Δ)\mathfrak{M}_{2,-4}\dashrightarrow\textrm{Prym}(\widetilde{\Delta},\Delta) actually turns out to be birational. Hence, by composing the birational maps, we get

(6) 𝔐Y3Prym(Δ~,Δ)\mathfrak{M}_{Y_{3}}\dashrightarrow\textrm{Prym}(\widetilde{\Delta},\Delta)

As a point in Prym(Δ~,Δ)\textrm{Prym}(\widetilde{\Delta},\Delta) can be interpreted as a twisted Higgs bundle on Δ\Delta by the spectral correspondence [7], the birational map (6) gives an explicit correspondence between instanton bundles of minimal charge on Y3Y_{3} and twisted Higgs bundles on Δ\Delta.

Moreover, as mentioned above, the moduli space of instanton bundles of minimal charge is birational to the intermediate Jacobian J(Y3)J(Y_{3}), so the birational map (6) gives a modular interpretation of the classical isomorphism J(Y3)Prym(Δ~,Δ)J(Y_{3})\cong\textrm{Prym}(\widetilde{\Delta},\Delta) in terms of instanton bundles of minimal charge and twisted Higgs bundles via 0\mathcal{B}_{0}-modules. From this viewpoint, the embedding functor Ξ3:𝒦u(Y3)Db(2,0)\Xi_{3}:\mathcal{K}u(Y_{3})\hookrightarrow\textrm{D}^{b}(\mathbb{P}^{2},\mathcal{B}_{0}) provides a reinterpretation of the classical isomorphism J(Y3)Prym(Δ~,Δ)J(Y_{3})\cong\textrm{Prym}(\widetilde{\Delta},\Delta).

Philosophically, the result allows us to think of Db(2,0)D^{b}(\mathbb{P}^{2},\mathcal{B}_{0}) as the categorical counterpart of Prym(Δ~,Δ)\textrm{Prym}(\widetilde{\Delta},\Delta) associated to a conic bundle, just as 𝒦u(Y3)\mathcal{K}u(Y_{3}) is the categorical counterpart of J(Y3).J(Y_{3}).

1.1. Convention

Throughout this paper we work over the complex numbers \mathbb{C}. All modules in this paper are assumed to be left modules. For a morphism f:XYf:X\to Y of two spaces (schemes or stacks) and a subspace ZYZ\subset Y, we will denote by X|Z:=X×YZX|_{Z}:=X\times_{Y}Z the fiber product and f|Z:X|ZZf|_{Z}:X|_{Z}\to Z.

1.2. Acknowledgement

This paper is written as part of my Ph.D. thesis at the University of Pennsylvania. I would like to thank both my advisors Ron Donagi and Tony Pantev for their constant help, many discussions and encouragement. I would also like to thank Emanuele Macrì for many useful discussions and Alexander Kuznetsov for all the useful and detailed comments on the earlier drafts of this paper. I would also like to thank the reviewers for their careful reading and insightful comments.

2. Special subvarieties in Prym varieties

In this section, we recall the special subvariety construction of Prym varieties following the work of Welters [23] and Beauville [5]. Let π:Σ~Σ\pi:\widetilde{\Sigma}\to\Sigma be an étale double cover of two smooth curves. Then we denote by Nm:JΣ~JΣ\textrm{Nm}:J\widetilde{\Sigma}\to J\Sigma the norm map on the Jacobians of curves. We also have the induced map Nmd:JdΣ~JdΣ\textrm{Nm}^{d}:J^{d}\widetilde{\Sigma}\to J^{d}\Sigma where JdΣ~J^{d}\widetilde{\Sigma} and JdΣJ^{d}\Sigma are the varieties of degree dd invertible sheaves on Σ~\widetilde{\Sigma} and Σ\Sigma respectively. We recall that the Prym variety is defined to be Prym(Σ~,Σ):=ker(Nm)\textrm{Prym}(\widetilde{\Sigma},\Sigma):=\textrm{ker}(\textrm{Nm})^{\circ} i.e. the connected component of the kernel of the norm map. .

Suppose gdrg^{r}_{d} is a linear system of degree dd and (projective) dimension rr on Σ\Sigma. Consider the Abel-Jacobi maps:

α~:Σ~(d)JdΣ~,\displaystyle\widetilde{\alpha}:\widetilde{\Sigma}^{(d)}\to J^{d}\widetilde{\Sigma}, x~1++x~d𝒪(x~1++x~d)\displaystyle\quad\widetilde{x}_{1}+...+\widetilde{x}_{d}\mapsto\mathcal{O}(\widetilde{x}_{1}+...+\widetilde{x}_{d})
α:Σ(d)JdΣ,\displaystyle\alpha:\Sigma^{(d)}\to J^{d}\Sigma, x1++xd𝒪(x1++xd)\displaystyle\quad x_{1}+...+x_{d}\mapsto\mathcal{O}(x_{1}+...+x_{d})

they fit in the following commutative diagram

(7) Σ~(d){\widetilde{\Sigma}^{(d)}}JdΣ~{J^{d}\widetilde{\Sigma}}Σ(d){\Sigma^{(d)}}JdΣ{J^{d}\Sigma}α~\scriptstyle{\widetilde{\alpha}}π(d)\scriptstyle{\pi^{(d)}}Nmd\scriptstyle{\textrm{Nm}^{d}}α\scriptstyle{\alpha}

The linear system gdrrg^{r}_{d}\cong\mathbb{P}^{r} is naturally a subvariety of Σ(d)\Sigma^{(d)}. We assume that the linear system gdrg^{r}_{d} contains a reduced divisor, so that gdrg^{r}_{d} is not contained in the branch locus of π(d)\pi^{(d)}.

Now, we define W=(π(d))1(gdr)W=(\pi^{(d)})^{-1}(g^{r}_{d}) as the preimage of gdrg^{r}_{d} in Σ~(d)\widetilde{\Sigma}^{(d)}. The image of WW under α~\widetilde{\alpha} is denoted by V=α~(W)V=\widetilde{\alpha}(W). The inverse image (Nmd)1(α(gdr))(\textrm{Nm}^{d})^{-1}(\alpha(g^{r}_{d})) consists of two disjoint components, each of which is isomorphic to the Prym variety Prym(Σ~,Σ)\textrm{Prym}(\widetilde{\Sigma},\Sigma) by translation, and we denote them by Pr0\textrm{Pr}^{0} and Pr1\textrm{Pr}^{1}. By construction, we have that V(Nmd)1(α(gdr))V\subset(\textrm{Nm}^{d})^{-1}(\alpha(g^{r}_{d})), so VV also has two disjoint components ViPriV^{i}\subset\textrm{Pr}^{i} for i=0,1i=0,1. Hence, WW also breaks into a disjoint union of two subvarieties W0W^{0} and W1W^{1} such that α~(Wi)=Vi\widetilde{\alpha}(W^{i})=V^{i}. It is proved in [23, Proposition 8.8] that W0W^{0} and W1W^{1} are the connected components of W.W. Welters [23] called WW the variety of divisors on Σ~\widetilde{\Sigma} lying over the gdrg^{r}_{d} and the two connected components W0W^{0} and W1W^{1} the halves of the variety of divisors WW. The subvarieties ViV^{i} are called the special subvarieties of Pri\textrm{Pr}^{i} associated to the linear system gdr.g^{r}_{d}.

Remark 2.1.

Let σ:Σ~Σ~\sigma:\widetilde{\Sigma}\to\widetilde{\Sigma} be the involution on Σ~.\widetilde{\Sigma}. By [20, Lemma 1], a line bundle Lker(Nm)L\in\textrm{ker}(\textrm{Nm}) can always be written as LMσ(M)L\cong M\otimes\sigma^{*}(M^{\vee}) such that if deg(M)0\textrm{deg}(M)\equiv 0 (resp. 1)1) mod 22, then LPr0L\in\textrm{Pr}^{0} (resp. Pr1\textrm{Pr}^{1}). It follows that if LPriL\in\textrm{Pr}^{i}, then L𝒪(xσ(x))Pr1iL\otimes\mathcal{O}(x-\sigma(x))\in\textrm{Pr}^{1-i} where xΣx\in\Sigma. This implies that if x1++xdWix_{1}+...+x_{d}\in W^{i}, then the divisor σ(x1)++xd=(σ(x1)x1)+(x1++xd)\sigma(x_{1})+...+x_{d}=(\sigma(x_{1})-x_{1})+(x_{1}+...+x_{d}) is contained in W1iW^{1-i}. In particular, we see that if we involute an even number of points xkx_{k} in x1++xdWix_{1}+...+x_{d}\in W^{i}, then the resulting divisor lies in the same component, i.e. kIσ(xk)+jIxjW1i\sum_{k\in I}\sigma(x_{k})+\sum_{j\not\in I}x_{j}\in W^{1-i} if x1++xdWix_{1}+...+x_{d}\in W^{i} and II has even cardinality.

Let Σ¯:=/2×Σ\overline{\Sigma}:=\mathbb{Z}/2\mathbb{Z}\times\Sigma be the constant group scheme over Σ\Sigma. The trivial double cover p:Σ¯Σp:\overline{\Sigma}\to\Sigma also induces a morphism on its dd-th symmetric products Σ¯(d)Σ(d)\overline{\Sigma}^{(d)}\to\Sigma^{(d)}. Let UΣ(d)U\subset\Sigma^{(d)} be the open subset of reduced effective divisors.

Proposition 2.2.

The scheme G:=Σ¯(d)|UG^{\prime}:=\overline{\Sigma}^{(d)}|_{U} is a group scheme over UU.

Proof.

Note that the map GUG^{\prime}\to U is étale. The multiplication map m:Σ¯×ΣΣ¯Σ¯m:\overline{\Sigma}\times_{\Sigma}\overline{\Sigma}\to\overline{\Sigma} induces the map m(d):(Σ¯×ΣΣ¯)(d)Σ¯(d)m^{(d)}:(\overline{\Sigma}\times_{\Sigma}\overline{\Sigma})^{(d)}\to\overline{\Sigma}^{(d)}. On the other hand, the natural projections prj:Σ¯×ΣΣ¯Σ¯pr_{j}:\overline{\Sigma}\times_{\Sigma}\overline{\Sigma}\to\overline{\Sigma} induces the maps prj(d):(Σ¯×ΣΣ¯)(d)Σ¯(d)pr_{j}^{(d)}:(\overline{\Sigma}\times_{\Sigma}\overline{\Sigma})^{(d)}\to\overline{\Sigma}^{(d)} and so r:(p(d)pr1(d))1(U)G×UGr:(p^{(d)}\circ pr_{1}^{(d)})^{-1}(U)\to G^{\prime}\times_{U}G^{\prime} by universal property. It is easy to see that rr is bijective on closed points. As GG^{\prime} and UU are smooth, so G×UGG^{\prime}\times_{U}G^{\prime} is also smooth and hence normal. Therefore, rr is an isomorphism. Then we define the multiplication map on GG^{\prime} to be

G×UGr1(p(d)pr1(d))1(U)m(d)|UGG^{\prime}\times_{U}G^{\prime}\xrightarrow{r^{-1}}(p^{(d)}\circ pr_{1}^{(d)})^{-1}(U)\xrightarrow{m^{(d)}|_{U}}G^{\prime}

At closed points, the group multiplication G()×U()G()G()G^{\prime}(\mathbb{C})\times_{U(\mathbb{C})}G^{\prime}(\mathbb{C})\to G^{\prime}(\mathbb{C}) simply sends

(i=1d(αi,xi),i=1d(βi,xi))i=1d(αi+βi,xi)\left(\sum_{i=1}^{d}(\alpha_{i},x_{i}),\sum_{i=1}^{d}(\beta_{i},x_{i})\right)\to\sum_{i=1}^{d}(\alpha_{i}+\beta_{i},x_{i})

over i=1dxiU()\sum_{i=1}^{d}x_{i}\in U(\mathbb{C}), where αi,βi/2\alpha_{i},\beta_{i}\in\mathbb{Z}/2\mathbb{Z}.

The trivial double cover Σ¯Σ\overline{\Sigma}\to\Sigma always has a section ΣΣ¯\Sigma\to\overline{\Sigma} mapping to q1(0)q^{-1}(0) where q:Σ¯/2q:\overline{\Sigma}\to\mathbb{Z}/2\mathbb{Z} is the projection map. The identity map is defined as the restriction of Σ(d)Σ¯(d)\Sigma^{(d)}\to\overline{\Sigma}^{(d)} to UU, i.e. e:UGe:U\to G^{\prime}.

The inverse map is simply the identity map ι:GG\iota:G^{\prime}\to G^{\prime}. ∎

The projection map q:Σ¯/2q:\overline{\Sigma}\to\mathbb{Z}/2\mathbb{Z} induces GΣ¯(d)q(d)(/2)(d)G^{\prime}\to\overline{\Sigma}^{(d)}\xrightarrow{q^{(d)}}(\mathbb{Z}/2\mathbb{Z})^{(d)} and there is the summation map /2(d)/2\mathbb{Z}/2\mathbb{Z}^{(d)}\to\mathbb{Z}/2\mathbb{Z}, and we denote by s:G/2s:G^{\prime}\to\mathbb{Z}/2\mathbb{Z} the composition of the two maps. Then we define the preimage G:=s1(0)G:=s^{-1}(0).

Corollary 2.3.

The scheme GG is a group scheme over UU.

We can denote a closed point of GG as (λi,xi)\sum(\lambda_{i},x_{i}) such that λi=0\sum\lambda_{i}=0 in /2\mathbb{Z}/2\mathbb{Z} where λi/2\lambda_{i}\in\mathbb{Z}/2\mathbb{Z} and xiΣx_{i}\in\Sigma. In other words, GG is the group UU-scheme of even cardinality subsets of reduced divisors in Σ\Sigma.

Proposition 2.4.

Let gdrg^{r}_{d} be a linear system and consider the half WiΣ~(d)W^{i}\subset\widetilde{\Sigma}^{(d)} of the variety of divisors WW lying over gdrg^{r}_{d}. If we denote by U0:=UgdrU_{0}:=U\cap g^{r}_{d} and G0:=G|U0G_{0}:=G|_{U_{0}}, then there is a G0G_{0}-action μ\mu on Wi|U0W^{i}|_{U_{0}} over U0U_{0}, making it a pseudo G0G_{0}-torsor on U0U_{0} i.e. the induced morphism (μ,pr2):G0×U0Wi|U0Wi|U0×U0Wi|U0(\mu,pr_{2}):G_{0}\times_{U_{0}}W^{i}|_{U_{0}}\to W^{i}|_{U_{0}}\times_{U_{0}}W^{i}|_{U_{0}} is an isomorphism.

Proof.

To simplify notation we denote U~:=Σ~(d)|U\widetilde{U}:=\widetilde{\Sigma}^{(d)}|_{U}. The construction of the group action by G0G_{0} is similar to the multiplication map defined in Proposition 2.2. We first define a group action G×UU~U~G^{\prime}\times_{U}\widetilde{U}\to\widetilde{U}. The involution action σ:Σ¯×ΣΣ~Σ~\sigma:\overline{\Sigma}\times_{\Sigma}\widetilde{\Sigma}\to\widetilde{\Sigma} induces the map σ(d):(Σ¯×ΣΣ~)(d)Σ~(d)\sigma^{(d)}:(\overline{\Sigma}\times_{\Sigma}\widetilde{\Sigma})^{(d)}\to\widetilde{\Sigma}^{(d)}. The natural projection pr1:Σ¯×ΣΣ~Σ¯pr_{1}:\overline{\Sigma}\times_{\Sigma}\widetilde{\Sigma}\to\overline{\Sigma} induces the map pr1(d):(Σ¯×ΣΣ~)(d)Σ~(d)pr_{1}^{(d)}:(\overline{\Sigma}\times_{\Sigma}\widetilde{\Sigma})^{(d)}\to\widetilde{\Sigma}^{(d)}. Then we get by universal property t:(p(d)pr1(d))1(U)G×UU~t:(p^{(d)}\circ pr_{1}^{(d)})^{-1}(U)\to G^{\prime}\times_{U}\widetilde{U}. Again, we can easily check that the map tt is bijective on closed points and G×UU~G^{\prime}\times_{U}\widetilde{U} is smooth and hence normal, the map tt is an isomorphism. We define the group action as

G×UU~t1(p(d)pr1(d))1(U)σ(d)|UU~G^{\prime}\times_{U}\widetilde{U}\xrightarrow{t^{-1}}(p^{(d)}\circ pr_{1}^{(d)})^{-1}(U)\xrightarrow{\sigma^{(d)}|_{U}}\widetilde{U}

This defines another group action by restricting to GGG\subset G^{\prime}. Finally, by Remark 2.1, we see that the restriction of the group action by GG to G0G_{0} defines a group action

μ:G0×U0Wi|U0Wi|U0\mu:G_{0}\times_{U_{0}}W^{i}|_{U_{0}}\to W^{i}|_{U_{0}}

At closed points, the group action μ():G0()×U0()(Wi|U0)()(Wi|U0)()\mu(\mathbb{C}):G_{0}(\mathbb{C})\times_{U_{0}(\mathbb{C})}(W^{i}|_{U_{0}})(\mathbb{C})\to(W^{i}|_{U_{0}})(\mathbb{C}) simply sends

(j=1d(λj,xj),j=1d(wj,xj))j=1d(σλj(wj),xj)\left(\sum^{d}_{j=1}(\lambda_{j},x_{j}),\sum_{j=1}^{d}(w_{j},x_{j})\right)\mapsto\sum_{j=1}^{d}(\sigma^{\lambda_{j}}(w_{j}),x_{j})

over j=1dxjU0()\sum_{j=1}^{d}x_{j}\in U_{0}(\mathbb{C}), where j=1d(wj,xj)(Wi|U0)()\sum_{j=1}^{d}(w_{j},x_{j})\in(W^{i}|_{U_{0}})(\mathbb{C}) with π(wj)=xj\pi(w_{j})=x_{j} and we denote by σ0=Id,σ1=σ\sigma^{0}=Id,\sigma^{1}=\sigma.

Also, it is clear that μ\mu is simply transitive on closed points. Then it follows by the normality of Wi|U0×U0Wi|U0W^{i}|_{U_{0}}\times_{U_{0}}W^{i}|_{U_{0}} that the induced morphism (μ,pr2):G0×U0Wi|U0Wi|U0×U0Wi|U0(\mu,pr_{2}):G_{0}\times_{U_{0}}W^{i}|_{U_{0}}\to W^{i}|_{U_{0}}\times_{U_{0}}W^{i}|_{U_{0}} is an isomorphism. ∎

Example 2.5.

Consider the linear system |KΣ||K_{\Sigma}| i.e. the linear system of canonical divisors. In this case, we have d=2g2d=2g-2 and r=g1r=g-1. Observe that dim(W)=dim(Pri)\textrm{dim}(W)=\textrm{dim}(\textrm{Pr}^{i}) and the fiber of the morphisms WiPriW^{i}\to\textrm{Pr}^{i} at a point [D]Pri[D]\in\textrm{Pr}^{i} is |D|.|D|. It can be shown (e.g. [21, Section 6], [20]) that:

  1. (1)

    α~|W1:W1Pr1\widetilde{\alpha}|_{W^{1}}:W^{1}\to\textrm{Pr}^{1} is birational.

  2. (2)

    α~|W0:W0Pr0\widetilde{\alpha}|_{W^{0}}:W^{0}\to\textrm{Pr}^{0} maps onto a divisor ΘPr0\Theta\subset\textrm{Pr}^{0} and is generically a 1\mathbb{P}^{1}-bundle.

3. Modules over the sheaf of even Clifford algebras

3.1. Conic fibrations and sheaves of even Clifford algebras

For simplicity, we will only discuss conic fibrations i.e. flat quadric fibrations of relative dimension 1. A conic fibration π:QS\pi:Q\to S over a smooth variety SS is defined by a rank 3 vector bundle FF on SS, together with an embedding of a line bundle q:LS2Fq:L\to S^{2}F^{\vee} which is also thought of as a section in S2FLS^{2}F^{\vee}\otimes L^{\vee}. Then QQ is embedded in (F)=Proj(k0SkF)\mathbb{P}(F)=\textrm{Proj}(\bigoplus_{k\geq 0}S^{k}F^{\vee}) as the zero locus of qH0(S,S2FL)=H0((F),𝒪(F)/S(2)(π)L)q\in H^{0}(S,S^{2}F^{\vee}\otimes L^{\vee})=H^{0}(\mathbb{P}(F),\mathcal{O}_{\mathbb{P}(F)/S}(2)\otimes(\pi^{\prime})^{*}L^{\vee}) where we denote by π:(F)S\pi^{\prime}:\mathbb{P}(F)\to S the projection morphism. The morphism π:QS\pi:Q\to S obtained by restricting π\pi^{\prime} to QQ is flat as q:LS2Fq:L\to S^{2}F^{\vee} is assumed to be an embedding and SS is smooth.

Given a conic fibration π:QS\pi:Q\to S, we define the sheaf of even Clifford algebras by following the approach of [2]111Note that we write a line bundle-valued quadratic form as q:LS2Fq:L\to S^{2}F^{\vee} where the authors in [2] write it as LS2FL^{\vee}\to S^{2}F^{\vee}.. Note that q:LS2Fq:L\to S^{2}F^{\vee} induces an 𝒪S\mathcal{O}_{S}-bilinear map q:F×FLq:F\times F\to L^{\vee} (again denoted by qq). Then we can consider the two ideals J1J_{1} and J2J_{2} of the tensor algebra T(FFL)T^{\bullet}(F\otimes F\otimes L) which are generated by

(8) vvfq(v,v),f,uvfvwgq(v,v),fuwgv\otimes v\otimes f-\langle q(v,v),f\rangle,\quad u\otimes v\otimes f\otimes v\otimes w\otimes g-\langle q(v,v),f\rangle u\otimes w\otimes g

respectively, where the sections u,v,wFu,v,w\in F and f,gLf,g\in L, and ,:LL𝒪S\langle\cdot,\cdot\rangle:L^{\vee}\otimes L\to\mathcal{O}_{S} is the natural map. Then the even Clifford algebra is defined as the quotient algebra

(9) 0:=T(FFL)/(J1+J2).\mathcal{B}_{0}:=T^{\bullet}(F\otimes F\otimes L)/(J_{1}+J_{2}).

The sheaf of algebra has naturally a filtration

(10) 𝒪S=F0F1=0\mathcal{O}_{S}=F_{0}\subset F_{1}=\mathcal{B}_{0}

obtained as the image of the truncation of the tensor algebra Ti(FFL)T^{\leq i}(F\otimes F\otimes L) in 0\mathcal{B}_{0}. Moreover, the associated graded piece F1/F02FLF_{1}/F_{0}\cong\wedge^{2}F\otimes L. As an 𝒪S\mathcal{O}_{S}-module, we actually have 0𝒪S(2FL)\mathcal{B}_{0}\cong\mathcal{O}_{S}\oplus(\wedge^{2}F\otimes L) which can be seen by defining the splitting 2FLFFLT(FFL)/(J1+J2)\wedge^{2}F\otimes L\to F\otimes F\otimes L\to T^{\bullet}(F\otimes F\otimes L)/(J_{1}+J_{2}) where 2F\wedge^{2}F is thought of as a subbundle of antisymmetric 2-tensors of FFF\otimes F.

Now, given a conic fibration π:QS\pi:Q\to S and its associated sheaf of even Clifford algebras 0\mathcal{B}_{0} as defined above, a 0\mathcal{B}_{0}-module is a coherent sheaf on SS with a left 0\mathcal{B}_{0}-module structure. We denote by Coh(S,0)\mathrm{Coh}(S,\mathcal{B}_{0}) the abelian category of 0\mathcal{B}_{0}-modules on SS.

3.2. Root stacks

The main objects in this paper are 0\mathcal{B}_{0}-modules. In order to study the category of 0\mathcal{B}_{0}-modules, it is easier to work with a root stack cover of SS. The advantage is that the category of 0\mathcal{B}_{0}-modules is equivalent to the category of modules over a sheaf of Azumaya algebras on the root stack. For more details about root stacks, we refer the readers to [10].

Let \mathcal{L} be a line bundle on a scheme XX and sH0(X,)s\in H^{0}(X,\mathcal{L}) and rr a positive integer. The pair (,s)(\mathcal{L},s) defines a morphism X[𝔸1/𝔾m]X\to\left[\mathbb{A}^{1}/\mathbb{G}_{m}\right], and the rr-th power maps on 𝔸1\mathbb{A}^{1} and 𝔾m\mathbb{G}_{m} induce a morphism θr:[𝔸1/𝔾m][𝔸1/𝔾m]\theta_{r}:\left[\mathbb{A}^{1}/\mathbb{G}_{m}\right]\to\left[\mathbb{A}^{1}/\mathbb{G}_{m}\right]. Following [10], we define the rr-th root stack X,s,rX_{\mathcal{L},s,r} as the fiber product

X×[𝔸1/𝔾m],θr[𝔸1/𝔾m].X\times_{\left[\mathbb{A}^{1}/\mathbb{G}_{m}\right],\theta_{r}}\left[\mathbb{A}^{1}/\mathbb{G}_{m}\right].

The rr-th root stack X,s,rX_{\mathcal{L},s,r} is a Deligne-Mumford stack. Locally on XX, when \mathcal{L} is trivial, X,s,rX_{\mathcal{L},s,r} is just the quotient stack [Spec(𝒪X[t]/(trs))/μr]\left[\textrm{Spec}\left(\mathcal{O}_{X}[t]/(t^{r}-s)\right)/\mu_{r}\right] where μr\mu_{r} is the group of rr-th roots of unity acting on tt by scalar action. The root stack X,s,rX_{\mathcal{L},s,r} has XX as its coarse moduli space. There is a tautological sheaf 𝒯\mathcal{T} on X,s,rX_{\mathcal{L},s,r} satisfying 𝒯rψ\mathcal{T}^{\otimes r}\cong\psi^{*}\mathcal{L} where ψ:X,s,rX\psi:X_{\mathcal{L},s,r}\to X is the projection. When the zero locus of ss is connected, every line bundle on X,s,rX_{\mathcal{L},s,r} is isomorphic to ψG𝒯k\psi^{*}G\otimes\mathcal{T}^{\otimes k} where k{0,,r1}k\in\{0,...,r-1\} is unique and GG is unique up to isomorphism [10, Corollary 3.1.2]. For our purposes, we will mainly consider the case =𝒪X(D)\mathcal{L}=\mathcal{O}_{X}(D) for an effective Cartier divisor DD and s=sDs=s_{D} the section vanishing at D.D. In this case, we will simply write X𝒪X(D),sD,r=XD,rX_{\mathcal{O}_{X}(D),s_{D},r}=X_{D,r} and the tautological sheaf 𝒯\mathcal{T} as 𝒪(Dr)\mathcal{O}(\frac{D}{r}).

Similarly, it is pointed out in [10, Lemma 2.1.1] that there is an equivalence of categories between the category of morphisms X[𝔸n/𝔾mn]X\to\left[\mathbb{A}^{n}/\mathbb{G}_{m}^{n}\right] and the category whose objects are nn-tuples (i,ti)i=1n(\mathcal{L}_{i},t_{i})^{n}_{i=1}, where i\mathcal{L}_{i} is a line bundle on XX and tiH0(X,i)t_{i}\in H^{0}(X,\mathcal{L}_{i}), and whose morphisms (i,ti)i=1n(i,ti)i=1n(\mathcal{L}_{i},t_{i})^{n}_{i=1}\to(\mathcal{L}^{\prime}_{i},t^{\prime}_{i})^{n}_{i=1} are nn-tuples (φi)i=1n(\varphi_{i})^{n}_{i=1} where φi:ii\varphi_{i}:\mathcal{L}_{i}\to\mathcal{L}_{i}^{\prime} is an isomorphism such that φi(ti)=ti\varphi_{i}(t_{i})=t^{\prime}_{i}. If we let 𝔻:=(D1,,Dn)\mathbb{D}:=(D_{1},...,D_{n}) be an nn-tuple of effective Cartier divisors and r=(r1,,rn)\vec{r}=(r_{1},...,r_{n}), then the nn-tuples (𝒪X(Di),sDi)i=1n(\mathcal{O}_{X}(D_{i}),s_{D_{i}})^{n}_{i=1} will determine a morphism X[𝔸n/𝔾mn].X\to\left[\mathbb{A}^{n}/\mathbb{G}_{m}^{n}\right]. Also, the morphisms on 𝔸n\mathbb{A}^{n} and 𝔾mn\mathbb{G}^{n}_{m} sending (x1,,xn)(x1r1,,xnrn)(x_{1},...,x_{n})\mapsto(x_{1}^{r_{1}},...,x_{n}^{r_{n}}) induce a morphism θr:[𝔸n/𝔾m][𝔸n/𝔾m]\theta_{\vec{r}}:\left[\mathbb{A}^{n}/\mathbb{G}_{m}\right]\to\left[\mathbb{A}^{n}/\mathbb{G}_{m}\right]. We define X𝔻,rX_{\mathbb{D},\vec{r}} as the fiber product

X×[𝔸n/𝔾mn],θr[𝔸n/𝔾mn].X\times_{\left[\mathbb{A}^{n}/\mathbb{G}^{n}_{m}\right],\theta_{\vec{r}}}\left[\mathbb{A}^{n}/\mathbb{G}^{n}_{m}\right].

This can be interpreted as iterating the rr-th root stack construction for n=1n=1. There are the tautological sheaves 𝒪(Diri)\mathcal{O}\left(\frac{D_{i}}{r_{i}}\right) on X𝔻,rX_{\mathbb{D},\vec{r}} satisfying 𝒪(Diri)riψ𝒪X(Di)\mathcal{O}\left(\frac{D_{i}}{r_{i}}\right)^{\otimes r_{i}}\cong\psi^{*}\mathcal{O}_{X}(D_{i}). Every line bundle FF on X𝔻,rX_{\mathbb{D},\vec{r}} can be written as

FψGi=1r𝒪(Diri)kiF\cong\psi^{*}G\otimes\prod^{r}_{i=1}\mathcal{\mathcal{O}}\left(\frac{D_{i}}{r_{i}}\right)^{\otimes k_{i}}

where 0kiri10\leq k_{i}\leq r_{i}-1 are unique, and GG is unique up to isomorphism and ψ:X𝔻,rX\psi:X_{\mathbb{D},\vec{r}}\to X is the projection [10, Corollary 3.2.1].

Lemma 3.1.

Let D=D1++DnD=D_{1}+...+D_{n} where DiD_{i} are pairwise disjoint effective Cartier divisors. If r=r1==rnr=r_{1}=...=r_{n}, we have

X𝔻,rXD,r.X_{\mathbb{D},\vec{r}}\xrightarrow{\sim}X_{D,r}.
Proof.

An object of XD,rX_{D,r} over a scheme TT consists of a quadruple (f,N,t,φ)(f,N,t,\varphi) where f:TXf:T\to X is a morphism, NN a line bundle, tH0(T,N)t\in H^{0}(T,N) and φ:Nrf𝒪(D)\varphi:N^{\otimes r}\xrightarrow{\sim}f^{*}\mathcal{O}(D) is an isomorphism such that φ(tr)=fs\varphi(t^{\otimes r})=f^{*}s and ss is the section of 𝒪X(D)\mathcal{O}_{X}(D) vanishing at DD.

On the other hand, an object of X𝔻,rX_{\mathbb{D},\vec{r}} consists of (f,(Ni)i=1n,(ti)i=1n,(φi)i=1n)(f,(N_{i})_{i=1}^{n},(t_{i})^{n}_{i=1},(\varphi_{i})^{n}_{i=1}) where f:TXf:T\to X a morphism, NiN_{i} is a line bundle, tiH0(T,Ni)t_{i}\in H^{0}(T,N_{i}) and φi:Nirifi𝒪(Di)\varphi_{i}:N_{i}^{\otimes r_{i}}\xrightarrow{\sim}f_{i}^{*}\mathcal{O}(D_{i}) is an isomorphism such that φi(tiri)=fisi\varphi_{i}(t_{i}^{\otimes r_{i}})=f_{i}^{*}s_{i} and sis_{i} is the section of 𝒪X(Di)\mathcal{O}_{X}(D_{i}) vanishing at DiD_{i}. We see that there is a natural morphism α:X𝔻,rXD,r\alpha:X_{\mathbb{D},\vec{r}}\to X_{D,r} over XX sending

(11) (f,(Ni)i=1n,(ti)i=1n,(φi)i=1n)(f,i=1nNi,i=1nti,i=1nφi).(f,(N_{i})_{i=1}^{n},(t_{i})^{n}_{i=1},(\varphi_{i})^{n}_{i=1})\mapsto(f,\bigotimes_{i=1}^{n}N_{i},\bigotimes_{i=1}^{n}t_{i},\bigotimes_{i=1}^{n}\varphi_{i}).

To see that this is an isomorphism, we restrict to each open neighborhood UiU_{i} of DiD_{i} away from DjD_{j} (jij\neq i) such that 𝒪(Dj)|Ui𝒪\mathcal{O}(D_{j})|_{U_{i}}\xrightarrow{\sim}\mathcal{O} for jij\neq i and 𝒪(Di)|Ui𝒪(D1+..+Dn)|Ui\mathcal{O}(D_{i})|_{U_{i}}\xrightarrow{\sim}\mathcal{O}(D_{1}+..+D_{n})|_{U_{i}}. Then it is clear that the functor (11) over UiU_{i} is essentially surjective i.e. the image of the quadruples (f,(Ni)i=1n,(ti)i=1n,(φi)i=1n)(f,(N_{i})_{i=1}^{n},(t_{i})^{n}_{i=1},(\varphi_{i})^{n}_{i=1}) where Nj𝒪N_{j}\cong\mathcal{O} and tj=1t_{j}=1 for jij\neq i is dense.

Example 3.2.

(Affine scheme) Let X=Spec(R)X=\textrm{Spec}(R), L=𝒪XL=\mathcal{O}_{X} and ss be a section of 𝒪X\mathcal{O}_{X}. Then XL,s,r[SpecR/μr]X_{L,s,r}\cong\left[\textrm{Spec}R^{\prime}/\mu_{r}\right], where R=R[t]/(trs)R^{\prime}=R[t]/(t^{r}-s), and γt=γ1t\gamma\cdot t=\gamma^{-1}t and γa=a\gamma\cdot a=a for aRa\in R and γμr\gamma\in\mu_{r}. A quasi-coherent sheaf on [SpecR/μr]\left[\textrm{Spec}R^{\prime}/\mu_{r}\right] is a RR^{\prime}-module MM with a μr\mu_{r}-action on MM such that for γμr,bR,mM\gamma\in\mu_{r},b\in R^{\prime},m\in M, we have

γ(bm)=(γb)(γm).\gamma\cdot(b\cdot m)=(\gamma\cdot b)\cdot(\gamma\cdot m).

As μr\mu_{r} is diagonalizable, there is a /r\mathbb{Z}/r\mathbb{Z}-grading MM0Mr1M\cong M_{0}\oplus...\oplus M_{r-1} where γmi=γimi\gamma\cdot m_{i}=\gamma^{i}m_{i} for miMim_{i}\in M_{i}. Note that the components are indexed by the group of characters of μr\mu_{r}, which is /r\mathbb{Z}/r\mathbb{Z}. Similarly, RR0Rr1R^{\prime}\cong R^{\prime}_{0}\oplus...\oplus R^{\prime}_{r-1} where R0=RR^{\prime}_{0}=R. In particular, we see that γ:MM\gamma:M\to M is an RR-module homomorphism, and so each MiM_{i} is an RR-module.

Example 3.3.

(Cyclic cover) When there exists a line bundle NN such that f:Nrf:N^{\otimes r}\cong\mathcal{L}, we can take the cyclic cover for section ss, defined as

ϕ:X~:=Spec¯(𝒜X)X,𝒜X:=𝒪XN(N)r1\phi:\widetilde{X}:=\underline{\textrm{Spec}}\left(\mathcal{A}_{X}\right)\to X,\qquad\mathcal{A}_{X}:=\mathcal{O}_{X}\oplus N^{\vee}\oplus...\oplus(N^{\vee})^{r-1}

where the algebra structure of 𝒜X\mathcal{A}_{X} is given by the map (N)rf()s𝒪(N^{\vee})^{\otimes r}\xrightarrow{f^{\vee}}(\mathcal{L})^{\vee}\xrightarrow{s^{\vee}}\mathcal{O}. By [9, Théorème 3.4], we know that

[X~/μr]X,s,r.\left[\widetilde{X}/\mu_{r}\right]\cong X_{\mathcal{L},s,r}.

Suppose XX is a smooth curve and D=p1++pkD=p_{1}+...+p_{k} is a reduced divisor and r=2r=2. The cyclic cover ϕ:X~X\phi:\widetilde{X}\to X is branched at pip_{i}, we denote by wiw_{i} the ramification points such that ϕ(wi)=pi\phi(w_{i})=p_{i}. Note that the points wiw_{i} are also the fixed points under the involution of X~\widetilde{X}. In this case, the line bundles on the root stack XD,2X_{D,2} can be described in terms of line bundles on X~\widetilde{X} as follows.

Since the root stack XD,2X_{D,2} is the quotient stack [X~/μ2]\left[\widetilde{X}/\mu_{2}\right], a line bundle on [X~/μ2]\left[\widetilde{X}/\mu_{2}\right] is the same as a μ2\mu_{2}-equivariant line bundle on X~\widetilde{X}. On 𝒪X~(wi)\mathcal{O}_{\widetilde{X}}(w_{i}), there is a group action on Tot(𝒪(wi))\textrm{Tot}(\mathcal{O}(w_{i})) which fixes the canonical section vanishing at wiw_{i}, we will denote by L(wi)L(w_{i}) the line bundle 𝒪(wi)\mathcal{O}(w_{i}) together with this μ2\mu_{2}-equivariant sheaf structure. In particular, the induced μ2\mu_{2}-action on the fiber of 𝒪(wi)\mathcal{O}(w_{i}) is Id.-Id. The pull back of a line bundle FF on XX to X~\widetilde{X} is equipped with a natural μ2\mu_{2}-equivariant sheaf structure, whose induced action on the fiber at wiw_{i} is IdId and the μ2\mu_{2}-equivariant bundle is again denoted by ϕF\phi^{*}F. Since ϕN𝒪(iwi)\phi^{*}N\cong\mathcal{O}\left(\sum_{i}w_{i}\right), we can write

𝒪(wi)𝒪(2wi+ijwj)ϕN𝒪(ijwj)ϕ(N𝒪(pi)).\mathcal{O}(w_{i})\cong\mathcal{O}\left(2w_{i}+\sum_{i\neq j}w_{j}\right)\otimes\phi^{*}N^{\vee}\cong\mathcal{O}\left(\sum_{i\neq j}w_{j}\right)\otimes\phi^{*}(N^{\vee}\otimes\mathcal{O}(p_{i})).

So we see that L(ijwj)ϕ(N𝒪(pi))L\left(\sum_{i\neq j}w_{j}\right)\otimes\phi^{*}(N^{\vee}\otimes\mathcal{O}(p_{i})) has the same underlying line bundle as L(wi).L(w_{i}).

As discussed above, every line bundle on XD,2X_{D,2} is of the form ψF𝒪(iIpi2)\psi^{*}F\otimes\mathcal{O}\left(\sum_{i\in I}\frac{p_{i}}{2}\right). In terms of the language of μ2\mu_{2}-equivariant line bundles, we see that 𝒪(pi2)\mathcal{O}\left(\frac{p_{i}}{2}\right) on XD,2X_{D,2} corresponds to L(wi)L(w_{i}) on X~\widetilde{X}. Moreover, the pushforward ψE^\psi_{*}\widehat{E} of a vector bundle E^\widehat{E} on XD,2X_{D,2} is the μ2\mu_{2}-invariant subbundle of the μ2\mu_{2}-equivariant bundle ϕE~\phi_{*}\widetilde{E}, denoted by (ϕE~)μ2(\phi_{*}\widetilde{E})^{\mu_{2}}, where E~\widetilde{E} is the μ2\mu_{2}-equivariant vector bundle corresponding to E^\widehat{E}.

Proposition 3.4.

([9, Proposition 3.12]) Suppose that div(s)div(s) is an effective Cartier divisor. Let \mathcal{F} be a locally free sheaf on X,s,rX_{\mathcal{L},s,r}. For each point xXx\in X, there exists a Zariski open neighborhood UU of xx such that |ψ1(U)\mathcal{F}|_{\psi^{-1}(U)} is a direct sum of invertible sheaves.

3.3. Root stack associated to a conic fibration

Let π:QS\pi:Q\to S be a conic fibration as defined in Section 3.1. We denote by S1SS_{1}\subset S the discriminant locus of degenerate conics. A conic fibration π:QS\pi:Q\to S is said to have simple degeneration if all the fibers are quadrics of corank 1\leq 1 and the discriminant locus is smooth. In other words, the degenerate fibers cannot be double lines. For the rest of the paper, we will assume all the conic fibrations to have simple degeneration.

We define the 22-nd root stack of SS along S1S_{1} as S^:=SS1,2\widehat{S}:=S_{S_{1},2} and ψ:S^S\psi:\widehat{S}\to S the projection. Then it is shown in [14, Section 3.6] that there is a sheaf of algebra ^0\widehat{\mathcal{B}}_{0} on S^\widehat{S} such that ψ^0=0\psi_{*}\widehat{\mathcal{B}}_{0}=\mathcal{B}_{0}, so there is an equivalence of categories

(12) ψ:Coh(S^,^0)Coh(S,0)\psi_{*}:\mathrm{Coh}(\widehat{S},\widehat{\mathcal{B}}_{0})\xrightarrow{\sim}\mathrm{Coh}(S,\mathcal{B}_{0})

where Coh(S^,^0)\mathrm{Coh}(\widehat{S},\widehat{\mathcal{B}}_{0}) is the abelian category of coherent sheaves on S^\widehat{S} with a left ^0\widehat{\mathcal{B}}_{0}-module structure. Moreover, the sheaf of algebra ^0\widehat{\mathcal{B}}_{0} is a sheaf of Azumaya algebra.

Suppose CSC\subset S is a smooth curve such that the intersection S1CS_{1}\cap C is tranverse, we restrict the conic fibration QSQ\to S to the smooth curve CSC\subset S to get a conic fibration π|C:Q|CC\pi|_{C}:Q|_{C}\to C with simple degeneration. We get the root stack C^:=CS1C,2S^|C\widehat{C}:=C_{S_{1}\cap C,2}\cong\widehat{S}|_{C} and denote by ^0\widehat{\mathcal{B}}_{0} the restriction ^0|C^\widehat{\mathcal{B}}_{0}|_{\widehat{C}} by abuse of notation. The sheaf of algebra ^0\widehat{\mathcal{B}}_{0} on C^\widehat{C} is a trivial Azumaya algebra since C^\widehat{C} is smooth and dim(C^)=1\textrm{dim}(\widehat{C})=1 [14, Corollary 3.16]. That means there exists a rank 2 vector bundle E0E_{0} on C^\widehat{C} (root stack construction is preserved under pull back) such that ^0nd(E0)\widehat{\mathcal{B}}_{0}\cong\mathcal{E}nd(E_{0}) and it induces the equivalence of categories:

Coh(C^)Coh(C^,^0)Coh(C,0)E0ψ(E0)\left.\begin{array}[]{rcccl}\mathrm{Coh}(\widehat{C})&\xrightarrow{\sim}&\mathrm{Coh}(\widehat{C},\widehat{\mathcal{B}}_{0})&\xrightarrow{\sim}&\mathrm{Coh}(C,\mathcal{B}_{0})\\ \mathcal{F}&\longmapsto&\mathcal{F}\otimes E_{0}&\longmapsto&\psi_{*}(\mathcal{F}\otimes E_{0})\end{array}\right.

Let us define the rank of a 0\mathcal{B}_{0}-module to be the rank of the underlying 𝒪C\mathcal{O}_{C}-module. In particular, by the equivalence of categories, we have the following:

Corollary 3.5.

The rank of a 0\mathcal{B}_{0}-module ψ(E0)\psi_{*}(\mathcal{F}\otimes E_{0}) on CC must be a multiple of 2.

Let pS1Cp\in S_{1}\cap C. According to Proposition 3.4, there exists a Zariski open neighbourhood UU of pp such that E0|ψ1(U)L1L2E_{0}|_{\psi^{-1}(U)}\cong L_{1}\oplus L_{2} for some line bundles LiL_{i} on ψ1(U).\psi^{-1}(U). As explained in Example 3.2 and by shrinking UU if necessary, we can assume that U=Spec(R)U=\textrm{Spec}(R) is an affine neighbourhood of pp and consider the double cover U~=Spec(R)\widetilde{U}=\textrm{Spec}(R^{\prime}) where R:=R[t]/(t2s)R^{\prime}:=R[t]/(t^{2}-s) and w=div(s)w=div(s) maps to pp. Then the root stack restricted over UU is simply U^=[(Spec(R[t]/(t2s))/μ2]\widehat{U}=\left[\left(\textrm{Spec}(R[t]/(t^{2}-s)\right)/\mu_{2}\right]. Then, as explained in Example 3.3, each LiL_{i} is an μ2\mu_{2}-equivariant bundle on U~\widetilde{U}, and in particular, each LiL_{i} defines a character χi,w:μ2\chi_{i,w}:\mu_{2}\to\mathbb{C}^{*} of μ2\mu_{2} at the fiber Li|wL_{i}|_{w}.

Proposition 3.6.

We have χ1,w(1)χ2,w(1)=1.\chi_{1,w}(-1)\cdot\chi_{2,w}(-1)=-1. In particular, the μ2\mu_{2}-invariant part of the fiber E0|wE_{0}|_{w} is one-dimensional.

Proof.

We will follow the notations in the preceding paragraph. We can further reduce to the localization of RR at pp, we will again write the local ring as RR.

As explained in Example 3.2, the rank 2 vector bundle E0E_{0} on U^\widehat{U} is an RR^{\prime}-module MM with a μ2\mu_{2}-action such that MM0M1M\cong M_{0}\oplus M_{1} where MiM_{i} are RR-modules. By Proposition 3.4, we can write MR[l1]R[l2]M\cong R^{\prime}[l_{1}]\oplus R^{\prime}[l_{2}] as /2\mathbb{Z}/2\mathbb{Z}-graded RR^{\prime}-modules where li{0,1}l_{i}\in\{0,1\}, or equivalently, choose e1Ml1e_{1}\in M_{l_{1}} and e2Ml2e_{2}\in M_{l_{2}} such that MRe1Re2M\cong R^{\prime}e_{1}\oplus R^{\prime}e_{2}. Since χi,w(1)=(1)li\chi_{i,w}(-1)=(-1)^{l_{i}} for i=1,2i=1,2, it suffices to check l1+l2=1/2l_{1}+l_{2}=1\in\mathbb{Z}/2\mathbb{Z}.

Suppose the contrary that l1=l2=0l_{1}=l_{2}=0 (or l1=l2=1l_{1}=l_{2}=1). Recall that E0E_{0} satisfies ψnd(E0)0\psi_{*}\mathcal{E}nd(E_{0})\cong\mathcal{B}_{0} as sheaf of algebras. Since the conic of QQ over pp is degenerate, its even Clifford algebra 0|p\mathcal{B}_{0}|_{p} is not isomorphic to the endomorphism algebra of rank 2.

On the other hand, there is a natural morphism

α:ψnd(E0)nd(ψE0).\alpha:\psi_{*}\mathcal{E}nd(E_{0})\to\mathcal{E}nd(\psi_{*}E_{0}).

Since E0E_{0} corresponds to a /2\mathbb{Z}/2\mathbb{Z}-graded RR^{\prime}-module, nd(E0)\mathcal{E}nd(E_{0}) also corresponds to a /2\mathbb{Z}/2\mathbb{Z}-graded RR^{\prime}-module and so ψnd(E0)\psi_{*}\mathcal{E}nd(E_{0}) corresponds to the μ2\mu_{2}-invariant part i.e. (nd(E0))0\left(\mathcal{E}nd(E_{0})\right)_{0} which is an RR-module. In terms of the RR^{\prime}-basis {e1,e2}\{e_{1},e_{2}\}, (nd(E0))0\left(\mathcal{E}nd(E_{0})\right)_{0} consists of the homogeneous RR-module homomorphisms δ\delta of degree 0:

e1\displaystyle e_{1} u0e1+u0e2\displaystyle\mapsto u_{0}e_{1}+u_{0}^{\prime}e_{2}
e2\displaystyle e_{2} v0e1+v0e2\displaystyle\mapsto v_{0}e_{1}+v_{0}^{\prime}e_{2}

where u0,u0,v0,v0R0=Ru_{0},u_{0}^{\prime},v_{0},v_{0}^{\prime}\in R^{\prime}_{0}=R. Similarly, the module ψE0\psi_{*}E_{0} is the μ2\mu_{2}-invariant part of Re1Re2R^{\prime}e_{1}\oplus R^{\prime}e_{2} which is freely generated by {f1=e1,f1=e2}\{f_{1}=e_{1},f_{1}=e_{2}\} (or {f1=te1,f2=te2}\{f_{1}=te_{1},f_{2}=te_{2}\} when l1=l2=1l_{1}=l_{2}=1) as RR-modules. For l1=l2=0l_{1}=l_{2}=0, δψnd(E0)\delta\in\psi_{*}\mathcal{E}nd(E_{0}) is mapped to an image in nd(ψE0)\mathcal{E}nd(\psi_{*}E_{0}) of the form

f1\displaystyle f_{1} =e1u0e1+u0e2=u0f1+u0f2\displaystyle=e_{1}\mapsto u_{0}e_{1}+u_{0}^{\prime}e_{2}=u_{0}f_{1}+u_{0}^{\prime}f_{2}
f2\displaystyle f_{2} =e2v0e1+v0e2=v0f1+v0f2\displaystyle=e_{2}\mapsto v_{0}e_{1}+v_{0}^{\prime}e_{2}=v_{0}f_{1}+v_{0}^{\prime}f_{2}

Since u0,v0,u0,v0u_{0},v_{0},u_{0}^{\prime},v_{0}^{\prime} are arbitrary elements in RR, the image of α\alpha will be the endomorphism algebra over RR i.e. α\alpha is an isomorphism of RR-algebras, which is a contradiction. For l1=l2=1l_{1}=l_{2}=1, the image of α\alpha is also surjective for the same reason.

The second assertion is clear because χ1,w(1)χ2,w(1)=1\chi_{1,w}(-1)\cdot\chi_{2,w}(-1)=-1 means that one and only one of L1|wL_{1}|_{w} and L2|wL_{2}|_{w} is μ2\mu_{2}-invariant. ∎

3.4. Moduli space of 0\mathcal{B}_{0}-modules

In order to guarantee the existence of a moduli space of 0\mathcal{B}_{0}-modules, we will use Simpson’s theory of moduli spaces of Λ\Lambda-modules [22]. Let us recall the definition of a sheaf of rings of differential operators from Simpson’s paper [22] and follow its notations closely. Suppose SS is a noetherian scheme over \mathbb{C}, and let f:XSf:X\to S be a scheme of finite type over S.S. A sheaf of rings of differential operators on XX over SS is a sheaf of (not necessarily commutative) 𝒪X\mathcal{O}_{X}-algebras Λ\Lambda over X,X, with a filtration Λ0Λ1\Lambda_{0}\subset\Lambda_{1}\subset... which satisfies the following properties:

  1. (1)

    Λ=i=0Λi\Lambda=\bigcup^{\infty}_{i=0}\Lambda_{i} and ΛiΛjΛi+j.\Lambda_{i}\cdot\Lambda_{j}\subset\Lambda_{i+j}.

  2. (2)

    The image of the morphism 𝒪XΛ\mathcal{O}_{X}\to\Lambda is equal to Λ0.\Lambda_{0}.

  3. (3)

    The image of f1(𝒪S)f^{-1}(\mathcal{O}_{S}) in 𝒪X\mathcal{O}_{X} is contained in the center of Λ.\Lambda.

  4. (4)

    The left and right 𝒪X\mathcal{O}_{X}-module structures on Gri(Λ):=Λi/Λi1Gr_{i}(\Lambda):=\Lambda_{i}/\Lambda_{i-1} are equal.

  5. (5)

    The sheaves of 𝒪X\mathcal{O}_{X}-modules Gri(Λ)Gr_{i}(\Lambda) are coherent.

  6. (6)

    The sheaf of graded 𝒪X\mathcal{O}_{X}-algebras Gr(Λ):=i=0Gri(Λ)Gr(\Lambda):=\bigoplus^{\infty}_{i=0}Gr_{i}(\Lambda) is generated by Gr1(Λ)Gr_{1}(\Lambda) in the sense that the morphism of sheaves

    Gr1(Λ)𝒪X.𝒪XGr1(Λ)Gri(Λ)Gr_{1}(\Lambda)\otimes_{\mathcal{O}_{X}}....\otimes_{\mathcal{O}_{X}}Gr_{1}(\Lambda)\to Gr_{i}(\Lambda)

    is surjective.

The definition of stability condition for a Λ\Lambda-module \mathcal{E} is similar as the case of coherent sheaves. We define d(),p(,n),r()d(\mathcal{E}),p(\mathcal{E},n),r(\mathcal{E}) to be the dimension, Hilbert polynomial and rank of the underlying coherent sheaf of \mathcal{E} respectively. As defined in [22], a Λ\Lambda-module \mathcal{E} is pp-semistable (resp. pp-stable) if it is of pure dimension, and if for any sub-Λ\Lambda-module \mathcal{F}\subset\mathcal{E} with 0<r()<r()0<r(\mathcal{F})<r(\mathcal{E}), there exists an NN such that

(13) p(,n)r()p(,n)r()\frac{p(\mathcal{F},n)}{r(\mathcal{F})}\leq\frac{p(\mathcal{E},n)}{r(\mathcal{E})}

(resp. <<) for nNn\geq N.

We will now specialize to the case where S=Spec()S=\textrm{Spec}(\mathbb{C}) and X=2X=\mathbb{P}^{2}. We fix a conic fibration π:Q2\pi:Q\to\mathbb{P}^{2} for the rest of the section and let 0\mathcal{B}_{0} be the associated sheaf of even Clifford algebras.

Proposition 3.7.

The sheaf of 𝒪2\mathcal{O}_{\mathbb{P}^{2}}-algebra 0\mathcal{B}_{0} is a sheaf of rings of differential operators.

Proof.

Recall that as an 𝒪2\mathcal{O}_{\mathbb{P}^{2}}-module, 0𝒪2(2F)\mathcal{B}_{0}\cong\mathcal{O}_{\mathbb{P}^{2}}\oplus(\wedge^{2}F\otimes\mathcal{L}) with the filtration Λ0=𝒪2,Λi=0\Lambda_{0}=\mathcal{O}_{\mathbb{P}^{2}},\Lambda_{i}=\mathcal{B}_{0} for i1i\geq 1. Properties (1), (2), and (5) are clearly satisfied. The center of 0\mathcal{B}_{0} is Λ0\Lambda_{0}, so (3) is also satisfied. The left and right 𝒪2\mathcal{O}_{\mathbb{P}^{2}}-module on 0\mathcal{B}_{0} coincide by definition, so the induced left and right 𝒪2\mathcal{O}_{\mathbb{P}^{2}}-module structure also coincide on Gri(Λ).Gr_{i}(\Lambda). Finally, since Gri(Λ)=0Gr_{i}(\Lambda)=0 for i>1i>1, property (6) is satisfied trivially. ∎

Since 0\mathcal{B}_{0} is a sheaf of rings of differential operators, [22, Theorem 4.7] guarantees the existence of a moduli space of semistable 0\mathcal{B}_{0}-modules with a fixed Hilbert polynomial whose closed points correspond to Jordan equivalence classes of 0\mathcal{B}_{0}-modules. Note that specifying a Hilbert polynomial is equivalent to specifying the Chern character for the case of 2\mathbb{P}^{2}. In this paper, we will be primarily interested in the moduli space of semistable 0\mathcal{B}_{0}-module on 2\mathbb{P}^{2} with Chern character (0,2d,e)(0,2d,e), denoted by 𝔐d,e\mathfrak{M}_{d,e}. The Chern character of a 0\mathcal{B}_{0}-module is defined as the Chern character of the underlying coherent sheaf.

Recall that in the case of a moduli space of one-dimensional coherent sheaves, one can define a support morphism by following Le Potier’s construction [18, Section 2.2]: a pure dimension one coherent sheaf GG on a smooth projective polarized surface (Y,𝒪Y(1))(Y,\mathcal{O}_{Y}(1)) is Cohen-Macaulay, so GG admits a resolution by vector bundles of rank rr:

0A𝑢BG00\to A\xrightarrow{u}B\to G\to 0

Then we can define the so-called Fitting support Fit(G)\textrm{Fit}(G) to be the vanishing subscheme of the induced morphism ru\wedge^{r}u. It can be checked that Fit(G)\textrm{Fit}(G) is independent of the resolution and Fit(G)\textrm{Fit}(G) represents its first Chern class c1(G)\textrm{c}_{1}(G). As the Fitting support construction works in families, it defines a support morphism from the moduli space of pure dimension one sheaves to the Hilbert scheme of curves of degree c1(G)𝒪Y(1)\mathrm{c}_{1}(G)\cdot\mathcal{O}_{Y}(1). When Y=2Y=\mathbb{P}^{2}, the latter space is simply the linear system |𝒪2(deg(G))||\mathcal{O}_{\mathbb{P}^{2}}(\textrm{deg}(G))|.

Similarly, for a pure dimension one 0\mathcal{B}_{0}-module MM of Chern character (0,2d,e)(0,2d,e), we can define the Fitting support of its underlying coherent sheaf Fit(M):=Fit(Forg(M))\textrm{Fit}(M):=\textrm{Fit}(\textrm{Forg}(M)). This construction again works in families and induces a support morphism on the moduli space of 0\mathcal{B}_{0}-modules

Υ:𝔐d,e|𝒪(2d)|.\Upsilon:\mathfrak{M}_{d,e}\to|\mathcal{O}(2d)|.

The following is observed in [17].

Proposition 3.8 ([17]).

The support morphism Υ:𝔐d,e|𝒪(2d)|\Upsilon:\mathfrak{M}_{d,e}\to|\mathcal{O}(2d)| factors through |𝒪(d)||𝒪(2d)|.|\mathcal{O}(d)|\subset|\mathcal{O}(2d)|.

Proof.

By Corollary 3.5, if a 0\mathcal{B}_{0}-module MM is (set-theoretically) supported on a smooth curve CC of degree 2d2d, its rank must be a multiple of 2. It follows that the Fitting support Fit(M)\textrm{Fit}(M) must be a nonreduced curve. So we see that the fiber over a smooth curve C|𝒪(2d)|C\in|\mathcal{O}(2d)| is empty.

If a 0\mathcal{B}_{0}-module MM is supported on an integral but singular curve CC of degree 2d2d, we can pull back MM to its normalization and apply Corollary 3.5, the rank of the pull back of MM must be a multiple of 2. Hence, MM has generically rank multiple of 2, its Fitting support will not be reduced. So again the fiber over C|𝒪(2d)|C\in|\mathcal{O}(2d)| is empty.

Finally, for a 0\mathcal{B}_{0}-module MM supported on a reduced but reducible curve, the same argument as in the previous paragraph shows that the fiber over such a curve is empty.

Hence, the image of Υ\Upsilon must be contained in the nonreduced locus |𝒪(d)||𝒪(2d)|.|\mathcal{O}(d)|\subset|\mathcal{O}(2d)|.

Remark 3.9.

From now on, we will write Υ:𝔐d,e|𝒪(d)|\Upsilon:\mathfrak{M}_{d,e}\to|\mathcal{O}(d)|.

Theorem 3.10 ([17]).

Let Ud|𝒪2(d)|U_{d}\subset|\mathcal{O}_{\mathbb{P}^{2}}(d)| be the open subset of smooth degree dd curves which intersect Δ\Delta transversally and CUdC\in U_{d}. Then

Υ1(C)IPic|I|/2C.\Upsilon^{-1}(C)\cong\bigsqcup_{I}\mathop{\rm Pic}\nolimits^{-|I|/2}C.

where II runs over the even cardinality subsets of {1,,dk}\{1,...,dk\} and k:=deg(Δ)k:=\textrm{deg}(\Delta).

Proof.

We will recall the description of the fiber Υ1(C)\Upsilon^{-1}(C) in [17, Theorem 2.12] and provide more details as it will be important for our purposes in later sections. A 0\mathcal{B}_{0}-module MΥ1(C)M\in\Upsilon^{-1}(C) is a rank 2 vector bundle supported on CC, so we can restrict our attention to 0\mathcal{B}_{0}-modules on C.C. Note that a 0\mathcal{B}_{0}-module MM that is a rank 2 vector bundle on CC is automatically pp-stable, since the rank of any 0\mathcal{B}_{0}-module on CC must be a multiple of 22.

As explained in previous section, there is a rank 2 vector bundle E0E_{0} on the 2nd-root stack C^:=CCΔ,2\widehat{C}:=C_{C\cap\Delta,2} and an equivalence of categories ψ:Coh(C^,nd(E0))Coh(C,0)\psi_{*}:\mathrm{Coh}(\widehat{C},\mathcal{E}nd(E_{0}))\xrightarrow{\sim}\mathrm{Coh}(C,\mathcal{B}_{0}) where ψ:C^C\psi:\widehat{C}\to C is the projection morphism. That means we are looking for line bundles L^\widehat{L} on C^\widehat{C} such that ch(iψ(E0L^))=(0,2d,e)\textrm{ch}(i_{*}\psi_{*}(E_{0}\otimes\widehat{L}))=(0,2d,e) where i:C2i:C\to\mathbb{P}^{2} is the inclusion map. It is clear that ch0(iψ(E0L^))=0\textrm{ch}_{0}(i_{*}\psi_{*}(E_{0}\otimes\widehat{L}))=0 and ch1(iψ(E0L^))=2d.\textrm{ch}_{1}(i_{*}\psi_{*}(E_{0}\otimes\widehat{L}))=2d. To compute ch2\textrm{ch}_{2}, we use the fact that is easily computed by the Grothendieck-Riemann-Roch theorem:

(14) ch2(iG)=deg(G)d22rk(G)\textrm{ch}_{2}(i_{*}G)=\textrm{deg}(G)-\frac{d^{2}}{2}\textrm{rk}(G)

for a vector bundle GG on C.C. So it is equivalent to finding all L^\widehat{L} on C^\widehat{C} such that e=deg(ψ(E0L^)d2e=\textrm{deg}(\psi_{*}(E_{0}\otimes\widehat{L})-d^{2} or deg(ψ(E0L^))=e+d2.\textrm{deg}(\psi_{*}(E_{0}\otimes\widehat{L}))=e+d^{2}.

Case 1: When k=deg(Δ)k=\textrm{deg}(\Delta) is even, in which case 𝒪2(k)|C\mathcal{O}_{\mathbb{P}_{2}}(k)|_{C} admits a square root 𝒪2(k/2)|C\mathcal{O}_{\mathbb{P}_{2}}(k/2)|_{C}, we can take the the cyclic cover ϕ:C~C\phi:\widetilde{C}\to C of order 2 branched at CΔC\cap\Delta with an involution action. As explained in Example 3.3 the root stack C^\widehat{C} is isomorphic to the quotient stack [C~/μ2]\left[\widetilde{C}/\mu_{2}\right]. Moreover, the morphism ϕ:C~C\phi:\widetilde{C}\to C factors as C~𝜂C^𝜓C\widetilde{C}\xrightarrow{\eta}\widehat{C}\xrightarrow{\psi}C.

Let wiC~,pi=ϕ(wi)CΔw_{i}\in\widetilde{C},p_{i}=\phi(w_{i})\in C\cap\Delta be the ramification and branch points respectively. Recall that a line bundle L^\widehat{L} on C^\widehat{C} can be written as ψJ𝒪(λipi2)\psi^{*}J\otimes\mathcal{O}\left(\sum\lambda_{i}\frac{p_{i}}{2}\right) such that JJ is a line bundle on CC, where λi{0,1}\lambda_{i}\in\{0,1\}. As a μ2\mu_{2}-equivariant line bundle L^=ϕ(J)L(λiwi)\widehat{L}=\phi^{*}(J)\otimes L\left(\sum\lambda_{i}w_{i}\right) on C~\widetilde{C} (following the notation in Example 3.3),

(15) c1(ϕ(E0ϕ(J)L(λiwi))μ2)\displaystyle\textrm{c}_{1}\left(\phi_{*}\left(E_{0}\otimes\phi^{*}(J)\otimes L\left(\sum\lambda_{i}w_{i}\right)\right)^{\mu_{2}}\right) =c1(Jϕ(E0L(λiwi))μ2)\displaystyle=\textrm{c}_{1}\left(J\otimes\phi_{*}\left(E_{0}\otimes L\left(\sum\lambda_{i}w_{i}\right)\right)^{\mu_{2}}\right)
=2c1(J)+c1(ϕ(E0L(λiwi))μ2)\displaystyle=2\textrm{c}_{1}(J)+\textrm{c}_{1}\left(\phi_{*}\left(E_{0}\otimes L\left(\sum\lambda_{i}w_{i}\right)\right)^{\mu_{2}}\right)

Note that we have the short exact sequence

0ϕ(E0)μ2ϕ(E0L(λiwi))μ2iIϕ(E0L(λiwi)𝒪wi)μ200\to\phi_{*}\left(E_{0}\right)^{\mu_{2}}\to\phi_{*}\left(E_{0}\otimes L\left(\sum\lambda_{i}w_{i}\right)\right)^{\mu_{2}}\to\bigoplus_{i\in I}\phi_{*}\left(E_{0}\otimes L\left(\sum\lambda_{i}w_{i}\right)\otimes\mathcal{O}_{w_{i}}\right)^{\mu_{2}}\to 0

where II is the subset of {1,2,,dk}\{1,2,...,dk\} such that λi=1\lambda_{i}=1 for iIi\in I. Since the dimension of the fiber ϕ(E0𝒪wi)μ2|pi(E0|wi)μ2\phi_{*}\left(E_{0}\otimes\mathcal{O}_{w_{i}}\right)^{\mu_{2}}|_{p_{i}}\cong(E_{0}|_{w_{i}})^{\mu_{2}} at pip_{i} is 1 by Proposition 3.6, we have c1(ϕ(E0𝒪wi)μ2)=1\textrm{c}_{1}\left(\phi_{*}\left(E_{0}\otimes\mathcal{O}_{w_{i}}\right)^{\mu_{2}}\right)=1 which implies that

c1(ϕ(E0𝒪wiL(λiwi))μ2)=1.\textrm{c}_{1}\left(\phi_{*}\left(E_{0}\otimes\mathcal{O}_{w_{i}}\otimes L\left(\sum\lambda_{i}w_{i}\right)\right)^{\mu_{2}}\right)=1.

Then the last expresssion of (15) becomes

2c1(J)+c1(ϕ(E0)μ2)+|I|.2\textrm{c}_{1}(J)+\textrm{c}_{1}\left(\phi_{*}(E_{0})^{\mu_{2}}\right)+\left|I\right|.

where |I||I| is its cardinality. Since E0E_{0} on C^\widehat{C} is determined up to tensorization by a line bundle, this expression means that we can assume deg((ϕE0)μ2)=e+d2\textrm{deg}\left(\left(\phi_{*}E_{0}\right)^{\mu_{2}}\right)=e+d^{2}.

We also see that the condition deg(ψ(E0L^))=e+d2\textrm{deg}(\psi_{*}(E_{0}\otimes\widehat{L}))=e+d^{2} becomes

e+d2=2deg(J)+deg(ψ(E0))+|I|0=2deg(F)+|I|,e+d^{2}=2\textrm{deg}(J)+\textrm{deg}(\psi_{*}(E_{0}))+|I|\implies 0=2\textrm{deg}(F)+|I|,

which is the same as saying that the degree of L^\widehat{L} as a line bundle on C~\widetilde{C} is 0. The condition 2deg(F)+|I|=02\textrm{deg}(F)+|I|=0 only makes sense if |I||I| is even. We also see that for each fixed II, the set of line bundles satisfying the condition above is Pic|I|/2(C)\mathop{\rm Pic}\nolimits^{-|I|/2}(C). Thus, Υ1(C)IPic|I|/2(C)\Upsilon^{-1}(C)\cong\bigsqcup_{I}\mathop{\rm Pic}\nolimits^{-|I|/2}(C) where II runs over the set of even cardinality subsets of dkdk.

Case 2: When k=deg(Δ)k=\textrm{deg}(\Delta) is odd, we will use the trick by choosing an auxiliary line H2H\subset\mathbb{P}^{2} which intersects CC transversally and Da:=HCD_{a}:=H\cap C is disjoint from D:=CΔD:=C\cap\Delta. Then the line bundle 𝒪C(D+Da)𝒪2(k+1)|C\mathcal{O}_{C}(D+D_{a})\cong\mathcal{O}_{\mathbb{P}^{2}}(k+1)|_{C} has a natural square root 𝒪2((k+1)/2)|C\mathcal{O}_{\mathbb{P}^{2}}((k+1)/2)|_{C}, so we can again consider the cyclic cover C~\widetilde{C} branched at C(Δ+H)C\cap(\Delta+H). The root stack C¯:=CD+Da,2\overline{C}:=C_{D+D_{a},2} is now isomorphic to the quotient stack [C~/μ2]\left[\widetilde{C}/\mu_{2}\right]. We again denote by C^\widehat{C} the root stack CCΔ,2C_{C\cap\Delta,2}. By Lemma 3.1, the stack C¯\overline{C} is isomorphic to CD,Da,(2,2)C_{D,D_{a},(2,2)} which is constructed as a fiber product, hence CD,Da,(2,2)C_{D,D_{a},(2,2)} projects to C^.\widehat{C}. We denote the composition by f:C¯CD,Da,(2,2)C^f:\overline{C}\xrightarrow{\sim}C_{D,D_{a},(2,2)}\to\widehat{C}.

C~{\widetilde{C}}C¯{{\overline{C}}}C^{\widehat{C}}C{C}q\scriptstyle{q}ϕ\scriptstyle{\phi}f\scriptstyle{f}ψ\scriptstyle{\psi}

Let L^\widehat{L} be a line bundle on C^\widehat{C}, we want to find all such line bundles such that ch(ψ(E0L^))=(0,2d,e)\textrm{ch}(\psi_{*}(E_{0}\otimes\widehat{L}))=(0,2d,e). Recall that this is equivalent to finding deg(ψ(E0L^))=e+d2\textrm{deg}(\psi_{*}(E_{0}\otimes\widehat{L}))=e+d^{2}. By [10, Theorem 3.1.1 (3)] (the proof there works for any vector bundle), we know that M^ffM^\widehat{M}\cong f_{*}f^{*}\widehat{M} for any vector bundles on C^\widehat{C}, so

ψ(E0L^)=ψf(f(E0L^))=ϕ(f(E0L^))μ2\displaystyle\psi_{*}(E_{0}\otimes\widehat{L})=\psi_{*}f_{*}\left(f^{*}\left(E_{0}\otimes\widehat{L}\right)\right)=\phi_{*}\left(f^{*}\left(E_{0}\otimes\widehat{L}\right)\right)^{\mu_{2}}

As C¯[C~/μ2]\overline{C}\cong\left[\widetilde{C}/\mu_{2}\right], f(E0L^)f^{*}\left(E_{0}\otimes\widehat{L}\right) on C¯\overline{C} is a μ2\mu_{2}-equivariant vector bundle on C~\widetilde{C} whose induced μ2\mu_{2}-characters at the fixed points wiDaw_{i}\in D_{a} is trivial. In other words, the problem now is to find all line bundles on C~\widetilde{C} of the form ϕ(F)𝒪(λiwi)\phi^{*}(F)\otimes\mathcal{O}(\sum\lambda_{i}w_{i}) where wiϕ1(D)w_{i}\in\phi^{-1}(D) such that

deg(ϕ(E0ϕ(F)𝒪(λiwi))μ2)=e+d2.\textrm{deg}\left(\phi_{*}\left(E_{0}\otimes\phi^{*}(F)\otimes\mathcal{O}\left(\sum\lambda_{i}w_{i}\right)\right)^{\mu_{2}}\right)=e+d^{2}.

The same argument as in Case 1 applies and implies that 2deg(F)+|I|=02\textrm{deg}(F)+|I|=0 where II is the subset of {1,,dk}\{1,...,dk\} such that λi=1\lambda_{i}=1 and |I||I| is its cardinality. Hence, Υ1(C)\Upsilon^{-1}(C) is again isomorphic to IPic|I|/2(C)\bigsqcup_{I}\mathop{\rm Pic}\nolimits^{-|I|/2}(C). Note that although we use the auxiliary line HH and the divisor DaD_{a} in the argument, the result is independent of them.

Remark 3.11.

Note that the isomorphism Υ1(C)IPic|I|/2(C)\Upsilon^{-1}(C)\cong\bigsqcup_{I}\mathop{\rm Pic}\nolimits^{-|I|/2}(C) here is not canonical, as E0E_{0} is only determined up to tensorization by line bundles.

Suppose d=1,2d=1,2. For d<deg(Δ)d<\textrm{deg}(\Delta), if we call the line bundle Ld:=O2(d)|ΔL_{d}:=O_{\mathbb{P}^{2}}(d)|_{\Delta} on Δ\Delta, it is easy to see that |O2(d)||Ld||O_{\mathbb{P}^{2}}(d)|\cong|L_{d}|. Recall the group scheme G|UdG|_{U_{d}} over UdU_{d} defined in Section 2.

Corollary 3.12.

With the same notation as above, for d=1,2d=1,2, Υ1(C)\Upsilon^{-1}(C) is a G|CG|_{C}-torsor.

Proof.

For d=1,2d=1,2, the Picard group Pica(C)\mathop{\rm Pic}\nolimits^{a}(C) is trivial for any aa. Let pi=CΔ\sum p_{i}=C\cap\Delta be the divisor corresponding to CC under |𝒪2(d)||Ld|.|\mathcal{O}_{\mathbb{P}^{2}}(d)|\cong|L_{d}|. We can denote a closed point of G|CG|_{C} by (λi,pi)\sum(\lambda_{i},p_{i}) where λi/2\lambda_{i}\in\mathbb{Z}/2\mathbb{Z} and λi=0\sum\lambda_{i}=0. Since any MΥ1(C)M\in\Upsilon^{-1}(C) can be written as ψM^\psi_{*}\widehat{M}, the group G|CG|_{C} acts on Υ1(C)\Upsilon^{-1}(C) by

((λi,pi))M=ψ(M^𝒪(λipi2)hC12λi)\left(\sum(\lambda_{i},p_{i})\right)\cdot M=\psi_{*}\left(\widehat{M}\otimes\mathcal{O}\left(\sum\lambda_{i}\frac{p_{i}}{2}\right)\otimes h_{C}^{-\frac{1}{2}\sum\lambda_{i}}\right)

where hC=ψ𝒪C(1)h_{C}=\psi^{*}\mathcal{O}_{C}(1). To see that G|CG|_{C} acts simply transitively, fix E0E_{0} such that for MΥ1(C)M\in\Upsilon^{-1}(C), Mψ(E0L^)M\cong\psi_{*}(E_{0}\otimes\widehat{L}), then the action becomes

((λi,pi))M=ψ(E0L^𝒪(λipi2)hC12λi)\left(\sum(\lambda_{i},p_{i})\right)\cdot M=\psi_{*}\left(E_{0}\otimes\widehat{L}\otimes\mathcal{O}\left(\sum\lambda_{i}\frac{p_{i}}{2}\right)\otimes h_{C}^{-\frac{1}{2}\sum\lambda_{i}}\right)

which is clearly simply transitive by the description of Υ1(C)\Upsilon^{-1}(C) in the proof of Theorem 3.10. ∎

4. Moduli spaces of 0\mathcal{B}_{0}-modules and special subvarieties of Prym varieties

In this section, we will construct the rational map from the moduli space 𝔐d,e\mathfrak{M}_{d,e} to the Prym variety Prym(Δ~,Δ)\textrm{Prym}(\widetilde{\Delta},\Delta). The key observation is that our 0\mathcal{B}_{0}-modules are supported on plane curves CC which intersect the discriminant curve Δ\Delta in finitely many points. The 0\mathcal{B}_{0}-modules restrict to a representation of the even Clifford algebra over each of these points. These representations then define a lift of the intersection CΔΔC\cap\Delta\subset\Delta to Δ~\widetilde{\Delta}, which will be a point in the variety of divisors lying over the linear system |𝒪2(C)|Δ||\mathcal{O}_{\mathbb{P}^{2}}(C)|_{\Delta}|, and maps to Prym(Δ~,Δ)\textrm{Prym}(\widetilde{\Delta},\Delta). So we begin by studying the representation theory for our purpose.

4.1. Representation theory of degenerate even Clifford algebras

In this subsection, we will restrict our attention to the fiber of the sheaf of the even Clifford algebras 0\mathcal{B}_{0} over a fixed pCΔp\in C\cap\Delta, which is a \mathbb{C}-algebra, denoted by AA. Note that all the fibers over the points in CΔC\cap\Delta are isomorphic as \mathbb{C}-algebra since the fiber 0|p\mathcal{B}_{0}|_{p} over a point pCΔp\in C\cap\Delta is defined by a degenerate quadratic form of corank 1 and all quadratic forms of corank 1 are isomorphic over \mathbb{C}. Let VV be a vector space of dimension 3, and qS2Vq\in S^{2}V^{*} a quadratic form of rank 2. The even Clifford algebra is defined as a vector space 2V\mathbb{C}\oplus\wedge^{2}V together with an algebra structure defined as follows. First, we can always find a basis {e1,e2,e3}\{e_{1},e_{2},e_{3}\} of VV such that qq is represented as the matrix diag(1,1,0)\textrm{diag}(1,1,0) and we denote by {1,x:=ie1e2,y:=ie2e3,z:=e1e3}\{1,x:=ie_{1}\wedge e_{2},y:=ie_{2}\wedge e_{3},z:=e_{1}\wedge e_{3}\} the basis of 2V\mathbb{C}\oplus\wedge^{2}V. The relations are given by

(16) x2=1,y2=z2=0,xy=z,xz=y,xy=yx,xz=zx,yz=zy=0.x^{2}=1,y^{2}=z^{2}=0,xy=-z,xz=-y,xy=-yx,xz=-zx,yz=zy=0.

Since AA is a finite-dimensional associative algebra, we can understand it via quivers and path algebras. We refer the reader to [1] for the basics of quivers and path algebras.

Proposition 4.1.

The algebra AA is isomorphic to the path algebra associated to the following quiver QQ

(17) +{+}{-}α\scriptstyle{\alpha}β\scriptstyle{\beta}

with relations αβ=βα=0.\alpha\beta=\beta\alpha=0.

Proof.

We begin by finding the idempotents i.e. elements bb in AA such that b2=bb^{2}=b. This is achieved by setting up the equations

(a0+a1x+a2y+a3z)2=(a0+a1x+a2y+a3z)(a_{0}+a_{1}x+a_{2}y+a_{3}z)^{2}=(a_{0}+a_{1}x+a_{2}y+a_{3}z)

and solving the equations in a0,a1,a2,a3a_{0},a_{1},a_{2},a_{3}. It is easy to check that the idempotents are:

0,1,12(1±x)+a2y+a3z0,1,\frac{1}{2}(1\pm x)+a_{2}y+a_{3}z

for a2,a3a_{2},a_{3}\in\mathbb{C} and that

e+:=12(1+x),e:=12(1x)e_{+}:=\frac{1}{2}(1+x),\quad e_{-}:=\frac{1}{2}(1-x)

is a complete set of primitive orthogonal idempotents of AA. From the description of idempotents, it is clear that the only central idempotents are 0,10,1, so AA is connected.

We also need to compute the radical of AA. Observe that the ideal I=(y,z)I=(y,z) is clearly nilpotent, i.e. I2=0I^{2}=0 and A/I[x]/(x21)A/I\cong\mathbb{C}[x]/(x^{2}-1)\cong\mathbb{C}\oplus\mathbb{C}. By [1, Corollary 1.4(c)], this implies that rad(A)=I=(y,z).\textrm{rad}(A)=I=(y,z). It also follows that AA is a basic algebra by [1, Proposition 6.2(a)].

The arrows between ++\to- of the associated quiver is described by

e(rad(A)/rad(A)2)e+=12(1x)(y,z)12(1+x)=(y+z).e_{-}(\textrm{rad}(A)/\textrm{rad}(A)^{2})e_{+}=\frac{1}{2}(1-x)(y,z)\frac{1}{2}(1+x)=\mathbb{C}(y+z).

Similarly, the arrows between +-\to+ is described by

e+(rad(A)/(rad(A)2))e=(yz)e_{+}(\textrm{rad}(A)/(\textrm{rad}(A)^{2}))e_{-}=\mathbb{C}(y-z)

and the arrows between -\to- and +++\to+

(18) e(rad(A)/(rad(A)2))e=e+(rad(A)/(rad(A)2))e+=0.e_{-}(\textrm{rad}(A)/(\textrm{rad}(A)^{2}))e_{-}=e_{+}(\textrm{rad}(A)/(\textrm{rad}(A)^{2}))e_{+}=0.

Hence, the associated quiver QQ [1, Definition 3.1] of AA is given by

+{+}{-}α\scriptstyle{\alpha}β\scriptstyle{\beta}

and we obtain a surjective map QA\mathbb{C}Q\to A from the path algebra Q\mathbb{C}Q associated to the quiver QQ to AA by sending the generators

(19) e+12(1+x),e12(1x),α12(y+z),β12(yz).e_{+}\mapsto\frac{1}{2}(1+x),\quad e_{-}\mapsto\frac{1}{2}(1-x),\quad\alpha\mapsto\frac{1}{2}(y+z),\quad\beta\mapsto\frac{1}{2}(y-z).

It is easy to see that αβ=βα=0\alpha\beta=\beta\alpha=0, and since any other paths of higher length must contain a factor of αβ\alpha\beta or βα\beta\alpha, we see that the kernel of kQAkQ\to A must be J=(αβ,βα)J=(\alpha\beta,\beta\alpha). Therefore, we have an isomorphism Q/JA\mathbb{C}Q/J\cong A.

Remark 4.2.

We can prove the isomorphism in Proposition 4.1 directly by checking that the map defined in (19) is indeed an isomorphism of \mathbb{C}-algebras. The detail with the idempotents and the radical ideal in the proof above is just to display a more systematic approach.

Since we are mainly interested in 0\mathcal{B}_{0}-modules that are locally free of rank 2, the fiber of such module over pCΔp\in C\cap\Delta is a representation of AA on 2\mathbb{C}^{2}. In light of the interpretation of AA as a path algebra, we can easily classify all the isomorphism classes of representations on 2\mathbb{C}^{2}. The isomorphism classes of representations of AQ/JA\cong\mathbb{C}Q/J on 2\mathbb{C}^{2} are listed as follows:

  1. (1)
    {\mathbb{C}}{\mathbb{C}}1\scriptstyle{1}0\scriptstyle{0}
  2. (2)
    {\mathbb{C}}{\mathbb{C}}0\scriptstyle{0}1\scriptstyle{1}
  3. (3)
    {\mathbb{C}}{\mathbb{C}}0\scriptstyle{0}0\scriptstyle{0}
  4. (4)
    2{\mathbb{C}^{2}}0{0}0\scriptstyle{0}0\scriptstyle{0}
  5. (5)
    0{0}2{\mathbb{C}^{2}}0\scriptstyle{0}0\scriptstyle{0}

4.2. Construction

Recall the geometric set-up: we have a rank 3 bundle FF on 2\mathbb{P}^{2} and an embedding of a line bundle q:LS2Fq:L\hookrightarrow S^{2}F^{\vee}. This defines a conic bundle Q(F)Q\subset\mathbb{P}(F) as the zero locus of qH0(2,S2FL)H0((F),𝒪(F)/2(2)(π)L)q\in H^{0}(\mathbb{P}^{2},S^{2}F^{\vee}\otimes L^{\vee})\cong H^{0}(\mathbb{P}(F),\mathcal{O}_{\mathbb{P}(F)/\mathbb{P}^{2}}(2)\otimes(\pi^{\prime})^{*}L^{\vee}) in (F)\mathbb{P}(F) where we denote π:(F)2\pi^{\prime}:\mathbb{P}(F)\to\mathbb{P}^{2}. The discriminant curve is assumed to be smooth and denoted by Δ\Delta. As an 𝒪2\mathcal{O}_{\mathbb{P}^{2}}-module, the sheaf of even Clifford algebras 0\mathcal{B}_{0} on 2\mathbb{P}^{2} is

0𝒪22FL.\mathcal{B}_{0}\cong\mathcal{O}_{\mathbb{P}^{2}}\oplus\wedge^{2}F\otimes L.

We will restrict our attention to 0\mathcal{B}_{0}-modules supported on a degree dd smooth curve C2C\subset\mathbb{P}^{2} which intersects Δ\Delta transversely. Given such a 0\mathcal{B}_{0}-module MM on CC, for each pCΔp\in C\cap\Delta we consider the vector subspace

K:=ker(0|pnd(M)|p).K:=\textrm{ker}(\mathcal{B}_{0}|_{p}\to\mathcal{E}nd(M)|_{p}).

As we will see in Proposition 4.3 (1), KK is a vector subspace of 2F|pL|p\wedge^{2}F|_{p}\otimes L|_{p}. The natural isomorphisms w:2Fdet(F)Fw:\wedge^{2}F\xrightarrow{\sim}\det(F)\otimes F^{\vee} and F(F)F\xrightarrow{\sim}(F^{\vee})^{\vee} give rise to another vector space

K:=ker(F|p(F)|pwpdet(F)|pL|pK(Ldet(F))|p)K^{\prime}:=\textrm{ker}(F|_{p}\xrightarrow{\sim}(F^{\vee})^{\vee}|_{p}\xrightarrow{w_{p}^{\vee}\otimes\det(F)|_{p}\otimes L|_{p}}K^{\vee}\otimes(L\otimes\det(F))|_{p})

where wp:K(2FL)|p(det(F)FL)|pw_{p}:K\hookrightarrow(\wedge^{2}F\otimes L)|_{p}\to(\det(F)\otimes F^{\vee}\otimes L)|_{p} is the composition of the inclusion map and the isomorphism ww restricted to p.p. Hence, (K)\mathbb{P}(K^{\prime}) is a linear subspace in (F|p)\mathbb{P}(F|_{p}). In the light of Proposition 4.3, KK^{\prime} is the two-dimensional vector space in F|pF|_{p} that corresponds to the line KK in 2F|p\wedge^{2}F|_{p} (identified with (2FL)|p\wedge^{2}F\otimes L)|_{p}).

Proposition 4.3.
  1. (1)

    K2F|pL|p𝒪|p2F|pL|pK\subset\wedge^{2}F|_{p}\otimes L|_{p}\subset\mathcal{O}|_{p}\oplus\wedge^{2}F|_{p}\otimes L|_{p} ;

  2. (2)

    dim(K)=1\textrm{dim}(K)=1 and dim(K)=2\textrm{dim}(K^{\prime})=2;

  3. (3)

    The line (K)(F|p)\mathbb{P}(K^{\prime})\subset\mathbb{P}(F|_{p}) is one of the two irreducible components of the degenerate conic Q|p(F|p)Q|_{p}\subset\mathbb{P}(F|_{p}).

Proof.

First of all, we can always choose a basis {e1,e2,e3}\{e_{1},e_{2},e_{3}\} of F|pF|_{p} and a trivialization i:L|pi:L|_{p}\cong\mathbb{C} so that q|pq|_{p} is represented by diag(1,1,0)\textrm{diag}(1,1,0). The trivialization ii induces an isomorphism of \mathbb{C}-algebras 0|p𝒪|p2F|p\mathcal{B}_{0}|_{p}\cong\mathcal{O}|_{p}\oplus\wedge^{2}F|_{p} where the latter is generated by 1𝒪|p1\in\mathcal{O}|_{p} and {x:=ie1e2,y:=ie2e3,z:=e1e3}2F|p\{x:=ie_{1}\wedge e_{2},y:=ie_{2}\wedge e_{3},z:=e_{1}\wedge e_{3}\}\subset\wedge^{2}F|_{p} with relation

x2=1,y2=z2=0,xy=z,xz=y,xy=yx,xz=zx,yz=0.x^{2}=1,y^{2}=z^{2}=0,xy=-z,xz=-y,xy=-yx,xz=-zx,yz=0.

The irreducible components of Q|p(F|p)Q|_{p}\subset\mathbb{P}(F|_{p}) are given by the projectivization of the isotropic planes in F|pF|_{p} with respect to qq. If we write a vector vF|pv\in F|_{p} as i=13aiei\sum_{i=1}^{3}a_{i}e_{i}, then the two isotropic planes are given by the two equations

(20) a1+ia2=0,a1ia2=0a_{1}+ia_{2}=0,\quad a_{1}-ia_{2}=0

which correspond to the lines in 2F|p\wedge^{2}F|_{p}

(21) ie2e3+e1e3=y+z,ie2e3e1e3=yz.\mathbb{C}\langle ie_{2}\wedge e_{3}+e_{1}\wedge e_{3}\rangle=\mathbb{C}\langle y+z\rangle,\quad\mathbb{C}\langle ie_{2}\wedge e_{3}-e_{1}\wedge e_{3}\rangle=\mathbb{C}\langle y-z\rangle.

To prove all the claims, it suffices to show that K0|pK\subset\mathcal{B}_{0}|_{p} corresponds to one of the these lines in the subspace 2F|p𝒪|p2F|p\wedge^{2}F|_{p}\subset\mathcal{O}|_{p}\oplus\wedge^{2}F|_{p}. Indeed, then KK^{\prime} will correspond to one of the isotropic planes.

Recall that with the choice of basis {1,x,y,z}\{1,x,y,z\} of 𝒪|p2F|p\mathcal{O}|_{p}\oplus\wedge^{2}F|_{p}, we have an isomorphism Q/J𝒪|p2F|p\mathcal{\mathbb{C}}Q/J\xrightarrow{\sim}\mathcal{O}|_{p}\oplus\wedge^{2}F|_{p}:

e+12(1+x),e12(1x),α12(y+z),β12(yz).e_{+}\mapsto\frac{1}{2}(1+x),\quad e_{-}\mapsto\frac{1}{2}(1-x),\quad\alpha\mapsto\frac{1}{2}(y+z),\quad\beta\mapsto\frac{1}{2}(y-z).

Then the kernel of 0|pnd(M)|p\mathcal{B}_{0}|_{p}\to\mathcal{E}nd(M)|_{p} can be computed by the composition

Q/J𝒪|p2F|p0|pnd(M)|p\mathbb{C}Q/J\xrightarrow{\sim}\mathcal{O}|_{p}\oplus\wedge^{2}F|_{p}\xrightarrow{\sim}\mathcal{B}_{0}|_{p}\to\mathcal{E}nd(M)|_{p}

which is a representation of the path algebra Q/J\mathbb{C}Q/J. As we will see in Proposition 4.5, the isomorphism classes of the representation of Q/J\mathbb{C}Q/J on 2\mathbb{C}^{2} in this case must be either type (1) and (2). Taking this as granted for a moment, we have:

  1. (1)

    For type (1), the kernel of Q/J0|pnd(M)|p\mathbb{C}Q/J\xrightarrow{\sim}\mathcal{B}_{0}|_{p}\to\mathcal{E}nd(M)|_{p} is β\mathbb{C}\langle\beta\rangle which corresponds to K=yz𝒪|p2F|p.K=\mathbb{C}\langle y-z\rangle\subset\mathcal{O}|_{p}\oplus\wedge^{2}F|_{p}.

  2. (2)

    For type (2), the kernel of Q/J0|pnd(M)|p\mathbb{C}Q/J\xrightarrow{\sim}\mathcal{B}_{0}|_{p}\to\mathcal{E}nd(M)|_{p} is α\mathbb{C}\langle\alpha\rangle which corresponds to K=y+z𝒪|p2F|pK=\mathbb{C}\langle y+z\rangle\subset\mathcal{O}|_{p}\oplus\wedge^{2}F|_{p}.

All the claims follow immediately. ∎

Remark 4.4.

The trivialization i:L|pi:L|_{p}\cong\mathbb{C} does not cause any ambiguity in the identifications; as we are only interested in identification of vector subspaces, other trivializations will only differ by a scalar multiplication.

Proposition 4.5.

The representation of 0|p\mathcal{B}_{0}|_{p} obtained from a 0\mathcal{B}_{0}-module MM as the fiber M|pM|_{p} over pCΔp\in C\cap\Delta must have isomorphism class of either type (1) or type (2).

Proof.

Fix pCΔp\in C\cap\Delta. Let n=3,4,5n=3,4,5 and MnM_{n} be a 0\mathcal{B}_{0}-modules such that its fiber over pp is a 0\mathcal{B}_{0}-representation of type nn isomorphism class. We can choose a local parameter t𝒪C,pt\in\mathcal{O}_{C,p} as 𝒪C,p\mathcal{O}_{C,p} is a discrete valuation ring.

Then MnM_{n} induces the homomorphisms over the local ring 𝒪C,p\mathcal{O}_{C,p} and over the residue field κ(p)\kappa(p) (i.e. fiber)

ρn:0𝒪C,pnd(Mn)𝒪C,p,ρn0:0κ(p)=0|pnd(Mn)κ(p)=nd(Mn)|p.\rho_{n}:\mathcal{B}_{0}\otimes\mathcal{O}_{C,p}\to\mathcal{E}nd(M_{n})\otimes\mathcal{O}_{C,p},\quad\rho_{n}^{0}:\mathcal{B}_{0}\otimes\kappa(p)=\mathcal{B}_{0}|_{p}\to\mathcal{E}nd(M_{n})\otimes\kappa(p)=\mathcal{E}nd(M_{n})|_{p}.

Again, we can always choose a basis {e1,e2,e3}\{e_{1},e_{2},e_{3}\} of F|pF|_{p} and a trivialization i:L|pi:L|_{p}\cong\mathbb{C} so that q|pq|_{p} is represented by diag(1,1,0)\textrm{diag}(1,1,0). The trivialization ii induces an isomorphism of \mathbb{C}-algebras 0|p𝒪|p2F|p\mathcal{B}_{0}|_{p}\cong\mathcal{O}|_{p}\oplus\wedge^{2}F|_{p} where the latter is generated by 1𝒪|p1\in\mathcal{O}|_{p} and {x:=ie1e2,y:=ie2e3,z:=e1e3}2F|p\{x:=ie_{1}\wedge e_{2},y:=ie_{2}\wedge e_{3},z:=e_{1}\wedge e_{3}\}\subset\wedge^{2}F|_{p} with relations

x2=1,y2=z2=0,xy=z,xz=y,xy=yx,xz=zx,yz=0.x^{2}=1,y^{2}=z^{2}=0,xy=-z,xz=-y,xy=-yx,xz=-zx,yz=0.

Recall that the algebra 𝒪|p2F|p\mathcal{O}|_{p}\oplus\wedge^{2}F|_{p} is isomorphic to Q/J\mathbb{C}Q/J generated by e+,e,α,β.e_{+},e_{-},\alpha,\beta. If we call the isomorphism j:Q/J𝒪|p2F|p0|pj:\mathbb{C}Q/J\xrightarrow{\sim}\mathcal{O}|_{p}\oplus\wedge^{2}F|_{p}\xrightarrow{\sim}\mathcal{B}_{0}|_{p}, then clearly ρn0(j(α))=ρn0(j(β))=0\rho^{0}_{n}(j(\alpha))=\rho^{0}_{n}(j(\beta))=0. It follows that ρn0(y)=ρn0(j(α+β))=0\rho^{0}_{n}(y)=\rho^{0}_{n}(j(\alpha+\beta))=0 and similarly ρn0(z)=0.\rho^{0}_{n}(z)=0. So that means ρn(y)=tPy\rho_{n}(y)=tP_{y} and ρn(z)=tPz\rho_{n}(z)=tP_{z} for some Py,Pznd(Mn)𝒪C,pP_{y},P_{z}\in\mathcal{E}nd(M_{n})\otimes\mathcal{O}_{C,p}.

As FF is locally free of rank 3, by Nakayama lemma, we can lift the basis {e1,e2,e3}\{e_{1},e_{2},e_{3}\} of F|pF|_{p} to a basis (also denoted as {e1,e2,e3}\{e_{1},e_{2},e_{3}\} by abuse of notation) of F𝒪C,pF\otimes\mathcal{O}_{C,p} over 𝒪C,p\mathcal{O}_{C,p}. The quadratic form qq is represented by the matrix

(22) (fij)=(f11f12f13f21f22f23f31f32f33)(f_{ij})=\begin{pmatrix}f_{11}&f_{12}&f_{13}\\ f_{21}&f_{22}&f_{23}\\ f_{31}&f_{32}&f_{33}\end{pmatrix}

where fijf_{ij} are elements in 𝒪C,p\mathcal{O}_{C,p}. By the choice of basis {e1,e2,e2}\{e_{1},e_{2},e_{2}\}, we have f11|p0f_{11}|_{p}\neq 0, f22|p0f_{22}|_{p}\neq 0 and fij|p=0f_{ij}|_{p}=0 for (i,j)(1,1),(2,2)(i,j)\neq(1,1),(2,2). It follows that fij=tfijf_{ij}=tf_{ij}^{\prime} for (i,j)(1,1),(2,2)(i,j)\neq(1,1),(2,2) and fij𝒪C,pf_{ij}^{\prime}\in\mathcal{O}_{C,p}. Then in 0𝒪C,p\mathcal{B}_{0}\otimes\mathcal{O}_{C,p} we have

yz\displaystyle yz =(ie2e3)(e1e3)\displaystyle=(ie_{2}e_{3})(e_{1}e_{3})
=if31e2e3ie2e1e3e3\displaystyle=if_{31}e_{2}e_{3}-ie_{2}e_{1}e_{3}e_{3}
=if31e2e3if33e2e1\displaystyle=if_{31}e_{2}e_{3}-if_{33}e_{2}e_{1}
=f31yif33f21+f33x\displaystyle=f_{31}y-if_{33}f_{21}+f_{33}x

(here we omit the """\wedge" between the (ei)s(e_{i})^{\prime}s) so it follows that

ρn(yz)=t2f31Pyit2f33f21+tf33ρn(x).\rho_{n}(yz)=t^{2}f_{31}^{\prime}P_{y}-it^{2}f_{33}^{\prime}f_{21}^{\prime}+tf_{33}^{\prime}\rho_{n}(x).

Since

ρn(yz)=ρn(y)ρn(z)=t2PyPz\rho_{n}(yz)=\rho_{n}(y)\rho_{n}(z)=t^{2}P_{y}P_{z}

by equating the two expression, we get

t2PyPz=t2f31Pyit2f33f21+tf33ρn(x)tPyPz=tf31Pyitf33f21+f33ρn(x).t^{2}P_{y}P_{z}=t^{2}f_{31}^{\prime}P_{y}-it^{2}f_{33}^{\prime}f_{21}^{\prime}+tf_{33}^{\prime}\rho_{n}(x)\implies tP_{y}P_{z}=tf_{31}^{\prime}P_{y}-itf_{33}^{\prime}f_{21}^{\prime}+f_{33}^{\prime}\rho_{n}(x).

Note that f33f^{\prime}_{33} is invertible in 𝒪C,p\mathcal{O}_{C,p} because otherwise det(fij)\det(f_{ij}) will have zeros of order 2 with respect to tt, which is not allowed since we assume that CC intersects Δ\Delta transversally. Hence, we can write ρn(x)=tPx\rho_{n}(x)=tP_{x} for some Pxnd(Mn)𝒪C,pP_{x}\in\mathcal{E}nd(M_{n})\otimes\mathcal{O}_{C,p}. In particular, we must have ρn0(x2)=0.\rho_{n}^{0}(x^{2})=0.

On the other hand, we have

x2\displaystyle x^{2} =(ie1e2)(ie1e2)\displaystyle=(ie_{1}e_{2})(ie_{1}e_{2})
=f21e1e2+e1e1e2e2\displaystyle=-f_{21}e_{1}e_{2}+e_{1}e_{1}e_{2}e_{2}
=if21x+f11f22\displaystyle=if_{21}x+f_{11}f_{22}

(again we omit the """\wedge" between the (ei)s(e_{i})^{\prime}s) and so ρn0(x2)=(f11f22)|p0\rho_{n}^{0}(x^{2})=(f_{11}f_{22})|_{p}\neq 0 as f21|p=0f_{21}|_{p}=0, f11|p0f_{11}|_{p}\neq 0 and f22|p0.f_{22}|_{p}\neq 0. Hence, a contradiction.

Let Ud|𝒪2(d)|U_{d}\subset|\mathcal{O}_{\mathbb{P}^{2}}(d)| be the subset of smooth curves of degree dd which intersect Δ\Delta transversally. For d<deg(Δ)=kd<\textrm{deg}(\Delta)=k, if we call the line bundle Ld:=O2(d)|ΔL_{d}:=O_{\mathbb{P}^{2}}(d)|_{\Delta} on Δ\Delta, it is easy to see that |O2(d)||Ld||O_{\mathbb{P}^{2}}(d)|\cong|L_{d}|. Hence, we can consider the variety of divisors WdW_{d} lying over |Ld||L_{d}| and its two components as WdiW^{i}_{d} for i=0,1i=0,1.

For each 0\mathcal{B}_{0}-module M𝔐d,eM\in\mathfrak{M}_{d,e} with support on CUd,C\in U_{d}, let j:CΔCj:C\cap\Delta\hookrightarrow C be the inclusion, by Proposition 4.3, the assignment

M𝒦:=ker(j0jnd(M))M\mapsto\mathcal{K}:=\textrm{ker}\left(j^{*}\mathcal{B}_{0}\to j^{*}\mathcal{E}nd(M)\right)

is argued to be contained in j(2F)j^{*}(\wedge^{2}F) and it defines exactly a point in Wd.W_{d}.

This construction also works in families. Let TT be a scheme and T\mathcal{M}_{T} be a flat family of 0\mathcal{B}_{0}-modules on 2\mathbb{P}^{2} with supports on curves of UdU_{d} and Chern character (0,2d,e)(0,2d,e) i.e. T\mathcal{M}_{T} is a p10p^{*}_{1}\mathcal{B}_{0}-module on 2×T\mathbb{P}^{2}\times T flat over TT with Fit(t)Ud\textrm{Fit}(\mathcal{M}_{t})\in U_{d} and ch(t)=(0,2d,e)\textrm{ch}(\mathcal{M}_{t})=(0,2d,e) for all t:Spec()Tt:\textrm{Spec}(\mathbb{C})\to T where t=tT\mathcal{M}_{t}=t^{*}\mathcal{M}_{T} and p1:2×T2p_{1}:\mathbb{P}^{2}\times T\to\mathbb{P}^{2} is the projection. Then we get a map T𝔐d,e|Ud|Ud|Δ(dk)T\to\mathfrak{M}_{d,e}|_{U_{d}}\to|U_{d}|\subset\Delta^{(dk)}. We can restrict the family of 0\mathcal{B}_{0}-modules to Δ×T2×T.\Delta\times T\subset\mathbb{P}^{2}\times T. Consider the universal divisor

𝒟Δ×Δ(dk){\mathcal{D}\subset\Delta\times\Delta^{(dk)}}Δ{\Delta}Δ(dk){\Delta^{(dk)}}

By pulling back 𝒟\mathcal{D} along the map Δ×TΔ×Δ(dk)\Delta\times T\to\Delta\times\Delta^{(dk)}, we get another divisor 𝒟TΔ×T\mathcal{D}_{T}\subset\Delta\times T and denote the inclusion by iT:𝒟TΔ×T2×Ti_{T}:\mathcal{D}_{T}\hookrightarrow\Delta\times T\hookrightarrow\mathbb{P}^{2}\times T.

We will write FT:=iTp1FF_{T}:=i_{T}^{*}p_{1}^{*}F and LT:=iTp1LL_{T}:=i_{T}^{*}p_{1}^{*}L. The sheaf

𝒦T:=ker(iTp10iTnd(T))\mathcal{K}_{T}:=\textrm{ker}\left(i_{T}^{*}p_{1}^{*}\mathcal{B}_{0}\to i_{T}^{*}\mathcal{E}nd(\mathcal{M}_{T})\right)

has constant fiber dimension one and is contained in the rank 3 vector bundle iTp1(2FTLT)i_{T}^{*}p_{1}^{*}\left(\wedge^{2}F_{T}\otimes L_{T}\right) on 𝒟T\mathcal{D}_{T} by Proposition 4.3, where FT=p1FF_{T}=p_{1}^{*}F and LT=p1LL_{T}=p_{1}^{*}L. Again, since there are the natural isomorphisms w:2FTdet(FT)FTw:\wedge^{2}F_{T}\xrightarrow{\sim}\det(F_{T})\otimes F_{T}^{\vee} and FT(FT)F_{T}\xrightarrow{\sim}(F_{T}^{\vee})^{\vee}, we can define

𝒦T:=ker(FT(FT)wTdet(FT)LT𝒦TLTdet(FT))\mathcal{K}_{T}^{\prime}:=\textrm{ker}(F_{T}\xrightarrow{\sim}(F_{T}^{\vee})^{\vee}\xrightarrow{w_{T}^{\vee}\otimes\det(F_{T})\otimes L_{T}}\mathcal{K}_{T}^{\vee}\otimes L_{T}\otimes\det(F_{T}))

where wT:𝒦T2FTLTdet(FT)FTLTw_{T}:\mathcal{K}_{T}\hookrightarrow\wedge^{2}F_{T}\otimes L_{T}\to\det(F_{T})\otimes F_{T}^{\vee}\otimes L_{T} is the composition. As we checked in Proposition 4.3 that each fiber of the projectivization (𝒦T)(FT)\mathbb{P}(\mathcal{K}^{\prime}_{T})\subset\mathbb{P}(F_{T}) is a component of the fiber of a degenerate conic in the conic bundle Q2Q\to\mathbb{P}^{2}, so we have (𝒦T)iTp1Q(FT)\mathbb{P}(\mathcal{K}^{\prime}_{T})\subset i_{T}^{*}p_{1}^{*}Q\subset\mathbb{P}(F_{T}). Since Δ~Δ\widetilde{\Delta}\to\Delta is the curve parametrizing the irreducible components of Q|ΔΔQ|_{\Delta}\to\Delta, it follows that (𝒦T)\mathbb{P}(\mathcal{K}^{\prime}_{T}) over 𝒟T\mathcal{D}_{T} defines a divisor 𝒟~TΔ~×T\widetilde{\mathcal{D}}_{T}\subset\widetilde{\Delta}\times T that maps to 𝒟T\mathcal{D}_{T} via Δ~×TΔ×T\widetilde{\Delta}\times T\to\Delta\times T. The divisor 𝒟~T\widetilde{\mathcal{D}}_{T} is a TT-family of degree dkdk divisors on Δ~\widetilde{\Delta}, so it defines a map TΔ~(dk)T\to\widetilde{\Delta}^{(dk)} which factors through Wd|UdW_{d}|_{U_{d}} since 𝒟T\mathcal{D}_{T} is induced from a map TUdT\to U_{d}. It is easy to check that the assignment from T\mathcal{M}_{T} to TWdT\to W_{d} is functorial, hence we obtain a morphism over UdU_{d}:

(23) 𝔐d,e|Ud{\mathfrak{M}_{d,e}|_{U_{d}}}Wd|Ud{W_{d}|_{U_{d}}}Ud{U_{d}}Φ\scriptstyle{\Phi}
Proposition 4.6.

The morphism Φ|C:𝔐d,e|C=Υ1(C)Wd|C\Phi|_{C}:\mathfrak{M}_{d,e}|_{C}=\Upsilon^{-1}(C)\to W_{d}|_{C} over CUdC\in U_{d} is G|CG|_{C}-equivariant.

Proof.

Let (λi,pi)G|C\sum(\lambda_{i},p_{i})\in G|_{C}. Recall the notations from Theorem 3.10 that C^=CCΔ,2\widehat{C}=C_{C\cap\Delta,2} is the 2nd-root stack and ψ:C^C\psi:\widehat{C}\to C is the projection morphism. By Theorem 3.10, we can write M𝔐d,e|CM\in\mathfrak{M}_{d,e}|_{C} as M=ψ(E0L^)M=\psi_{*}(E_{0}\otimes\widehat{L}) for a L^\widehat{L} line bundle on C^\widehat{C} by choosing a rank 2 bundle E0E_{0} (recall that E0E_{0} is determined up to a line bundle) on C^\widehat{C}. We need to show that

(24) Φ((λi,pi)M)=Φ(ψ(E0L^𝒪(iλi2pi)hC12λi))=(λipi)Φ(M)\Phi\left(\sum\left(\lambda_{i},p_{i}\right)\cdot M\right)=\Phi\left(\psi_{*}\left(E_{0}\otimes\widehat{L}\otimes\mathcal{O}\left(\sum_{i}\frac{\lambda_{i}}{2}p_{i}\right)\otimes h_{C}^{-\frac{1}{2}\sum\lambda_{i}}\right)\right)=\left(\sum\lambda_{i}p_{i}\right)\cdot\Phi(M)

Since Φ(M)\Phi(M) is determined at each point in CΔC\cap\Delta, it suffices to check the equivariance property over a point pCΔp\in C\cap\Delta. As we checked that ker(0|pnd(ψ(E0L^))|p)\textrm{ker}\left(\mathcal{B}_{0}|_{p}\to\mathcal{E}nd\left.\left(\psi_{*}\left(E_{0}\otimes\widehat{L}\right)\right)\right|_{p}\right) always determines one of the two preimages of pΔp\in\Delta in the double cover Δ~,\widetilde{\Delta}, to prove the proposition it suffices to show that

ker(0|pnd(ψ(E0L^))|p)ker(0|pnd(ψ(E0L^𝒪(p2)))|p)\textrm{ker}\left(\mathcal{B}_{0}|_{p}\to\mathcal{E}nd\left.\left(\psi_{*}\left(E_{0}\otimes\widehat{L}\right)\right)\right|_{p}\right)\neq\textrm{ker}\left(\mathcal{B}_{0}|_{p}\to\mathcal{E}nd\left.\left(\psi_{*}\left(E_{0}\otimes\widehat{L}\otimes\mathcal{O}\left(\frac{p}{2}\right)\right)\right)\right|_{p}\right)

or equivalently,

(25) ker(0|pnd(ψ(E0L^))|p)ker(0|pnd(ψ(E0L^𝒪(p2)))|p).\textrm{ker}\left(\mathcal{B}_{0}|_{p}\to\mathcal{E}nd\left.\left(\psi_{*}\left(E_{0}\otimes\widehat{L}\right)\right)\right|_{p}\right)\neq\textrm{ker}\left(\mathcal{B}_{0}|_{p}\to\mathcal{E}nd\left.\left(\psi_{*}\left(E_{0}\otimes\widehat{L}\otimes\mathcal{O}\left(-\frac{p}{2}\right)\right)\right)\right|_{p}\right).

In fact, we can simplify further by assuming L^=𝒪C^\widehat{L}=\mathcal{O}_{\widehat{C}}.

The 0\mathcal{B}_{0}-module structure on ψ(E0L^)\psi_{*}(E_{0}\otimes\widehat{L}) can be described concretely by the composition of the isomorphism 0ψnd(E0)ψnd(E0L^)\mathcal{B}_{0}\cong\psi_{*}\mathcal{E}nd(E_{0})\cong\psi_{*}\mathcal{E}nd(E_{0}\otimes\widehat{L}) and the natural morphism

α:ψnd(E0L^)nd(ψ(E0L^)).\alpha:\psi_{*}\mathcal{E}nd(E_{0}\otimes\widehat{L})\to\mathcal{E}nd(\psi_{*}(E_{0}\otimes\widehat{L})).

In particular, we can define

α0\displaystyle\alpha^{0} :ψnd(E0)ψnd(E0𝒪(p2))nd(ψ(E0𝒪(p2)))\displaystyle:\psi_{*}\mathcal{E}nd(E_{0})\xrightarrow{\sim}\psi_{*}\mathcal{E}nd\left(E_{0}\otimes\mathcal{O}\left(-\frac{p}{2}\right)\right)\to\mathcal{E}nd\left(\psi_{*}\left(E_{0}\otimes\mathcal{O}\left(-\frac{p}{2}\right)\right)\right)
α1\displaystyle\alpha^{1} :ψnd(E0)nd(ψ(E0))\displaystyle:\psi_{*}\mathcal{E}nd(E_{0})\to\mathcal{E}nd\left(\psi_{*}(E_{0})\right)

Hence, to check that (25) holds, it is equivalent to show that ker(α0|p)ker(α1|p)\textrm{ker}(\alpha^{0}|_{p})\neq\textrm{ker}(\alpha^{1}|_{p}).

To check this, we proceed as in Example 3.2 and Proposition 3.6 and work in an affine neighborhood Z=Spec(R)Z=\textrm{Spec}(R) of pp and the double cover Z~=Spec(R)\widetilde{Z}=\textrm{Spec}(R^{\prime}) where R:=R[t]/(t2s))R^{\prime}:=R[t]/(t^{2}-s)) and div(s)=pdiv(s)=p. So that the root stack restricted over ZZ is simply Z^=[Spec(R[t]/(t2s))/μ2]\widehat{Z}=[\textrm{Spec}(R[t]/(t^{2}-s))/\mu_{2}]. Recall the notations that a /2\mathbb{Z}/2\mathbb{Z}-graded RR^{\prime}-module is written as A=A0A1A=A_{0}\oplus A_{1} with AiA_{i} being the graded pieces. We can further reduce to the localization of RR at pp, we will again write the local ring as RR and its unique maximal ideal 𝔪\mathfrak{m} which contains s.s.

As argued in Proposition 3.6, E0E_{0} is a /2\mathbb{Z}/2\mathbb{Z}-graded RR^{\prime}-module N=N0N1N=N_{0}\oplus N_{1} and we can choose e1N0e_{1}\in N_{0} and e2N1e_{2}\in N_{1} such that NRe1Re2N\cong R^{\prime}e_{1}\oplus R^{\prime}e_{2}. In terms of the RR^{\prime}-basis {e1,e2}\{e_{1},e_{2}\}, ψ(nd(E0))(nd(E0))0\psi_{*}(\mathcal{E}nd(E_{0}))\cong\left(\mathcal{E}nd(E_{0})\right)_{0} consists of homogeneous RR-module homomorphisms δ\delta of degree 0:

(26) e1u0e1+u1e2\displaystyle e_{1}\mapsto u_{0}e_{1}+u_{1}e_{2}
e2v1e1+v0e2\displaystyle e_{2}\mapsto v_{1}e_{1}+v_{0}e_{2}

where ui,vi(R)i.u_{i},v_{i}\in(R^{\prime})_{i}. Then we can write u1=tu~1u_{1}=t\widetilde{u}_{1} and v1=tv~1v_{1}=t\widetilde{v}_{1} where u~1,v~1(R)0=R.\widetilde{u}_{1},\widetilde{v}_{1}\in(R^{\prime})_{0}=R.

As before, the module ψE0\psi_{*}E_{0} is freely generated by {f1=e1,f2=te2}\{f_{1}=e_{1},f_{2}=te_{2}\} as RR-module. Suppose that δψEnd(E0)\delta\in\psi_{*}End(E_{0}) is of the form (26), then its image in nd(ψE0)\mathcal{E}nd(\psi_{*}E_{0}) under α\alpha will be a map of the form

f1=e1\displaystyle f_{1}=e_{1} u0e1+u~1(te2)=u0f1+u~1f2\displaystyle\mapsto u_{0}e_{1}+\widetilde{u}_{1}(te_{2})=u_{0}f_{1}+\widetilde{u}_{1}f_{2}
f2=te2\displaystyle f_{2}=te_{2} v~1t(te1)+v0(te2)=sv~1f1+v0f2\displaystyle\mapsto\widetilde{v}_{1}t(te_{1})+v_{0}(te_{2})=s\widetilde{v}_{1}f_{1}+v_{0}f_{2}

If we choose the generators of ψnd(E0)\psi_{*}\mathcal{E}nd(E_{0}) to be the following RR-valued matrices (with respect to the basis {e1,e2})\{e_{1},e_{2}\})

(27) I=(1001),a:=(1001),b:=(0tt0),c:=(0tt0)I=\begin{pmatrix}1&0\\ 0&1\end{pmatrix},\quad a:=\begin{pmatrix}-1&0\\ 0&1\end{pmatrix},\quad b:=\begin{pmatrix}0&t\\ t&0\end{pmatrix},\quad c:=\begin{pmatrix}0&t\\ -t&0\end{pmatrix}

their images in nd(ψE0)\mathcal{E}nd(\psi_{*}E_{0}) are the corresponding RR-matrices (with respect to the basis {f1,f2}\{f_{1},f_{2}\}):

(28) (1001),(1001),(0s10),(0s10)\begin{pmatrix}1&0\\ 0&1\end{pmatrix},\quad\begin{pmatrix}-1&0\\ 0&1\end{pmatrix},\quad\begin{pmatrix}0&s\\ 1&0\end{pmatrix},\quad\begin{pmatrix}0&s\\ -1&0\end{pmatrix}

As the homomorphism 𝒪(p2)𝒪C^\mathcal{O}\left(-\frac{p}{2}\right)\to\mathcal{O}_{\widehat{C}} is represented as the inclusion RtRR^{\prime}t\hookrightarrow R^{\prime} of RR^{\prime}-modules, the homomorphism E0𝒪(p2)E0E_{0}\otimes\mathcal{O}\left(-\frac{p}{2}\right)\to E_{0} corresponds to taking the RR^{\prime}-module homomorphism

(29) R(te1)R(te2)R(e1)R(e2).R^{\prime}(te_{1})\oplus R^{\prime}(te_{2})\to R^{\prime}(e_{1})\oplus R^{\prime}(e_{2}).

Note that ψ(E0𝒪(p2))\psi_{*}\left(E_{0}\otimes\mathcal{O}\left(-\frac{p}{2}\right)\right) corresponds to the RR-module R(sf1)Rf2R(sf_{1})\oplus Rf_{2} as the μ2\mu_{2}-invariant submodule of R(te1)R(te2)R^{\prime}(te_{1})\oplus R^{\prime}(te_{2}). Pushing the homomorphism (29) forward ψ(E0𝒪(p2))ψE0\psi_{*}\left(E_{0}\otimes\mathcal{O}\left(-\frac{p}{2}\right)\right)\to\psi_{*}E_{0} corresponds to taking the μ2\mu_{2}-invariant part R(sf1)Rf2Rf1Rf2R(sf_{1})\oplus Rf_{2}\to Rf_{1}\oplus Rf_{2}, which is represented by the RR-valued matrix

(s001)\begin{pmatrix}s&0\\ 0&1\end{pmatrix}

with respect to the bases {sf1,f2}\{sf_{1},f_{2}\} of ψE0𝒪(p2)\psi_{*}E_{0}\otimes\mathcal{O}(-\frac{p}{2}) and {f1,f2}\{f_{1},f_{2}\} of ψE0\psi_{*}E_{0}.

For any δψ(nd(E0))\delta\in\psi_{*}(\mathcal{E}nd(E_{0})), there are 𝒪Z\mathcal{O}_{Z}-module homomorphisms α0(δ):ψ(E0𝒪(p2))ψ(E0𝒪(p2))\alpha^{0}(\delta):\psi_{*}(E_{0}\otimes\mathcal{O}(-\frac{p}{2}))\to\psi_{*}(E_{0}\otimes\mathcal{O}(-\frac{p}{2})) and α1(δ):ψE0ψE0\alpha^{1}(\delta):\psi_{*}E_{0}\to\psi_{*}E_{0}. They form a commutative diagram by the definition of a ψnd(E0)\psi_{*}\mathcal{E}nd(E_{0})-module homomorphism

ψ(E0𝒪(p2)){\psi_{*}(E_{0}\otimes\mathcal{O}(-\frac{p}{2}))}ψ(E0𝒪(p2)){\psi_{*}(E_{0}\otimes\mathcal{O}(-\frac{p}{2}))}ψE0{\psi_{*}E_{0}}ψE0{\psi_{*}E_{0}}α0(δ)\scriptstyle{\alpha^{0}(\delta)}α1(δ)\scriptstyle{\alpha^{1}(\delta)}

In terms of the bases {f1,f2},{sf1,f2}\{f_{1},f_{2}\},\{sf_{1},f_{2}\} of ψE0\psi_{*}E_{0} and ψ(E0𝒪(p2))\psi_{*}(E_{0}\otimes\mathcal{O}(-\frac{p}{2})), the morphisms above can be written as

RR{R\oplus R}RR{R\oplus R}RR{R\oplus R}RR{R\oplus R}α0(δ)\scriptstyle{\alpha^{0}(\delta)}(s001)\scriptstyle{\begin{pmatrix}s&0\\ 0&1\end{pmatrix}}(s001)\scriptstyle{\begin{pmatrix}s&0\\ 0&1\end{pmatrix}}α1(δ)\scriptstyle{\alpha^{1}(\delta)}

Since α1(b)=(0s10)\alpha^{1}(b)=\begin{pmatrix}0&s\\ 1&0\end{pmatrix}, it is easy to check that α0(b)\alpha^{0}(b) must be (01s0)\begin{pmatrix}0&1\\ s&0\end{pmatrix}. Similarly, we have the following α0(δ)\alpha^{0}(\delta) when α1(δ)\alpha^{1}(\delta) is the other generator:

α1(I)=(1001)α0(I)=(1001)\displaystyle\alpha^{1}(I)=\begin{pmatrix}1&0\\ 0&1\end{pmatrix}\implies\alpha^{0}(I)=\begin{pmatrix}1&0\\ 0&1\end{pmatrix}
α1(a)=(1001)α0(a)=(1001)\displaystyle\alpha^{1}(a)=\begin{pmatrix}-1&0\\ 0&1\end{pmatrix}\implies\alpha^{0}(a)=\begin{pmatrix}-1&0\\ 0&1\end{pmatrix}
α1(c)=(0s10)α0(c)=(01s0)\displaystyle\alpha^{1}(c)=\begin{pmatrix}0&s\\ -1&0\end{pmatrix}\implies\alpha^{0}(c)=\begin{pmatrix}0&1\\ -s&0\end{pmatrix}

Finally, when s=0s=0 i.e. over pp, we have

  1. (1)

    the kernel of α1|p\alpha^{1}|_{p} is spanned by (b+c)|p(b+c)|_{p},

  2. (2)

    the kernel of α0|p\alpha^{0}|_{p} is spanned by (bc)|p(b-c)|_{p}.

Hence, ker(α0|p)ker(α1|p)\textrm{ker}(\alpha^{0}|_{p})\neq\textrm{ker}(\alpha^{1}|_{p}) and we are done.

Theorem 4.7.

Let d=1,2d=1,2 and ee\in\mathbb{Z}.

  1. (1)

    The moduli space 𝔐d,e\mathfrak{M}_{d,e} is birational to one of the two connected components WdiW_{d}^{i} of Wd.W_{d}.

  2. (2)

    If 𝔐d,e\mathfrak{M}_{d,e} is birational to WdiW^{i}_{d}, then 𝔐d,e+1\mathfrak{M}_{d,e+1} is birational to Wd1i.W_{d}^{1-i}. In particular, the birational type of 𝔐d,e\mathfrak{M}_{d,e} only depends on dd and (e(e mod 2)2).

Proof.

In this proof, we will write Φ\Phi as Φd,e:𝔐d,e|UdWd|Ud\Phi_{d,e}:\mathfrak{M}_{d,e}|_{U_{d}}\to W_{d}|_{U_{d}} to indicate dd and ee explicitly.

Since the morphism Υ|Ud:𝔐d,e|UdUd\Upsilon|_{U_{d}}:\mathfrak{M}_{d,e}|_{U_{d}}\to U_{d} is quasi-finite (Theorem 3.10), there exists an open dense subset VUdV\subset U_{d} such that the restriction Υ|V:𝔐d,e|VV\Upsilon|_{V}:\mathfrak{M}_{d,e}|_{V}\to V is finite. Note that Corollary 4.6 implies that each fiber Υ1(C)\Upsilon^{-1}(C) for CVC\in V must be contained in a connected component of Wd|VW_{d}|_{V}. It follows that Φd,e(𝔐d,e|V)\Phi_{d,e}(\mathfrak{M}_{d,e}|_{V}) is contained in one of the connected components Wdi|VW^{i}_{d}|_{V} of Wd|VW_{d}|_{V}. Indeed, if this is not the case, the finite morphism Υ|V\Upsilon|_{V} would send the disjoint nonempty closed subsets Φd,e1(Wd1|V)\Phi_{d,e}^{-1}(W^{1}_{d}|_{V}) and Φd,e1(Wd2|V)\Phi_{d,e}^{-1}(W^{2}_{d}|_{V}) to disjoint nonempty closed subsets in VV, contradicting the irreducibility of VV. Then, the combination of Proposition 2.4, Corollary 3.12, Proposition 4.6 shows that the morphism Φd,e|Υ1(V)\Phi_{d,e}|_{\Upsilon^{-1}(V)} is a bijection on closed points from 𝔐d,e|V\mathfrak{M}_{d,e}|_{V} to Wdi|VW^{i}_{d}|_{V}. As Wdi|VW_{d}^{i}|_{V} is smooth and hence normal, the morphism Φd,e|Υ1(V)\Phi_{d,e}|_{\Upsilon^{-1}(V)} is an isomorphism and this proves part (1).

Now, for any ee\in\mathbb{Z}, suppose M=ψM^𝔐d,eM=\psi_{*}\widehat{M}\in\mathfrak{M}_{d,e} and Φd,e(M)=x1++xdkWdi\Phi_{d,e}(M)=x_{1}+...+x_{dk}\in W^{i}_{d} where xjΔ~x_{j}\in\widetilde{\Delta} and k=deg(Δ)k=\textrm{deg}(\Delta). The computation in Theorem 3.10 shows that ch(ψ(M^𝒪(pj2)))=(0,2d,e+1)\textrm{ch}\left(\psi_{*}\left(\widehat{M}\otimes\mathcal{O}\left(\frac{p_{j}}{2}\right)\right)\right)=(0,2d,e+1). By the proof of Proposition 4.6, we see that Φd,e+1(ψ(M^𝒪(pj2)))=x1++σ(xj)++xdkWd1i\Phi_{d,e+1}\left(\psi_{*}\left(\widehat{M}\otimes\mathcal{O}\left(\frac{p_{j}}{2}\right)\right)\right)=x_{1}+...+\sigma(x_{j})+...+x_{dk}\in W^{1-i}_{d}. Hence, it follows that 𝔐d,e+1\mathfrak{M}_{d,e+1} is birational to Wd1iW^{1-i}_{d} by part (1).

Remark 4.8.

The need for the assumption d=1,2d=1,2 in Theorem 4.7 can already be seen by comparing the fiber dimension of Υ|Ud:𝔐d,e|UdUd\Upsilon|_{U_{d}}:\mathfrak{M}_{d,e}|_{U_{d}}\to U_{d} and Wd|UdUdW_{d}|_{U_{d}}\to U_{d}: the dimension of the fibers of the latter is 0, while the dimension of the fibers of the former is dim(Pic(C))=g(C)\textrm{dim}(\mathop{\rm Pic}\nolimits(C))=g(C) (by Theorem 3.10) for CUdC\in U_{d}, which is positive for d3.d\geq 3.

5. Cubic threefolds

We will apply the construction of the rational map Φ:𝔐d,eWd\Phi:\mathfrak{M}_{d,e}\dashrightarrow W_{d} for the conic bundles obtained by blowing up smooth cubic threefolds along a line. As a consequence, this yields an explicit correspondence between instanton bundles on cubic threefolds and twisted Higgs bundles on the discriminant curve.

Let Y4Y\subset\mathbb{P}^{4} be a cubic threefold and l0Yl_{0}\subset Y a general line. The blow-up σ:Y~:=Bll0YY\sigma:\widetilde{Y}:=Bl_{l_{0}}Y\to Y of YY along l0l_{0} is known to be a conic bundle π:Y~2\pi:\widetilde{Y}\to\mathbb{P}^{2}. In this case, the rank 3 vector bundle is F=𝒪22𝒪2(1)F=\mathcal{O}_{\mathbb{P}^{2}}^{\oplus 2}\oplus\mathcal{O}_{\mathbb{P}^{2}}(-1) and the line bundle is L=𝒪2(1)L=\mathcal{O}_{\mathbb{P}^{2}}(-1). The discriminant curve Δ\Delta of the conic bundle π:Y~2\pi:\widetilde{Y}\to\mathbb{P}^{2} is a degree 5 curve and its étale double cover is denoted by Δ~Δ.\widetilde{\Delta}\to\Delta. Then we can consider the variety of divisors W2Δ~(10)W_{2}\subset\widetilde{\Delta}^{(10)} lying over the linear system |𝒪2(2)|Δ||\mathcal{O}_{\mathbb{P}^{2}}(2)|_{\Delta}|, its two components W2=W20W21W_{2}=W_{2}^{0}\cup W_{2}^{1} and the associated sheaf of even Clifford algebras 0\mathcal{B}_{0}, and the moduli space 𝔐d,e\mathfrak{M}_{d,e} as considered in previous sections. Note that in the case of YY, 𝒪2(2)|ΔKΔ\mathcal{O}_{\mathbb{P}^{2}}(2)|_{\Delta}\cong K_{\Delta}, so we can apply Example 2.5. Recall that in Example 2.5 the Abel-Jacobi map α~:Δ~(10)J10Δ~\widetilde{\alpha}:\widetilde{\Delta}^{(10)}\to J^{10}\widetilde{\Delta} induces the morphism α~|W21:W21Pr1\widetilde{\alpha}|_{W_{2}^{1}}:W^{1}_{2}\to\textrm{Pr}^{1} that maps birationally to the abelian variety Pr1\textrm{Pr}^{1} and the morphism α~|W20:W20Pr0\widetilde{\alpha}|_{W_{2}^{0}}:W^{0}_{2}\to\textrm{Pr}^{0} which is a generically 1\mathbb{P}^{1}-bundle over the theta divisor.

Proposition 5.1.

Let ee\in\mathbb{Z} be even. The image of Φ:𝔐2,eW2\Phi:\mathfrak{M}_{2,e}\dashrightarrow W_{2} is contained in the component W21W_{2}^{1}. In particular, 𝔐2,e\mathfrak{M}_{2,e} is birational to the Prym variety Prym(Δ~,Δ)\textrm{Prym}(\widetilde{\Delta},\Delta).

Proof.

Recall that Theorem 4.7 says that Φ\Phi maps 𝔐2,4\mathfrak{M}_{2,-4} birationally to one of the connected components W2iW^{i}_{2} of W2W_{2}, it suffices to show that 𝔐2,4\mathfrak{M}_{2,-4} cannot be birational to W20W^{0}_{2}. By the work of [17] (see Theorem 5.3 and Theorem 5.4 in the next subsection), it is known that 𝔐2,4\mathfrak{M}_{2,-4} is birational to another abelian variety, namely the intermediate Jacobian of the cubic threefold YY. In particular, the component of W2W_{2} that is birational to 𝔐2,4\mathfrak{M}_{2,-4} is birational to an abelian variety.

But recall from Example 2.5 (2) that W20W^{0}_{2} is generically a 1\mathbb{P}^{1}-bundle, which cannot happen for a variety birational to an abelian variety. Hence, the image of Φ\Phi must be contained in W21.W^{1}_{2}. It follows immediately that the composition 𝔐2,4W21α~|W21Pr1\mathfrak{M}_{2,-4}\dashrightarrow W^{1}_{2}\xrightarrow{\widetilde{\alpha}|_{W^{1}_{2}}}\textrm{Pr}^{1} is a birational map. By Theorem 4.7, the same holds for 𝔐2,e\mathfrak{M}_{2,e} when ee is even.

Proposition 5.2.

Let ee\in\mathbb{Z} be even. The image of Φ:𝔐2,e+1W2\Phi:\mathfrak{M}_{2,e+1}\dashrightarrow W_{2} is contained in the component W20W^{0}_{2} and its image in Pr0Prym(Δ~,Δ)\textrm{Pr}^{0}\cong Prym(\widetilde{\Delta},\Delta) is an open subset of the theta divisor of the Prym variety.

Proof.

This follows immediately from Theorem 4.7, Proposition 5.1, and Example 2.5. ∎

5.1. Instanton bundles on cubic threefolds and twisted Higgs bundles

A rank 2 vector bundle EE on YY is called an instanton bundle of minimal charge if EE is Gieseker semistable and c1(E)=0,c2(E)=2c_{1}(E)=0,c_{2}(E)=2 and c3(E)=0.c_{3}(E)=0. We will simply call it an instanton bundle for the rest of this section.

It is known that (see e.g. [12]) there exist the moduli space of stable instanton bundles 𝔐Y\mathfrak{M}_{Y} and its compactification by the moduli space of semistable instanton sheaves 𝔐¯Y\overline{\mathfrak{M}}_{Y}. Now, the intermediate Jacobian J(Y)J(Y) of a cubic threefold YY has birationally a modular interpretation as the moduli space 𝔐Y\mathfrak{M}_{Y} of instanton bundles, via Serre’s construction by the works of Markushevich, Tikhomirov, Iliev and Druel:

Theorem 5.3 ([19][13][12][6]).

The compactification of 𝔐Y\mathfrak{M}_{Y} by the moduli space 𝔐¯Y\overline{\mathfrak{M}}_{Y} of rank 2 semistable sheaves with c1=0,c2=2,c3=0c_{1}=0,c_{2}=2,c_{3}=0 is isomorphic to the blow-up of J(Y)J(Y) along a translate of F(Y)-F(Y). Moreover, it induces an open immersion of 𝔐Y\mathfrak{M}_{Y} into J(Y).J(Y).

We recall a theorem in [17] relating instanton bundles and 0\mathcal{B}_{0}-modules. Recall that we can embed the Fano surface of lines F(Y)F(Y) in J(Y)J(Y) as F(Y)Alb(F(Y))J(Y)F(Y)\hookrightarrow Alb(F(Y))\xrightarrow{\sim}J(Y) by picking l0l_{0} as the base point. We denote by F(Y)¯\overline{F(Y)} the strict transform of F(Y)F(Y) under the blow-up in Theorem 5.3.

Theorem 5.4 ([17]).

The moduli space 𝔐2,4\mathfrak{M}_{2,-4} is isomorphic to the blow-up of 𝔐¯Y\overline{\mathfrak{M}}_{Y} along the strict transform F(Y)¯\overline{F(Y)} of F(Y).F(Y). In particular, 𝔐Y\mathfrak{M}_{Y} is birational to 𝔐2,4.\mathfrak{M}_{2,-4}.

For a stable instanton bundle E𝔐¯YE\in\overline{\mathfrak{M}}_{Y}, the image of EE in 𝔐2,4\mathfrak{M}_{2,-4} under the birational map in Theorem 5.4 is induced by a functor Ξ:𝒦u(Y)Db(2,0)\Xi:\mathcal{K}u(Y)\hookrightarrow\textrm{D}^{b}(\mathbb{P}^{2},\mathcal{B}_{0}) which can be described explicitly as follows. First, we define the functor

Ψ:Db(Y~)Db(2,0),Eπ((E)𝒪Y~𝒪Y~detF[1])\Psi:\textrm{D}^{b}(\widetilde{Y})\to\textrm{D}^{b}(\mathbb{P}^{2},\mathcal{B}_{0}),\quad E\mapsto\pi_{*}((E)\otimes_{\mathcal{O}_{\widetilde{Y}}}\mathcal{E}\otimes_{\mathcal{O}_{\widetilde{Y}}}\det F^{\vee}[1])

where \mathcal{E} is a rank 2 vector bundle with a natural structure of flat left π0\pi^{*}\mathcal{B}_{0}-module. For details of the definition, we refer to [14]. Then Ξ(E)=Ψ(σ(E))\Xi(E)=\Psi(\sigma^{*}(E)) which makes sense as it can be checked that instanton bundles on YY are naturally objects in 𝒦u(Y)\mathcal{K}u(Y) [17, Section 3.2]. While Ξ(E)\Xi(E) is a priori a complex, it turns out that Ξ(E)\Xi(E) is concentrated in only one degree [17, Lemma 3.9], so Ξ(E)\Xi(E) is indeed a 0\mathcal{B}_{0}-module.

On the other hand, recall that for an étale double cover p:Δ~Δp:\widetilde{\Delta}\to\Delta, there is an associated 2-torsion line bundle π:ξΔ\pi:\xi\to\Delta such that Δ~\widetilde{\Delta} is recovered as the cyclic cover of ξ\xi and the section 1ξ2𝒪Δ1\in\xi^{\otimes 2}\cong\mathcal{O}_{\Delta} i.e. Δ~\widetilde{\Delta} is embedded in Tot(ξ)\textbf{Tot}(\xi) as the zero locus of t2π1t^{\otimes 2}-\pi^{*}1 where tt is the tautological section of πξ.\pi^{*}\xi. Recall that a ξ\xi-twisted SL2SL_{2}-Higgs bundle on a curve Σ\Sigma is a pair (V,ϕ)(V,\phi) consisting of a rank 2 vector bundle VV with a fixed determinant line bundle LL and ϕH0(Σ,nd0(V)ξ)\phi\in H^{0}(\Sigma,\mathcal{E}nd_{0}(V)\otimes\xi). Since we will only deal with this case, We simply call it a twisted Higgs bundle. The spectral correspondence [7] says that pushing forward a line bundle NN on Δ~\widetilde{\Delta} gives a twisted Higgs bundle (pN,pt)(p_{*}N,p_{*}t) on Δ\Delta. In fact, Prym(Δ~,Δ)\textrm{Prym}(\widetilde{\Delta},\Delta) parametrizes all twisted Higgs bundles on Δ\Delta with the spectral curve defined by t2π1t^{\otimes 2}-\pi^{*}1. Since the Hitchin base H0(Δ,ξ2)=H0(Δ,𝒪Δ)=H^{0}(\Delta,\xi^{\otimes 2})=H^{0}(\Delta,\mathcal{O}_{\Delta})=\mathbb{C}, all smooth spectral curves (defined away from 00\in\mathbb{C}) are isomorphic to each other.

Combining the functor Ξ\Xi which induces a birational map 𝔐Y𝔐2,4\mathfrak{M}_{Y}\dashrightarrow\mathfrak{M}_{2,-4}, the birational map Φ:𝔐2,4W21\Phi:\mathfrak{M}_{2,-4}\dashrightarrow W^{1}_{2}, the Abel-Jacobi map α~:Δ~(10)J10Δ~\widetilde{\alpha}:\widetilde{\Delta}^{(10)}\to J^{10}\widetilde{\Delta} and the spectral correspondence, we obtain an explicit correspondence between instanton bundles on YY and ξ\xi-twisted Higgs bundles on Δ\Delta:

Instanton bundles on YY0\mathcal{B}_{0}-modules on 2\mathbb{P}^{2} Line bundles on Δ~\widetilde{\Delta}twisted Higgs bundles on Δ\Delta with spectral curve Δ~Δ\widetilde{\Delta}\to\DeltaΞ\Xiα~Φ\widetilde{\alpha}\circ\Phipp_{*}

The correspondences of different objects here hold as birational maps between the corresponding moduli spaces.

References

  • [1] Ibrahim Assem, Daniel Simson and Andrzej Skowroński “Elements of the representation theory of associative algebras. Vol. 1” Techniques of representation theory 65, London Mathematical Society Student Texts Cambridge University Press, Cambridge, 2006, pp. x+458 DOI: 10.1017/CBO9780511614309
  • [2] Asher Auel, Marcello Bernardara and Michele Bolognesi “Fibrations in complete intersections of quadrics, Clifford algebras, derived categories, and rationality problems” In J. Math. Pures Appl. (9) 102.1, 2014, pp. 249–291 DOI: 10.1016/j.matpur.2013.11.009
  • [3] Arend Bayer, Marti Lahoz, Emanuele Macrì, Howard Nuer, Alexander Perry and Paolo Stellari “Stability conditions in families” In Publ. Math. Inst. Hautes Études Sci. 133, 2021, pp. 157–325 DOI: 10.1007/s10240-021-00124-6
  • [4] Arend Bayer, Martí Lahoz, Emanuele Macrì and Paolo Stellari “Stability conditions on Kuznetsov components”, 2017 arXiv:1703.10839 [math.AG]
  • [5] Arnaud Beauville “Sous-variétés spéciales des variétés de Prym” In Compositio Mathematica 45.3, 1982, pp. 357–383
  • [6] Arnaud Beauville “Vector bundles on the cubic threefold” In Symposium in Honor of C. H. Clemens (Salt Lake City, UT, 2000) 312, Contemp. Math. Amer. Math. Soc., Providence, RI, 2002, pp. 71–86 DOI: 10.1090/conm/312/04987
  • [7] Arnaud Beauville, M.. Narasimhan and S. Ramanan “Spectral curves and the generalised theta divisor” In J. Reine Angew. Math. 398, 1989, pp. 169–179 DOI: 10.1515/crll.1989.398.169
  • [8] Marcello Bernardara, Emanuele Macrì, Sukhendu Mehrotra and Paolo Stellari “A categorical invariant for cubic threefolds” In Adv. Math. 229.2, 2012, pp. 770–803 DOI: 10.1016/j.aim.2011.10.007
  • [9] Niels Borne “Fibrés paraboliques et champ des racines” In Int. Math. Res. Not. IMRN, 2007, pp. Art. ID rnm049\bibrangessep38 DOI: 10.1093/imrn/rnm049
  • [10] Charles Cadman “Using stacks to impose tangency conditions on curves” In Amer. J. Math. 129.2, 2007, pp. 405–427 DOI: 10.1353/ajm.2007.0007
  • [11] C Herbert Clemens and Phillip A Griffiths “The intermediate Jacobian of the cubic threefold” In Annals of Mathematics JSTOR, 1972, pp. 281–356
  • [12] S. Druel “Espace des modules des faisceaux de rang 2 semi-stables de classes de Chern c 1 = 0, c 2 = 2 et c 3 = 0 sur la cubique de 4\mathbb{P}^{4} In International Mathematics Research Notices 2000.19, 2000, pp. 985–1004
  • [13] A. Iliev and D. Markushevich “The Abel-Jacobi map for cubic threefold and periods of Fano threefolds of degree 14.” In Documenta Mathematica 5 Universiät Bielefeld, Fakultät für Mathematik, 2000, pp. 23–47 URL: http://eudml.org/doc/48423
  • [14] Alexander Kuznetsov “Derived categories of quadric fibrations and intersections of quadrics” In Advances in Mathematics 218.5 Elsevier, 2008, pp. 1340–1369
  • [15] Alexander Kuznetsov “Derived categories of cubic fourfolds” In Cohomological and geometric approaches to rationality problems 282, Progr. Math. Birkhäuser Boston, Boston, MA, 2010, pp. 219–243 DOI: 10.1007/978-0-8176-4934-0“˙9
  • [16] Alexander Kuznetsov “Instanton bundles on Fano threefolds” In Cent. Eur. J. Math. 10.4, 2012, pp. 1198–1231 DOI: 10.2478/s11533-012-0055-1
  • [17] Martí Lahoz, Emanuele Macrì and Paolo Stellari “Arithmetically Cohen-Macaulay bundles on cubic threefolds” In Algebr. Geom. 2.2, 2015, pp. 231–269 DOI: 10.14231/AG-2015-011
  • [18] J. Le Potier “Faisceaux semi-stables de dimension 11 sur le plan projectif” In Rev. Roumaine Math. Pures Appl. 38.7-8, 1993, pp. 635–678
  • [19] Dimitri Markushevich and Alexander Tikhomirov “The Abel-Jacobi map of a moduli component of vector bundles on the cubic threefold” In Journal of Algebraic Geometry 10, 1998
  • [20] David Mumford “Theta characteristics of an algebraic curve” In Ann. Sci. École Norm. Sup. (4) 4, 1971, pp. 181–192 URL: http://www.numdam.org/item?id=ASENS_1971_4_4_2_181_0
  • [21] David Mumford “Prym varieties. I” In Contributions to analysis (a collection of papers dedicated to Lipman Bers), 1974, pp. 325–350
  • [22] Carlos T. Simpson “Moduli of representations of the fundamental group of a smooth projective variety. I” In Inst. Hautes Études Sci. Publ. Math., 1994, pp. 47–129 URL: http://www.numdam.org/item?id=PMIHES_1994__79__47_0
  • [23] G.. Welters “Abel-Jacobi isogenies for certain types of Fano threefolds” 141, Mathematical Centre Tracts Mathematisch Centrum, Amsterdam, 1981, pp. i+139