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

On the moduli spaces of parabolic symplectic/orthogonal bundles on curves

Jianping Wang Address of author: School of Mathematical Sciences, University of Science and Technology of China, Hefei, 230026, China. [email protected]  and  Xueqing Wen Address of author: Yau Mathematical Sciences Center, Beijing, 100084, China. [email protected]
Abstract.

We prove that the moduli spaces of parabolic symplectic/orthogonal bundles on a smooth curve are globally F regular type. As a consequence, all higher cohomology of theta line bundle vanish. During the proof, we develop a method to estimate codimension, and consider the infinite grassmannians for parabolic GG bundles.

1. Introduction

Let XX be a variety over an algebraically closed field of positive characteristic and FX:XXF_{X}:X\rightarrow X be the absolute Frobenius map. In [14] Mehta and Ramanathan introduced the notion “F split”: XX is said to be F split if the natural morphism FX#:𝒪XFX𝒪XF_{X}^{\#}:\mathcal{O}_{X}\rightarrow F_{X*}\mathcal{O}_{X} splits as an 𝒪X\mathcal{O}_{X} module morphism. Later in [21] Smith studied a special kind of F split varieties: globally F regular varieties(see Section 6 for details). F split varieties and globally F regular varieties have many nice properties, for example, the vanishing of higher cohomologies of ample line bundles(nef line bundles in the case of globally F regular varieties).

Although almost all varieties are not F split, some important kind of varieties are, such as flag varieties, toric varieties. In [13] Mehta and Ramadas proved that the moduli space of semistable parabolic rank two vector bundles with fixed determinant on a generic nonsingular projective curve, are F split. They conjectured that the generic condition can be moved. On the other hand, as mentioned in [25], this conjecture should be extended into the following: the moduli spaces of semistable parabolic bundles with fixed determinant on any nonsingular projective curve are globally F regular.

In [25] Sun and Zhou studied the characteristic zero analogy of this extended conjecture. A variety over a field of characteristic zero is said to be of globally F regular type if its modulo pp reduction are globally F regular for all p0p\gg 0. They proved that the moduli spaces of semistable parabolic vector bundles on a smooth projective curve over an algebraically closed field of characteristic zero are of globally F regular type. As an application, they can give a finite dimensional proof of the so called Verlinde formula in GLnGL_{n} and SLnSL_{n} case([24]).

Globally F regular type varieties have similar vanishing properties as globally F regular varieties, namely all higher cohomologies of nef line bundles are vanishing. Unlike the positive characteristic case, in characteristic zero, all Fano varieties with rational singularities are globally F regular type varieties([21]). So globally F regular type varieties can be regarded as a generalization of Fano varieties in characteristic zero, with the vanishing properties retained and hence it is interesting to find examples of globally F regular type varieties.

On the other hand, properties of moduli spaces is a central topic in the study of moduli problems. We already know that for connected simply connected algebraic group GG, the moduli space of semistable GG bundles on a smooth curve is a Fano variety ([11]). However, if one consider the moduli space of semistable GG bundles with parabolic structure on a smooth curve, then one may not get a Fano variety. As mentioned before, in the case of G=SLnG=SL_{n}, Sun and Zhou proved that the moduli spaces of semistable parabolic vector bundles with fixed determinant are globally F regular type varieties([25]). So it encourage us to consider globally F regularity as a reasonable property of moduli spaces of GG bundles with parabolic structure on curves.

In this paper, we consider parabolic symplectic and orthogonal bundles over smooth curves. Our main theorem is the following:

Theorem 1.1 (Main theorem, see Theorem 6.5).

The moduli spaces of semistable parabolic symplectic/orthogonal bundles over any smooth projective curve are globally F regular type varieties. As a consequence, any higher cohomologies of nef line bundles on these moduli spaces vanish.

We now describe how this paper is organized:

In Section 2, we recall some basics about parabolic vector bundles, parabolic symplectic/orthogonal bundles and the equivalence between parabolic bundles and orbifold bundles.

In Section 3, we construct the moduli space of semistable parabolic symplectic/orthogonal bundles explicitly, using Geometric Invariant Theory.

In Section 4, we develop a technique to estimate the codimension of unsemistable locus in a given family, not only for parabolic symplectic/orthogonal bundles, but also GG bundles and parabolic vector bundles.

In Section 5, to evaluate the canonical line bundle on the moduli spaces we constructed in Section 3, we introduce the infinite Grassmannians for parabolic GG bundles, here GG is a connected simply connected simple algebraic group; we also define the theta line bundles for any family of symplectic/orthogonal bundle then we can show that under certain choice of rank and weights, the moduli spaces we constructed in Section 3 are Fano varieties.

In Section 6, we recall definition and properties of globally F regular type varieties, with the help of key Proposition 6.9, we can prove our main theorem.

Acknowledgements We would like to thank our supervisor, Prof. Xiaotao Sun, who brought this problem to us and kindly answer our questions. The second author would like to thank Dr. Bin Wang and Dr. Xiaoyu Su, for helpful discussions.

2. Basics of parabolic principal bundle over curve

2.1. Parabolic vector bundles and parabolic symplectic/orthogonal bundles

Let CC be a smooth projective curve of genus g0g\geq 0. We fix a reduced effective divisor DD of CC, and an integer K>0K>0.

EE is a vector bundle of rank rr and degree dd over CC, by a parabolic structure on EE, we mean the following:

  1. (1)

    At each xDx\in D, we have a choice of flag of ExE_{x}:

    0=Flx(Ex)Flx1(Ex)F0(Ex)=Ex0=F_{l_{x}}(E_{x})\subseteq F_{l_{x}-1}(E_{x})\subseteq\cdots\subseteq F_{0}(E_{x})=E_{x}

    Let ni(x)=dimFi1(Ex)/Fi(Ex)n_{i}(x)=\text{dim}F_{i-1}(E_{x})/F_{i}(E_{x}) and n(x)=(n1(x),n2(x),,nlx(x))\overrightarrow{n}(x)=\big{(}n_{1}(x),n_{2}(x),\cdots,n_{l_{x}}(x)\big{)}. Notice that all these filtrations together are equivalent to a filtration:

    E(D)=Fl(E)Fl1(E)F0(E)=EE(-D)=F_{l}(E)\subseteq F_{l-1}(E)\subseteq\cdots\subseteq F_{0}(E)=E
  2. (2)

    At each xDx\in D, we fix a choice of sequence of integers, which are called weights:

    0a1(x)<a2(x)<alx(x)<K0\leq a_{1}(x)<a_{2}(x)\cdots<a_{l_{x}}(x)<K

    Put a(x)=(a1(x),a2(x),,alx(x))\overrightarrow{a}(x)=\big{(}a_{1}(x),a_{2}(x),\cdots,a_{l_{x}}(x)\big{)}.

We say that (E,D,K,{n(x)}xD,{a(x)}xD)\big{(}E,D,K,\{\overrightarrow{n}(x)\}_{x\in D},\{\overrightarrow{a}(x)\}_{x\in D}\big{)}, or simply EE, is a parabolic vector bundle, and σ=({n(x)}xD,{a(x)}xD)\sigma=\big{(}\{\overrightarrow{n}(x)\}_{x\in D},\{\overrightarrow{a}(x)\}_{x\in D}\big{)} is the parabolic type of EE.

For any subbundle FF of the vector bundle EE, it is clearly that there is an induced parabolic structure on FF, with induced flags structures and same weights; similarly there is an induced parabolic structure on E/FE/F.

Let E1E_{1} and E2E_{2} be two parabolic vector bundle with same weights, the space of parabolic homomorphisms Hompar(E1,E2)\text{Hom}_{par}(E_{1},E_{2}) given by 𝒪C\mathcal{O}_{C}-homomorphisms between E1E_{1} and E2E_{2} preserving filtrations at each xDx\in D. We can also define parabolic sheaf of parabolic homomorphisms ompar(E1,E2)\mathcal{H}om_{par}(E_{1},E_{2}) in a similar way, which inherits a parabolic structure naturally. In fact, in [26], it is shown that the category of parabolic bundles is contained in an abelian category with enough injectives. So we have the derived functors of parabolic homomorphism. We use Extpar1(E1,E2)\text{Ext}^{1}_{par}(E_{1},E_{2}) to denote the space of parabolic extensions.

Definition 2.1.

The parabolic degree of EE is defined by

pardegE=degE+1KxDi=1lxai(x)ni(x)pardegE=degE+\dfrac{1}{K}\sum_{x\in D}\sum_{i=1}^{l_{x}}a_{i}(x)n_{i}(x)

and EE is said to be stable(resp. semistable) if for all nontrivial subbundle FEF\subset E, concerning the induced parabolic structure, we have:

pardegFrankF<pardegErankE(resp.)\dfrac{pardegF}{rankF}<\dfrac{pardegE}{rankE}\ \ \ \ \ \ (\text{resp.}\ \leq)

Now let us talk about family of parabolic vector bundles. Let SS be a scheme of finite type, a family of parabolic vector bundle with type σ\sigma over CC parametrized by SS is a vector bundle \mathcal{E} over S×CS\times C, together with filtrations of vector bundles on x\mathcal{E}_{x} of type n(x)\overrightarrow{n}(x) and weights a(x)\overrightarrow{a}(x) for each xDx\in D. As before, such filtrations are equivalent to the following:

((S×D))=Fl()Fl1()F0()=\mathcal{E}\big{(}-(S\times D)\big{)}=F_{l}(\mathcal{E})\subseteq F_{l-1}(\mathcal{E})\subseteq\cdots\subseteq F_{0}(\mathcal{E})=\mathcal{E}

where S×DS\times D is considered as an effective divisor of S×CS\times C. Following [26], we say \mathcal{E} is a flat family if all Fi()F_{i}(\mathcal{E}) are flat families.

Definition 2.2.

EE is a vector bundle of rank rr degree dd over CC. By a symplectic/orthogonal parabolic structure on EE, we mean the following:

  1. (1)

    A non-degenerated anti-symmetric/symmetric two form

    ω:EE𝒪C(D)\omega:E\otimes E\longrightarrow\mathcal{O}_{C}(-D)
  2. (2)

    At each xDx\in D, a choice of flag:

    0=F2lx+1(Ex)F2lx(Ex)Flx+1(Ex)Flx(Ex)F0(Ex)=Ex0=F_{2l_{x}+1}(E_{x})\subseteq F_{2l_{x}}(E_{x})\subseteq\cdots F_{l_{x}+1}(E_{x})\subseteq F_{l_{x}}(E_{x})\subseteq\cdots\subseteq F_{0}(E_{x})=E_{x}

    where Fi(Ex)F_{i}(E_{x}) are isotropic subspaces of ExE_{x} respect to the form ω\omega and F2lx+1i(Ex)=Fi(Ex)F_{2l_{x}+1-i}(E_{x})=F_{i}(E_{x})^{\perp} for lx+1i2lx+1l_{x}+1\leq i\leq 2l_{x}+1.

  3. (3)

    At each xDx\in D, we fix a choice of weights:

    0a1(x)<a2(x)<alx(x)<alx+1(x)<<a2lx+1(x)K0\leq a_{1}(x)<a_{2}(x)\cdots<a_{l_{x}}(x)<a_{l_{x}+1}(x)<\cdots<a_{2l_{x}+1}(x)\leq K

    satisfying ai(x)+a2lx+2i(x)=Ka_{i}(x)+a_{2l_{x}+2-i}(x)=K, 1ilx+11\leq i\leq l_{x}+1.

As before, we put ni(x)=dim(Fi1(Ex)/Fi(Ex))n_{i}(x)=\text{dim}\big{(}F_{i-1}(E_{x})/F_{i}(E_{x})\big{)} , and

n(x)\displaystyle\overrightarrow{n}(x) =(n1(x),n2(x),,n2lx+1(x))\displaystyle=\big{(}n_{1}(x),n_{2}(x),\cdots,n_{2l_{x}+1}(x)\big{)}
a(x)\displaystyle\overrightarrow{a}(x) =(a1(x),a2(x),,a2lx+1(x))\displaystyle=\big{(}a_{1}(x),a_{2}(x),\cdots,a_{2l_{x}+1}(x)\big{)}

We say that (E,ω,D,K,{n(x)}xD,{a(x)}xD)\big{(}E,\omega,D,K,\{\overrightarrow{n}(x)\}_{x\in D},\{\overrightarrow{a}(x)\}_{x\in D}\big{)}, or simply EE, is a parabolic symplectic/orthogonal bundle and σ=({n(x)}xD,{a(x)}xD)\sigma=\big{(}\{\overrightarrow{n}(x)\}_{x\in D},\{\overrightarrow{a}(x)\}_{x\in D}\big{)} is the parabolic type of EE.

Convention: when talked about parabolic symplectic/orthogonal bundles, we always assume that degDD is even, and we fix a line bundle LL over CC and an isomorphism L2𝒪C(D)L^{\otimes 2}\cong\mathcal{O}_{C}(D).

Remark 2.3.
  1. (1)

    The original definition of parabolic principal bundles is just a principal bundle together with additional structures [18]. Later in [1] Balaji, Biswas and Nagaraj establish a different definition, which share some nice properties as in the case of parabolic vector bundles, for example, a parabolic symplectic/orthogonal bundle admits an Einstein–Hermitian connection if and only if it is polystable([6]).

  2. (2)

    Although in our definition, EE is not a principal symplectic/orthogonal bundle, but ELE\otimes L is, we call EE twisted orthogonal/symplectic bundle.

The parabolic degree of EE is given by

pardegE=degE+1KxDi=12lx+1ai(x)ni(x)\displaystyle pardegE=degE+\dfrac{1}{K}\sum_{x\in D}\sum_{i=1}^{2l_{x}+1}a_{i}(x)n_{i}(x)

By relations between n(x)\overrightarrow{n}(x) and a(x)\overrightarrow{a}(x) we see that pardegE=degE+r2degDpardegE=degE+\frac{r}{2}degD, noticing that ω:EE𝒪X(D)\omega:E\otimes E\rightarrow\mathcal{O}_{X}(-D) is non-degenerated, so EE(D)E\simeq E^{\vee}(D). Thus degE+r2degD=0degE+\dfrac{r}{2}degD=0 and then pardegE=0pardegE=0.

For any subbundle FF of EE, we can define the parabolic degree of FF by

pardegF=degF+1KxDi=12lx+1ai(x)niF(x)pardegF=degF+\dfrac{1}{K}\sum_{x\in D}\sum_{i=1}^{2l_{x}+1}a_{i}(x)n_{i}^{F}(x)

where niF(x)=dim(Fi1(Ex)Fx/Fi(Ex)Fx)n_{i}^{F}(x)=\text{dim}\big{(}F_{i-1}(E_{x})\cap F_{x}/F_{i}(E_{x})\cap F_{x}\big{)} .

Definition 2.4.

A parabolic orthogonal/symplectic bundle EE is said to be stable(resp. semistable) if for all nontrivial isotropic subbundle FEF\subset E(by isotropic we mean ω(FF)=0\omega(F\otimes F)=0), we have

pardegF<0(resp.)pardegF<0\ \ \ \ (\text{resp.}\ \leq)
Lemma 2.5.

A parabolic symplectic/orthogonal bundle is semistable iff for any subbundle FF, not necessarily isotropic, we have pardegF0pardegF\leq 0, i.e. semistable as a parabolic vector bundle.

Proof.

If E is semistable as a parabolic vector bundle, then it is semistable as parabolic symplectic/orthogonal bundle.

Conversely, if E is a semistable parabolic symplectic/orthogonal bundle and a subbundle F is given. We want to show that pardeg(F)0pardeg(F)\leq 0.

If FF=0F\cap F^{\perp}=0, then E=FFE=F\oplus F^{\perp}. Hence 2pardeg(F)=pardeg(F)+pardeg(F)=deg(E)=02pardeg(F)=pardeg(F)+pardeg(F^{\perp})=deg(E)=0 and we are done.

If FF0F\cap F^{\perp}\neq 0, then we have the exact sequece of parabolic bundles:

0FFFFF+F00\rightarrow F\cap F^{\perp}\rightarrow F\oplus F^{\perp}\rightarrow F+F^{\perp}\rightarrow 0

This shows that

pardeg(FF)+pardeg(F+F)=pardeg(F)+pardeg(F)=2pardeg(F)pardeg(F\cap F^{\perp})+pardeg(F+F^{\perp})=pardeg(F)+pardeg(F^{\perp})=2pardeg(F)

It is easy to see pardeg(FF)pardeg(F+F)pardeg(F\cap F^{\perp})\geq pardeg(F+F^{\perp}) and hence we have 2pardeg(F)2pardeg(FF)02pardeg(F)\leq 2pardeg(F\cap F^{\perp})\leq 0, where pardeg(FF)0pardeg(F\cap F^{\perp})\leq 0 since FFF\cap F^{\perp} is isotropic. ∎

2.2. Equivalence between parabolic bundles and orbifold bundles

There is an interesting and useful correspondence between parabolic bundles and orbifold bundles, which is developed in [15], and [4] for general case. We will recall the correspondence briefly as follows:

Given C,D,KC,D,K as before, By Kawamata covering, there is a smooth projective curve YY and a morphism p:YCp:Y\rightarrow C such that pp is only ramified over DD with pD=KxDp1(x)p^{*}D=K\sum_{x\in D}p^{-1}(x), moreover, if we put Γ=Gal(Rat(Y)/Rat(C))\Gamma=\text{Gal}\big{(}Rat(Y)/Rat(C)\big{)} to be the Galois group, then pp is exactly the quotient map of YY by Γ\Gamma.

Definition 2.6.

An orbifold bundle over YY is a vector bundle WW over YY such that the action of Γ\Gamma lifts to WW.

And an orbifold symplectic/orthogonal bundle is an orbifold bundle such that the correspondence 2-form ω\omega is a morphism of orbifold bundles.

Given an orbifold bundle WW, for any y=p1(x)pDy=p^{-1}(x)\in p^{*}D, the stabilizer Γy\Gamma_{y}, which is a cyclic group of order KK, acts on the fiber WyW_{y} by some representation(after choosing suitable basis):

ξKdiag{ξKa1(x),,ξKa1(x),ξKa2(x),,ξKalx(x)}\xi_{K}\longmapsto\text{diag}\{\xi_{K}^{a_{1}(x)},\cdots,\xi_{K}^{a_{1}(x)},\xi_{K}^{a_{2}(x)},\cdots,\xi_{K}^{a_{l_{x}}(x)}\}

where 0a1(x)<a2(x)alx(x)<K0\leq a_{1}(x)<a_{2}(x)\cdots a_{l_{x}}(x)<K are integers, ξK\xi_{K} is the KK-th root of unity and the multiplicity of ξKai(x)\xi_{K}^{a_{i}(x)} is given by ni(x)n_{i}(x). Similar in the definition of parabolic bundle, we use σ=({n(x)}xD,{a(x)}xD)\sigma=\big{(}\{\overrightarrow{n}(x)\}_{x\in D},\{\overrightarrow{a}(x)\}_{x\in D}\big{)} to denote the type of orbifold bundle WW.

Proposition 2.7 ([15],[4]).

There is an equivalence between the category of orbifold bundles over YY with type σ\sigma and the category of parabolic vector bundles over CC with type σ\sigma.

Roughly speaking, given an orbifold bundle WW, then (pW)Γ(p_{*}W)^{\Gamma} is a parabolic vector bundle over CC, with parabolic structures given by the action of stablizers. Conversely, EE is a parabolic vector bundle, we put W1=pEW_{1}=p^{*}E, after some elementary transformations of W1W_{1}, we would have an orbifold bundle of type σ\sigma. Moreover, we have

#ΓpardegE=degW\#\Gamma\cdot pardegE=degW

and E is (semi)stable as parabolic bundle if and only if W is (semi)stable as orbifold bundle.

Now we will talk about orbifold symplectic/orthogonal bundles over YY: an orbifold symplectic/orthogonal bundles is symplectic/orthogonal bundle WW over YY such that the action of Γ\Gamma lifts to WW compatible with the symplectic/orthogonal structure. For any y=p1(x)p(D)y=p^{-1}(x)\in p*(D), the action of stabilizer is given by:

ξKdiag{ξKa1(x),,ξKa1(x),ξKa2(x),,ξKalx(x),ξKalx(x),,ξKa1(x)}\xi_{K}\longmapsto\text{diag}\{\xi_{K}^{a_{1}(x)},\cdots,\xi_{K}^{a_{1}(x)},\xi_{K}^{a_{2}(x)},\cdots,\xi_{K}^{a_{l_{x}}(x)},\xi_{K}^{-a_{l_{x}}(x)},\cdots,\xi_{K}^{-a_{1}(x)}\}

As before, we use σ\sigma to denote the type of this orbifold symplectic bundle. Similarly, we have:

Proposition 2.8.

There is an equivalence between the category of orbifold symplectic/orthogonal bundles over YY with type σ\sigma and the category of parabolic symplectic/orthogonal bundles over CC with type σ\sigma. Moreover, this equivalence induces an equivalence between orbifold isotropic subbundles and isotropic subbundles.

Proof.

See [6]. ∎

3. Moduli space of semistable parabolic symplectic/orthogonal bundles

In this section, we construct the moduli space of semistable parabolic symplectic/orthogonal bundles with fixed parabolic type σ\sigma over CC. Although the moduli space is already constructed in [3] for general algebraic groups, but for our purpose, we will construct the moduli spaces explicitly using GIT constructions.

3.1. Construction of the moduli space

In this section we will use EE to denote a parabolic symplectic/orthogonal bundle of rank rr, degree dd and parabolic type σ\sigma.

We will fix an ample line bundle 𝒪(1)\mathcal{O}(1) on CC with degree cc, then the Hilbert polynomial of E is PE(m)=crm+χ(E).P_{E}(m)=crm+\chi(E).

Firstly we notice that by Lemma 2.3 of [8], the class of semistable parabolic orthogonal/symplectic bundles with fixed rank, degree and parabolic type are bounded. So we may choose an integer N0N_{0} large enough so that E(N)E(N) is globally generated for all semistable parabolic bundle EE and all integers NN0N\geq N_{0} ; which means, we have a quotient

q:V𝒪X(N)Eq:V\otimes\mathcal{O}_{X}(-N)\twoheadrightarrow E

where V is the vector space P(N)\mathbb{C}^{P(N)} and PP is the Hilbert polynomial of E.

Let QQ be the Quot scheme of quotients of V𝒪X(N)V\otimes\mathcal{O}_{X}(-N) with Hilbert polynomial PP.

The orthogonal/symplectic structure on EE will induce a morphism :

(V𝒪C)(V𝒪C)E(N)E(N)𝒪C(2ND)(V\otimes\mathcal{O}_{C})\otimes(V\otimes\mathcal{O}_{C})\longrightarrow E(N)\otimes E(N)\longrightarrow\mathcal{O}_{C}(2N-D)

which is equivalent to a bilinear map on VV:

ϕ:VVH0(C,𝒪C(2ND))\phi:V\otimes V\longrightarrow\text{H}^{0}(C,\mathcal{O}_{C}(2N-D))

here 𝒪C(2ND)=𝒪C(2N)𝒪C(D)\mathcal{O}_{C}(2N-D)=\mathcal{O}_{C}(2N)\otimes\mathcal{O}_{C}(-D) and we use HH to denote the space H0(C,𝒪C(2ND))\text{H}^{0}(C,\mathcal{O}_{C}(2N-D)).

Now we can regard every semistable EE as a point in the space Q×Hom(VV,H).Q\times\mathbb{P}Hom(V\otimes V,H). However, not every element in Hom(VV,H)\mathbb{P}Hom(V\otimes V,H) would give a nondegenerated form on EE. To fix this, we will use the following lemma:

Lemma 3.1 (Lemma 3.1 of [8]).

Let XX be a smooth projective variety and YY be a scheme. Consider a morphism of sheaves f:f:\mathcal{E}\rightarrow\mathcal{F} over X×YX\times Y, moreover, we assume \mathcal{F} is flat over YY. Then there is a unique closed subscheme ZZ of YY satisfying the following universal property: for any scheme SS and a Cartesian diagram:

X×S\textstyle{X\times S\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}h¯\scriptstyle{\bar{h}}pS\scriptstyle{p_{S}}X×Y\textstyle{X\times Y\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p\scriptstyle{p}S\textstyle{S\ignorespaces\ignorespaces\ignorespaces\ignorespaces}h\scriptstyle{h}Y\textstyle{Y}

then f¯(f)=0\bar{f}^{*}(f)=0 if and only if hh factors through ZZ.

Now we let ZQ×Hom(VV,H)Z\subset Q\times\mathbb{P}Hom(V\otimes V,H) be the closed subscheme such that every closed point (q:V𝒪X(N)E,ϕ:VVHq:V\otimes\mathcal{O}_{X}(-N)\twoheadrightarrow E,\phi:V\otimes V\rightarrow H) of ZZ represents a twisted symplectic/orthogonal bundle EE.

So over Z×CZ\times C, we have a universal quotient q:VpC𝒪C(N)0q:V\otimes p_{C}^{*}\mathcal{O}_{C}(-N)\rightarrow\mathcal{E}\rightarrow 0 on X×ZX\times Z and a nondegenerated anti-symmetric/symmetric two form ω:pC𝒪C(D)\omega:\mathcal{E}\otimes\mathcal{E}\rightarrow p_{C}^{*}\mathcal{O}_{C}(-D) where pC:Z×CCp_{C}:Z\times C\rightarrow C is the projection. For any xDx\in D, let x\mathcal{E}_{x} be the restriction of \mathcal{E} on Z×{x}ZZ\times\{x\}\cong Z and we put Flagn(x)(x)ZFlag_{\vec{n}(x)}(\mathcal{E}_{x})\rightarrow Z be the relative isotropic flag scheme of type n(x)\vec{n}(x).

Let :=×ZxDFlagn(x)(x)Z,\mathcal{R}:=\underset{x\in D}{\times_{Z}}Flag_{\vec{n}(x)}(\mathcal{E}_{x})\rightarrow Z, then a closed point of \mathcal{R} is represented by

((q,ϕ),(q1(x),q2(x),,q2lx(x))xD)((q,\phi),(q_{1}(x),q_{2}(x),\dots,q_{2l_{x}}(x))_{x\in D})

where (q,ϕ)(q,\phi) is a point of Z, and qi(x)q_{i}(x) is the composition qi(x):V𝒪X(N)EExQi(x)q_{i}(x):V\otimes\mathcal{O}_{X}(-N)\rightarrow E\rightarrow E_{x}\twoheadrightarrow Q_{i}(x). We denote by Qi(x)Q_{i}(x) the quotients Ex/Fi(E)x{E_{x}}/{F_{i}(E)_{x}}, and let ri(x)=dimQi(x).r_{i}(x)=dimQ_{i}(x).

For mm large enough, let 𝒢=GrassP(m)(VWm)×Hom(VV,H)×Flag\mathcal{G}=Grass_{P(m)}(V\otimes W_{m})\times\mathbb{P}Hom(V\otimes V,H)\times\textbf{Flag}, where, Wm=H0(V𝒪(mN))W_{m}=H^{0}(V\otimes\mathcal{O}(m-N)), and Flag is defined as:

Flag=xD(Grassr1(x)(V)××Grassr2lx(x)(V))\textbf{Flag}=\prod_{x\in D}\big{(}Grass_{r_{1}(x)}(V)\times\dots\times Grass_{r_{2l_{x}}(x)}(V)\big{)}

Now, consider the SL(V)SL(V)-equivariant embedding

𝒢=GrassP(m)(VWm)×Hom(VV,H)×Flag\mathcal{R}\hookrightarrow\mathcal{G}=Grass_{P(m)}(V\otimes W_{m})\times\mathbb{P}Hom(V\otimes V,H)\times\textbf{Flag}

Which maps the point ((q,ϕ),(q1(x),q2(x),,q2lx(x))xD)((q,\phi),(q_{1}(x),q_{2}(x),\dots,q_{2l_{x}}(x))_{x\in D}) of \mathcal{R} to the point

(g,ϕ,(g1(x),g2(x),,g2lx(x))xD)(g,\phi,(g_{1}(x),g_{2}(x),\dots,g_{2l_{x}}(x))_{x\in D})

of 𝒢\mathcal{G}, where g:VWmH0(E(mN))g:V\otimes W_{m}\twoheadrightarrow H^{0}(E(m-N)) and gi(x):VQi(x)g_{i}(x):V\twoheadrightarrow Q_{i}(x).

We give the polarisation on 𝒢\mathcal{G} by:

n1×1×xDi=12lxdi(x)n_{1}\times 1\times\prod_{x\in D}\prod_{i=1}^{2l_{x}}d_{i}(x)

Where n1=l+KcNc(mN),di(x)=ai+1(x)ai(x)n_{1}=\mathnormal{\frac{l+KcN}{c(m-N)}},d_{i}(x)=a_{i+1}(x)-a_{i}(x) and l\mathnormal{l} is the number satisfying

xDi=12lxdi(x)ri(x)+rl=Kχ\mathnormal{\sum_{x\in D}\sum_{i=1}^{2l_{x}}d_{i}(x)r_{i}(x)+rl=K\chi}

We will analyse the action of SL(V)SL(V) on \mathcal{R} using a method in [8]. Let s\mathcal{R}^{s}(resp. ss\mathcal{R}^{ss}) to denote the sublocus of \mathcal{R} where the corresponding parabolic symplectic/orthogonal bundles are stable(resp. semistable) and the map H0(q):VH0(C,E(m))\text{H}^{0}(q):V\rightarrow\text{H}^{0}(C,E(m)) is an isomorphism. We are going to show s\mathcal{R}^{s}(respectively, ss\mathcal{R}^{ss}) is the stable (respectively, semistable) locus of the action in the sense of GIT. Firstly let us recall a definition in [8]:

Definition 3.2.

A weighted filtration (E,m)(E_{\textbf{\textbullet}},m_{\textbf{\textbullet}}) of a parabolic symplectic/orthogonal bundle EE consists of

  1. (1)

    a filtration of subsheaves

    0E1E2EtEt+1=E0\subset E_{1}\subset E_{2}\subset\dots\subset E_{t}\subset E_{t+1}=E

    We denote rk(Ei)rk(E_{i}) by sis_{i};

  2. (2)

    a sequence of positive numbers m1,m2,mtm_{1},m_{2}\dots,m_{t}, called the weights of this filtraion.

Let Γ=i=1tmiΓsir\Gamma=\sum_{i=1}^{t}m_{i}\Gamma^{s_{i}}\in\mathbb{C}^{r}, where

Γk=(kr,kr,,krk,k,krk)\Gamma^{k}=(\overbrace{k-r,k-r,\dots,k-r}^{k},\overbrace{k\dots,k}^{r-k})

Now, given a weighted filtration (E,m)(E_{\textbf{\textbullet}},m_{\textbf{\textbullet}}) of a parabolic symplectic/orthogonal bundle EE, let Γj\Gamma_{j} be the jj-th component of Γ\Gamma, and we define

μ(ω,E,m):=min{Γsi1+Γsi2:ω|Ei1Ei20}\mu(\omega,E_{\textbf{\textbullet}},m_{\textbf{\textbullet}}):=min\{\Gamma_{s_{i_{1}}}+\Gamma_{s_{i_{2}}}:\omega|_{E_{i_{1}}\otimes E_{i_{2}}}\neq 0\}

We have the following result(see [8], Lemma 5.6):

Lemma 3.3.

If ω\omega is nondegenerate, then μ(ω,E,m)0\mu(\omega,E_{\textbf{\textbullet}},m_{\textbf{\textbullet}})\leq 0.

Proof.

we can take the index i and j such that μ(ω,E,m)=Γsi+Γsj\mu(\omega,E_{\textbf{\textbullet}},m_{\textbf{\textbullet}})=\Gamma_{s_{i}}+\Gamma_{s_{j}} and ω|EiEj0\omega|_{E_{i}\otimes E_{j}}\neq 0. Then there exist a point xCx\in C away from DD such that the restriction ωx=ω|Ei,xEj,x0\omega_{x}=\omega|_{E_{i,x}\otimes E_{j,x}}\neq 0.

Let W=ExW=E_{x}. Then the nondegenerate form ω\omega over EE induces a nondegenerate form over the vector space WW. We still write this form as ω:WW.\omega:W\otimes W\rightarrow\mathbb{C}. By the nondegenrate condition, using Hilbert-Mumford criterion, one can see that ω¯Hom(WW,)\overline{\omega}\in\mathbb{P}Hom(W\otimes W,\mathbb{C}) is GIT semistable with the natural SL(W)SL(W) action. It implies that μ(ω,W,m)0\mu(\omega,W_{\textbf{\textbullet}},m_{\textbf{\textbullet}})\leq 0 for all weighted filtrations of W.

It’s easy to see μ(φ,E,m)μ(ω,W,m),\mu(\varphi,E_{\textbf{\textbullet}},m_{\textbf{\textbullet}})\leq\mu(\omega,W_{\textbf{\textbullet}},m_{\textbf{\textbullet}}), hence μ(ω,E,m)0\mu(\omega,E_{\textbf{\textbullet}},m_{\textbf{\textbullet}})\leq 0. ∎

In the following we use Hilbert-Mumford criterion to determine the (semi)stable locus for the action of SL(V)SL(V) of \mathcal{R}.

Proposition 3.4.

A point ((q,ϕ),(q1(x),q2(x),,q2lx(x))xD)((q,\phi),(q_{1}(x),q_{2}(x),\dots,q_{2l_{x}}(x))_{x\in D}) of \mathcal{R} is GIT stable (resp. GIT semistable) for the action of SL(V)SL(V), with respect to the polarisation defined in definition 2.1, if and only if for all weighted filtration (E,m)(E_{\textbf{\textbullet}},m_{\textbf{\textbullet}}), we have

kP(N)(i=1t(pardeg(Ei))+μ(ω,E,m)<0(resp.)kP(N)(\sum_{i=1}^{t}(pardeg(E_{i}))+\mu(\omega,E_{\textbf{\textbullet}},m_{\textbf{\textbullet}})<0\ \ (\text{resp.}\leq)
Proof.

By the Hilbert-Mumford criterion, a point ((q,ϕ),(q1(x),q2(x),,q2lx(x))xD)((q,\phi),(q_{1}(x),q_{2}(x),\dots,q_{2l_{x}}(x))_{x\in D}) is GIT semistable if and only if any one parameter subgroup λ:𝔾mSL(V)\lambda:\mathbb{G}_{m}\rightarrow SL(V), the corresponding Hilbert-Mumford weight is greater or equal than zero. But a one parameter subgroup of SL(V)SL(V) is equivalent to a weighted filtration of VV and hence gives a weight filtration (E,m)(E_{\textbf{\textbullet}},m_{\textbf{\textbullet}}) for the corresponding bundle EE. In terms of weight filtration for EE, we see that the Hilbert-Mumford weight is given by

s(E):=n1(i=1tmi(χ(Ei(N))P(m)P(N)χ(Ei(m))))+μ(ω,E,m)s(E):=n_{1}(\sum_{i=1}^{t}m_{i}(\chi(E_{i}(N))P(m)-P(N)\chi(E_{i}(m))))+\mu(\omega,E_{\textbf{\textbullet}},m_{\textbf{\textbullet}})
+xDj=12lxdj(x)(i=1tmi(χ(Ei(N))rj(x)P(N)rjEi(x)))+\sum_{x\in D}\sum_{j=1}^{2l_{x}}d_{j}(x)(\sum_{i=1}^{t}m_{i}(\chi(E_{i}(N))r_{j}(x)-P(N)r_{j}^{E_{i}}(x)))

where rjEi(x):=dim(Im(EiEQj(x)))r_{j}^{E_{i}}(x):=dim(Im(E_{i}\rightarrow E\rightarrow Q_{j}(x))). Hence the point is GIT semistable if and only if s(E)0s(E)\leq 0.

However, one can show that (see Proposition 2.9 of [23])

s(E)=kP(N)(i=1tmipardeg(Ei))+μ(ω,E,m)s(E)=kP(N)\big{(}\sum_{i=1}^{t}m_{i}pardeg(E_{i})\big{)}+\mu(\omega,E_{\textbf{\textbullet}},m_{\textbf{\textbullet}})

In fact, the coefficients of mim_{i} in s(E)s(E) is

n1(χ(Ei(N))P(m)P(N)χ(Ei(m)))+xDj=12lxdj(x)(χ(Ei(N))rj(x)P(N)rjEi(x))n_{1}\big{(}\chi(E_{i}(N))P(m)-P(N)\chi(E_{i}(m))\big{)}+\sum_{x\in D}\sum_{j=1}^{2l_{x}}d_{j}(x)\big{(}\chi(E_{i}(N))r_{j}(x)-P(N)r_{j}^{E_{i}}(x)\big{)}
=(rl+rKcN)(deg(Ei)r(Ei)rdeg(E))+xDj=12lxdj(x)(χ(Ei(N))rj(x)P(N)rjEi(x))=KP(N)(deg(Ei)r(Ei)rdeg(E))+P(N)(r(Ei)rxDj=12lxdj(x)rj(x)xDj=12lxdj(x)rjEi(x))=KP(N)(deg(Ei)r(Ei)rdeg(E))+r(Ei)rP(N)(rxDa2lx+1xDj=12lx+1aj(x)nj(x))P(N)(r(Ei)xDa2lx+1xDj=12lx+1aj(x)njEi(x))=KP(N)(pardeg(Ei))\begin{split}&=(rl+rKcN)\big{(}deg(E_{i})-\dfrac{r(E_{i})}{r}deg(E)\big{)}+\sum_{x\in D}\sum_{j=1}^{2l_{x}}d_{j}(x)\big{(}\chi(E_{i}(N))r_{j}(x)-P(N)r_{j}^{E_{i}}(x)\big{)}\\ &=KP(N)\big{(}deg(E_{i})-\dfrac{r(E_{i})}{r}deg(E)\big{)}+P(N)\big{(}\dfrac{r(E_{i})}{r}\sum_{x\in D}\sum_{j=1}^{2l_{x}}d_{j}(x)r_{j}(x)-\sum_{x\in D}\sum_{j=1}^{2l_{x}}d_{j}(x)r_{j}^{E_{i}}(x)\big{)}\\ &=KP(N)\big{(}deg(E_{i})-\dfrac{r(E_{i})}{r}deg(E)\big{)}+\dfrac{r(E_{i})}{r}P(N)\big{(}r\sum_{x\in D}a_{2l_{x}+1}-\sum_{x\in D}\sum_{j=1}^{2l_{x}+1}a_{j}(x)n_{j}(x)\big{)}\\ &\quad-P(N)\big{(}r(E_{i})\sum_{x\in D}a_{2l_{x}+1}-\sum_{x\in D}\sum_{j=1}^{2l_{x}+1}a_{j}(x)n_{j}^{E_{i}}(x)\big{)}\\ &=KP(N)(pardeg(E_{i}))\\ \end{split}

Proposition 3.5.

A parabolic symplectic/orthogonal bundle EE is stable(resp. semistable) if and only if the correspondence point ((q,ϕ),(q1(x),q2(x),,q2lx(x))xD)((q,\phi),(q_{1}(x),q_{2}(x),\dots,q_{2l_{x}}(x))_{x\in D}) of \mathcal{R} is GIT stable(resp. semistable) for the action of SL(V)SL(V).

Proof.

Let EE be a stable(resp. semistable) bundle. For any weighted filtration (E,m)(E_{\textbf{\textbullet}},m_{\textbf{\textbullet}}), we have pardeg(Ei)<0( resp.)pardeg(E_{i})<0\ (\text{ resp.}\ \leq) by Lemma 2.5. Furthermore,by Lemma 3.3, μ(ω,E,m)0\mu(\omega,E_{\textbf{\textbullet}},m_{\textbf{\textbullet}})\leq 0,hence

kP(N)(i=1t(pardeg(Ei))+μ(ω,E.,m.)<0(resp.)kP(N)(\sum_{i=1}^{t}(pardeg(E_{i}))+\mu(\omega,E.,m.)<0\ (\text{resp.}\ \leq)

By Proposition 3.4, this tells that the corresponding point ((q,ϕ),(q1(x),q2(x),,q2lx(x))xD)((q,\phi),(q_{1}(x),q_{2}(x),\dots,q_{2l_{x}}(x))_{x\in D}) of \mathcal{R} is GIT stable(resp. semistable).

Conversely, Let EE be a parabolic orthogonal/symplectic bundle such that the corresponding point ((q,ϕ),(q1(x),q2(x),,q2lx(x))xD)((q,\phi),(q_{1}(x),q_{2}(x),\dots,q_{2l_{x}}(x))_{x\in D}) is GIT stable(resp. GIT semistable). We want to show that E is a stable (resp. semistable). That is, for any isotropic subbundle F of E, we have pardeg(F)<0(resp.)pardeg(F)<0\ (\text{resp.}\ \leq).

Since EE is stable(resp. semistable), the inequality in Proposition 3.4 must hold for all weighted filtrations (E,m)(E_{\textbf{\textbullet}},m_{\textbf{\textbullet}}). In particular, if we take the weighted filtration as: 0FFE0\subset F\subset F^{\perp}\subset E, and weights m1=m2=1m_{1}=m_{2}=1, then the inequality becomes

kP(N)((pardeg(F)+pardeg(F))+μ(ω,E,m)<0( resp.).kP(N)((pardeg(F)+pardeg(F^{\perp}))+\mu(\omega,E_{\textbf{\textbullet}},m_{\textbf{\textbullet}})<0\ (\text{ resp.}\ \leq).

However, in this case we have μ(ω,E.,m.)=0\mu(\omega,E.,m.)=0 and pardeg(F)=pardeg(F)pardeg(F)=pardeg(F^{\perp}), hence we have pardeg(F)<0(resp.).pardeg(F)<0\ (\text{resp.}\ \leq).

Therefore,let ss\mathcal{R}^{ss}\subset\mathcal{R} be the open set of \mathcal{R} which consists of semistable parabolic orthogonal(symplectic,resp) sheaves. In the rest of this section, we will show that ss\mathcal{R}^{ss} is smooth. Therefore let MG,P=ss//SL(V)M_{G,P}=\mathcal{R}^{ss}//SL(V) be the GIT quotient, then we have

Theorem 3.6.

MG,PM_{G,P} is the coarse moduli space of semistable parabolic orthogonal/symplectic sheaves of rank rr and degree dd with fixed parabolic type σ\sigma. Moreover, MG,PM_{G,P} is a normal Cohen-Macaulay projective variety, with only rational singularities.

Proof.

Since we have show that ss\mathcal{R}^{ss} is smooth in the next subsection, especially ss\mathcal{R}^{ss} is normal with only rational singularities, so is its GIT quotient MG,PM_{G,P}. Finally the fact that ss\mathcal{R}^{ss} is regular implies that MG,PM_{G,P} is Cohen-Macaulay(see [16]). ∎

3.2. Smoothness of ss\mathcal{R}^{ss}

The smoothness of ss\mathcal{R}^{ss} has essentially proved in [18]. We will reformulate the proof here.

Let QFQ_{F} be the open subscheme of QQ consisting of quotients [q:V𝒪X(N)E]Q[q:V\otimes\mathcal{O}_{X}(-N)\twoheadrightarrow E]\in Q such that H1(E(N))=0H^{1}(E(N))=0. Let ZFZ_{F} be the inverse image of QFQ_{F} under the projection ZQZ\rightarrow Q and RFR_{F} be the inverse image of ZFZ_{F} under the projection RZR\rightarrow Z. If we can show that ZFZ_{F} is smooth, then F\mathcal{R}_{F} is smooth because is a flag bundle over ZFZ_{F}. Thus ss\mathcal{R}^{ss} is smooth as it is an open subscheme of ZFZ_{F}. So the smoothness of ss\mathcal{R}^{ss} reduce to the smoothness of ZFZ_{F}. We will prove the smoothness of ZFZ_{F} in the rest part of this subsection. First of all, let us recall the definition of Atiyah bundle of a principal GG-bundle ([5])

Definition 3.7.

Let p:EXp:E\rightarrow X be a principal GG-bundle, the Atiyah bundle At(E)At(E) of EE is defined as At(E)(U):=H0(p1U,Tp1U)GAt(E)(U):=H^{0}(p^{-1}U,T_{p^{-1}U})^{G} for any open subset UXU\subseteq X.

Proposition 3.8.

Let p:EXp:E\rightarrow X be a principal GG-bundle and At(E)At(E) is the Atiyah bundle, then

  1. (1)

    We have an exact sequence (the Atiyah sequence): 0ad(E)At(E)TX00\rightarrow ad(E)\rightarrow At(E)\rightarrow T_{X}\rightarrow 0.

  2. (2)

    There is a natural isomorphism μ:pAt(E)TE\mu:p^{*}At(E)\stackrel{{\scriptstyle\cong}}{{\longrightarrow}}T_{E}.

Proof.

See [5] section 1. ∎

Remark 3.9.

For the Grassmann variety Gn,rG_{n,r}, let p:𝒜Gn,rp:\mathcal{A}\rightarrow G_{n,r} be the universal GL(r)GL(r) bundle and 0𝒦V𝒪Gn,r𝒜00\rightarrow\mathcal{K}\rightarrow V\otimes\mathcal{O}_{G_{n,r}}\rightarrow\mathcal{A}\rightarrow 0 be the universal exact sequence, then we have an isomorphism At(𝒜)V𝒜At(\mathcal{A})\cong V\otimes\mathcal{A} and the Atiyah sequence becomes 0𝒜𝒜V𝒜𝒦𝒜00\rightarrow\mathcal{A}^{*}\otimes\mathcal{A}\rightarrow V\otimes\mathcal{A}\rightarrow\mathcal{K}^{*}\otimes\mathcal{A}\rightarrow 0.

Let GGL(r)G\hookrightarrow GL(r) be the orthogonal/symplectic subgroup of GL(r)GL(r). In [18], the author has construct the moduli space of principal GG-bundles. The author also shows that QFQ_{F} and ZFZ_{F} can be the open subschemes of some Hilbert scheme:

Consider the Grassmann variety Gn,rG_{n,r}, where n=dimVn=dimV. Denote the universal family over Gn,rG_{n,r} by 𝒜Gn,r\mathcal{A}\rightarrow G_{n,r}. Let Y=GL(r)//GY=GL(r)//G and 𝒜(Y)=𝒜//G\mathcal{A}(Y)=\mathcal{A}//G be the fibre bundle with fibre YY over Gn,rG_{n,r} associated to 𝒜Gn,r\mathcal{A}\rightarrow G_{n,r}. Then we have the following proposition:

Proposition 3.10.

QFQ_{F} is an open subscheme of Hom(C,Gn,r)Hom(C,G_{n,r}) and ZFZ_{F} is an open subscheme of Hom(C,𝒜(Y))Hom\big{(}C,\mathcal{A}(Y)\big{)}.

Proof.

See [18] section 4.13. ∎

Proposition 3.11.

The semistable locus ss\mathcal{R}^{ss} is smooth.

Proof.

As mentioned before, we just need to show that ZFZ_{F} is smooth. By proposition 3.10, ZFHom(C,𝒜(Y))Z_{F}\subset Hom\big{(}C,\mathcal{A}(Y)\big{)}. So let f:C𝒜(Y)f:C\rightarrow\mathcal{A}(Y) be a point of ZFZ_{F}, We need to show that ZFZ_{F} is smooth at ff, that is Hom(C,𝒜(Y))Hom\big{(}C,\mathcal{A}(Y)\big{)} is smooth at ff. However by associating ff to the graph Γf\Gamma_{f}, we may consider Hom(C,𝒜(Y))Hom\big{(}C,\mathcal{A}(Y)\big{)} as an open subscheme of Hilb(C×𝒜(Y))Hilb\big{(}C\times\mathcal{A}(Y)\big{)}. Hence we should prove the that Hilb(C×𝒜(Y))Hilb\big{(}C\times\mathcal{A}(Y)\big{)} is smooth at Γf\Gamma_{f}. By obstruction theory, this is equivalent to show that H1(Γf,NΓf)=0H^{1}(\Gamma_{f},N_{\Gamma_{f}})=0 where NΓfN_{\Gamma_{f}} is the normal bundle of Γf\Gamma_{f} in C×𝒜(Y)C\times\mathcal{A}(Y). However, since ΓfC\Gamma_{f}\simeq C by the projection C×𝒜(Y)CC\times\mathcal{A}(Y)\rightarrow C, we have NΓffT𝒜(Y)N_{\Gamma_{f}}\simeq f^{*}T_{\mathcal{A}(Y)} where T𝒜(Y)T_{\mathcal{A}(Y)} is the tangent bundle of 𝒜(Y)\mathcal{A}(Y). We will show that H1(C,fT𝒜(Y))=0H^{1}(C,f^{*}T_{\mathcal{A}(Y)})=0.

Let p:𝒜(Y)Gn,rp:\mathcal{A}(Y)\rightarrow G_{n,r} and q:𝒜Gn,rq:\mathcal{A}\rightarrow G_{n,r} be the caonical maps. Dnote by θ:𝒜𝒜(Y)=𝒜//G\theta:\mathcal{A}\rightarrow\mathcal{A}(Y)=\mathcal{A}//G the natural quotient. It is easy to see q=pθq=p\circ\theta. Then we get the following exact sequence by taking the differential of the projection θ:𝒜𝒜(Y)\theta:\mathcal{A}\rightarrow\mathcal{A}(Y)

0θT𝒜T𝒜(Y)0\displaystyle 0\longrightarrow\mathcal{M}\longrightarrow\theta_{*}T_{\mathcal{A}}\longrightarrow T_{\mathcal{A}(Y)}\longrightarrow 0

On the other hand, by proposition 3.8, we have

T𝒜qAt(𝒜)=(pθ)At(𝒜)=θpAt(𝒜)\displaystyle T_{\mathcal{A}}\cong q^{*}At(\mathcal{A})=(p\circ\theta)^{*}At(\mathcal{A})=\theta^{*}p^{*}At(\mathcal{A})

So we have a surjective map

θθpAt(𝒜)T𝒜(Y)\displaystyle\theta_{*}\theta^{*}p^{*}At(\mathcal{A})\longrightarrow T_{\mathcal{A}(Y)}

Compose the above map with the canonical map pAt(𝒜)θθpAt(𝒜)p^{*}At(\mathcal{A})\longrightarrow\theta_{*}\theta^{*}p^{*}At(\mathcal{A}), we get a surjective morphism

pAt(𝒜)T𝒜(Y)\displaystyle p^{*}At(\mathcal{A})\longrightarrow T_{\mathcal{A}(Y)}

Assume the kernel is 𝒩\mathcal{N}, then we have an exact sequence

0𝒩pAt(𝒜)T𝒜(Y)0\displaystyle 0\longrightarrow\mathcal{N}\longrightarrow p^{*}At(\mathcal{A})\longrightarrow T_{\mathcal{A}(Y)}\longrightarrow 0

Taking the pullback functor ff^{*}, we get an exact sequence over CC

0f𝒩fpAt(𝒜)fT𝒜(Y)0\displaystyle 0\longrightarrow f^{*}\mathcal{N}\longrightarrow f^{*}p^{*}At(\mathcal{A})\longrightarrow f^{*}T_{\mathcal{A}(Y)}\longrightarrow 0

We get a exact sequence of cohomologies from the above sequence

H1(C,f𝒩)H1(C,fpAt(𝒜))H1(C,fT𝒜(Y))0\displaystyle H^{1}(C,f^{*}\mathcal{N})\longrightarrow H^{1}(C,f^{*}p^{*}At(\mathcal{A}))\longrightarrow H^{1}(C,f^{*}T_{\mathcal{A}(Y)})\longrightarrow 0 (3.1)

However, by Remark 3.9, we have At(𝒜)V𝒜At(\mathcal{A})\cong V\otimes\mathcal{A}. So

fpAt(𝒜)(pf)(V𝒜)f^{*}p^{*}At(\mathcal{A})\cong(p\circ f)^{*}(V\otimes\mathcal{A})

Notice that pfZFp\circ f\in Z_{F}, so it correspondence to a quotient bundle

0FfV𝒪C(N)Ef0\displaystyle 0\longrightarrow F_{f}\longrightarrow V\otimes\mathcal{O}_{C}(-N)\longrightarrow E_{f}\longrightarrow 0

Then we have (pf)(V𝒜)VEf(N)(p\circ f)^{*}(V\otimes\mathcal{A})\cong V\otimes E_{f}(N), so we have that

H1(C,fpAt(𝒜))H1(C,VEf(N))=0\displaystyle H^{1}\big{(}C,f^{*}p^{*}At(\mathcal{A})\big{)}\cong H^{1}\big{(}C,V\otimes E_{f}(N)\big{)}=0 (3.2)

Combine with (3.1) and (3.2) we finally get H1(C,fT𝒜(Y))=0H^{1}(C,f^{*}T_{\mathcal{A}(Y)})=0.

4. Codimention estimate

In this section, we fix SS to be a scheme of finite type. Let \mathcal{E} be a flat family of vector bundle, principal GG bundle, parabolic vector bundle or parabolic symplectic/orthogonal bundle over CC parametrized by SS, under certain conditions, we want to estimate the codimension of the unstable (unsemistable) locus, i.e. the locally closed subscheme SusSS^{us}\subset S (SussSS^{uss}\subset S) parametrizing all t\mathcal{E}_{t} which is not stable (semistable). Our main method is taken from [10].

4.1. The case of vector bundle and principal GG bundle

In fact, the case of vector bundle and principal GG bundle have been already done in [10] and [11]. For later use, we reformulate the results and give a short proof if necessary.

All the stories begin with the following proposition:

Proposition 4.1.

\mathcal{E} is a flat family of vector bundles over S×CS\times C. Let ϕ:QS\phi:Q\rightarrow S be the relative Quot-scheme parametrizing all flat quotients of \mathcal{E} with certain fixed rank and degree. For any sSs\in S and qϕ1(s)q\in\phi^{-1}(s), corresponding to exact sequence:

0FsG00\longrightarrow F\longrightarrow\mathcal{E}_{s}\longrightarrow G\longrightarrow 0

we have the following exact sequence:

0Hom(F,G)TqQTsSExt1(F,G).\displaystyle 0\longrightarrow\text{Hom}(F,G)\longrightarrow T_{q}Q\longrightarrow T_{s}S\longrightarrow\text{Ext}^{1}(F,G). (4.1)
Proof.

See [9] Proposition 2.2.7. ∎

Let EE be a vector bundle over CC, the classical Harder-Narasimhan filtration and Jordan-Holder filtration show that if EE is not stable(resp. semistable), then there is a maximal stable subbundle F0EF_{0}\subset E with the property degom(F0,E/F0)0\text{deg}\mathcal{H}om(F_{0},E/F_{0})\leq 0 (resp. <0<0). F0F_{0} is taken to be the first term of the Jordan-Holder filtration of the maximal destabilizing subbundle of EE(so different choice of F0F_{0} have same slope). Moreover, if we say F0F_{0} is of type μ=(r,d)\mu=(r^{\prime},d^{\prime}), i.e. FF is of rank rr^{\prime} and degree dd^{\prime}, Then for a flat family of vector bundle \mathcal{E} over S×CS\times C, the locus SμSS^{\mu}\subset S parametrizing t\mathcal{E}_{t} having a subbundle described above with type μ\mu, is locally closed and non-empty for finitely many μ\mu.

Similarly properties hold for principal GG bundles. Let EE be a principal GG bundle, then there is a unique standard parabolic subgroup PP and a unique reduction EPE_{P}, and if we denote E𝗌E_{\mathsf{s}} to be the vector bundle associated to EPE_{P} by the natural representation of PP on the vector space 𝗌:=𝗀/𝗉\mathsf{s}:=\mathsf{g}/\mathsf{p}, where 𝗀\mathsf{g} and 𝗉\mathsf{p} are Lie algebras of GG and PP, then degE𝗌<0\text{deg}E_{\mathsf{s}}<0. More over, we have similar concept of SμS^{\mu}. For details and proof, please refer to [11].

Proposition 4.2.

Let \mathcal{E} be a flat family of vector bundles or principal GG bundles over S×CS\times C. Assume that for each closed point tSt\in S, the Kodaira-Spencer maps

TtSExt1(t,t) or TtSH1(C,t(Ad))T_{t}S\rightarrow\text{Ext}^{1}(\mathcal{E}_{t},\mathcal{E}_{t})\text{\ \ \ \ or\ \ \ \ }T_{t}S\rightarrow H^{1}(C,\mathcal{E}_{t}(Ad))

are surjective. Then:

  1. (1)

    In the vector bundle case, for any sSμs\in S^{\mu}, the normal space NsSμN_{s}S^{\mu} is isomorphic to Ext1(F0,s/F0)\text{Ext}^{1}(F_{0},\mathcal{E}_{s}/F_{0}), where F0F_{0} is a maximal stable bundle described above.

  2. (2)

    In the principal GG bundles case, for any sSμs\in S^{\mu}, the normal space NsSμN_{s}S^{\mu} is isomorphic to H1(C,s,𝗌)H^{1}(C,\mathcal{E}_{s,\mathsf{s}}) where s,𝗌\mathcal{E}_{s,\mathsf{s}} is described above.

Proof.

For the vector bundle case, we first consider the Quot-scheme ϕ:QS\phi:Q\rightarrow S parametrizing all subbundles of type μ\mu, then analyse the exact sequence 4.1. Firstly the image of ϕ\phi covers SμS^{\mu}, we see that the map TqQTsST_{q}Q\rightarrow T_{s}S factors as TqQTsSμTsST_{q}Q\twoheadrightarrow T_{s}S^{\mu}\hookrightarrow T_{s}S. Secondly, by the proof of exactness of 4.1, we see that the map TsSExt1(F0,s/F0)T_{s}S\rightarrow\text{Ext}^{1}(F_{0},\mathcal{E}_{s}/F_{0}) indeed factors as

TsSExt1(s,s)Ext1(F0,s/F0).T_{s}S\rightarrow\text{Ext}^{1}(\mathcal{E}_{s},\mathcal{E}_{s})\rightarrow\text{Ext}^{1}(F_{0},\mathcal{E}_{s}/F_{0}).

The first map is Kodaira-Spencer map which is surjective by assumption; the second map is induced by the exact sequence:

0F0ss/F000\longrightarrow F_{0}\longrightarrow\mathcal{E}_{s}\longrightarrow\mathcal{E}_{s}/F_{0}\longrightarrow 0

which is surjective naturally. Thus we see that Ext1(F0,s/F0)\text{Ext}^{1}(F_{0},\mathcal{E}_{s}/F_{0}) is isomorphic to the cokernel of TqQTsST_{q}Q\rightarrow T_{s}S, i.e. the normal space NsSμN_{s}S^{\mu}.

The principal bundle case is similar, except we need a variety to parametrize all reductions to PP. But this is already done in [18], it is an open subscheme 𝒰\mathcal{U} of Hilb(/P)/S\text{Hilb}_{(\mathcal{E}/P)_{/S}}, parametrizing all sections of /PS\mathcal{E}/P\rightarrow S. Now we apply Proposition 4.1 to this 𝒰\mathcal{U}, with similar method above, we have our proposition. ∎

Corollary 4.3.

With same notation and assumptions as above, if we assume SS is smooth, we have:

  1. (1)

    In the vector bundle case, the rank of \mathcal{E} is assumed to be rr, then we have

    codim(Sus)\displaystyle\text{codim}(S^{us}) (r1)(g1)\displaystyle\geq(r-1)(g-1)
    codim(Suss)\displaystyle\text{codim}(S^{uss}) >(r1)(g1)\displaystyle>(r-1)(g-1)
  2. (2)

    In the principal bundle case, we have

    codim(Sus)\displaystyle\text{codim}(S^{us}) rank(t,𝗌)(g1)\displaystyle\geq\text{rank}(\mathcal{E}_{t,\mathsf{s}})(g-1)
    codim(Suss)\displaystyle\text{codim}(S^{uss}) >rank(t,𝗌)(g1)\displaystyle>\text{rank}(\mathcal{E}_{t,\mathsf{s}})(g-1)
Proof.

Since SμS^{\mu} is non-empty for only finitely many μ\mu, by proposition above, we only need to calculate dimExt1(F0,t/F0)\text{dim}\text{Ext}^{1}(F_{0},\mathcal{E}_{t}/F_{0}) and dimH1(C,t,𝗌)\text{dim}H^{1}(C,\mathcal{E}_{t,\mathsf{s}}). Using Riemann-Roch, we have

dimExt1(F0,t/F0)\displaystyle\text{dim}\text{Ext}^{1}(F_{0},\mathcal{E}_{t}/F_{0}) =dimHom(F0,t/F0)degom(F0,t/F0)+r(rr)(g1)\displaystyle=\text{dim}\text{Hom}(F_{0},\mathcal{E}_{t}/F_{0})-\text{deg}\mathcal{H}om(F_{0},\mathcal{E}_{t}/F_{0})+r^{\prime}(r-r^{\prime})(g-1)
dimH1(C,t,𝗌)\displaystyle\text{dim}H^{1}(C,\mathcal{E}_{t,\mathsf{s}}) =dimH0(C,t,𝗌)degt,𝗌+rankt,𝗌(g1)\displaystyle=\text{dim}H^{0}(C,\mathcal{E}_{t,\mathsf{s}})-\text{deg}\mathcal{E}_{t,\mathsf{s}}+\text{rank}\mathcal{E}_{t,\mathsf{s}}(g-1)

where rr^{\prime} is the rank of FF. Thus our corollary holds by analyse of degrees of om(F0,t/F0)\mathcal{H}om(F_{0},\mathcal{E}_{t}/F_{0}) and t,𝗌\mathcal{E}_{t,\mathsf{s}} before. ∎

4.2. The case of parabolic vector bundle

We fix \mathcal{E} to be a flat family of parabolic vector bundles of type σ\sigma over S×CS\times C.To apply our method to parabolic vector bundle case, we need to construct an SS-scheme parametrizing all flat quotients of \mathcal{E}, with fixed parabolic type σ\sigma^{\prime}.

We begin with a functor

𝖥:(Sch/S)op(Set)\mathsf{F}:(Sch/S)^{op}\longrightarrow(Set)

as follows: for any f:TSf:T\rightarrow S, 𝖥(f:TS)\mathsf{F}(f:T\rightarrow S) is the set of isomorphism classes of all quotients fC𝒢0f^{*}_{C}\mathcal{E}\rightarrow\mathcal{G}\rightarrow 0, such that the induced parabolic structure on 𝒢\mathcal{G} makes 𝒢\mathcal{G} a flat family of parabolic vector bundle of rank rr^{\prime} and degree dd^{\prime} with fixed type σ\sigma^{\prime}.

Proposition 4.4.

𝖥\mathsf{F} is represented by a finite type scheme ϕP:QPS\phi_{P}:Q_{P}\rightarrow S.

Proof.

Thanks to Proposition 2.7, we will translate parabolic bundle and orbifold bundle interchangeably.

\mathcal{E} gives a flat family of orbifold bundle 𝒲\mathcal{W} over S×YS\times Y. Firstly we consider the Quot-scheme QSQ\rightarrow S, parametrizing all flat quotients of 𝒲\mathcal{W} with certain fixed rank and degree. Secondly, since 𝒲\mathcal{W} is an orbifold bundle, we see that Γ\Gamma acts on QQ, and the closed subscheme QΓQ^{\Gamma} of Γ\Gamma-invariant points parametrizes all the orbifold quotients of 𝒲\mathcal{W}([20]). At last, by[20] again, there is an open subscheme QPQΓQ_{P}\subset Q^{\Gamma}, parametrizing all locally free orbifold quotients with fixed type σ\sigma^{\prime}. We claim that QPQ_{P} represents 𝖥\mathsf{F}.

For any f:TSf:T\rightarrow S, and any quotient fC𝒢0f^{*}_{C}\mathcal{E}\rightarrow\mathcal{G}\rightarrow 0, using the correspondence in Proposition 2.7, we see easily that there is an SS-morphism: TQPT\rightarrow Q_{P}. Conversely, Given an SS-morphism φ:TQP\varphi:T\rightarrow Q_{P}, this would give a flat orbifold bundle quotient fY𝒲𝒢~0f_{Y}^{*}\mathcal{W}\rightarrow\tilde{\mathcal{G}}\rightarrow 0. By our correspondence, we have a quotient

fC𝒢0f^{*}_{C}\mathcal{E}\rightarrow\mathcal{G}\rightarrow 0

where 𝒢\mathcal{G} is a flat family of parabolic vector bundles with type σ\sigma^{\prime}. Notice that this is a quotient since taking Γ\Gamma invariant sections of \mathbb{C}-modules is an exact functor. ∎

Remark 4.5.

In [7], a similar scheme is constructed in a different way.

Corollary 4.6.

For any sSs\in S and qϕP1(s)q\in\phi_{P}^{-1}(s), corresponding to exact sequence:

0FsG00\longrightarrow F\longrightarrow\mathcal{E}_{s}\longrightarrow G\longrightarrow 0

Then we have an exact sequence:

0Hompar(F,G)TqQPTsSExtpar1(F,G).0\longrightarrow\text{Hom}_{par}(F,G)\longrightarrow T_{q}Q_{P}\longrightarrow T_{s}S\longrightarrow\text{Ext}^{1}_{par}(F,G).
Proof.

Let 0F~𝒲sG~00\rightarrow\tilde{F}\rightarrow\mathcal{W}_{s}\rightarrow\tilde{G}\rightarrow 0 be the corresponding exact sequence of orbifold bundles over YY. When we regard qq as a point of QQ, apply the exact sequence 4.1, we have an exact sequence:

0Hom(F~,G~)TqQTsSExt1(F~,G~)0\longrightarrow\text{Hom}(\tilde{F},\tilde{G})\longrightarrow T_{q}Q\longrightarrow T_{s}S\longrightarrow\text{Ext}^{1}(\tilde{F},\tilde{G})

However, this sequence is in fact a Γ\Gamma-exact sequence, Thus we have:

0Hom(F~,G~)Γ(TqQ)ΓTsSExt1(F~,G~)Γ0\longrightarrow\text{Hom}(\tilde{F},\tilde{G})^{\Gamma}\longrightarrow(T_{q}Q)^{\Gamma}\longrightarrow T_{s}S\longrightarrow\text{Ext}^{1}(\tilde{F},\tilde{G})^{\Gamma}

which is exact since taking Γ\Gamma-invariant sections of \mathbb{C}-modules is an exact functor. Now, it is known that Hom(F~,G~)Γ=Hompar(F,G)\text{Hom}(\tilde{F},\tilde{G})^{\Gamma}=\text{Hom}_{par}(F,G) and (TqQ)Γ=TqQP(T_{q}Q)^{\Gamma}=T_{q}Q_{P}. Finally, spectral sequence argument tells Ext1(F~,G~)Γ=Extpar1(F,G)\text{Ext}^{1}(\tilde{F},\tilde{G})^{\Gamma}=\text{Ext}^{1}_{par}(F,G), we are done. ∎

Before going further, we mention that there are Harder-Narasimhan filtration and Jordan-Holder filtration for parabolic bundles. So similar as in the previous subsection, for a parabolic bundle which is not stable (resp. semistable), there is a maximal stable subbundle F0F_{0} such that pardegompar(F0,E/F0)0pardeg\mathcal{H}om_{par}(F_{0},E/F_{0})\leq 0 (resp. <0<0). Moreover, for a family of parabolic vector bundle as above, SμS^{\mu} defined as before, is locally closed and non-empty for finitely many μ\mu.

Proposition 4.7.

Assume that for any tSt\in S, the Kodaira-Spencer map

TtSExtpar1(t,t)T_{t}S\longrightarrow\text{Ext}^{1}_{par}(\mathcal{E}_{t},\mathcal{E}_{t})

is surjective. Let SμSS^{\mu}\subset S be the locally closed described before. Then for any sSμs\in S^{\mu}, we have NsSμExtpar1(F0,s/F0)N_{s}S^{\mu}\cong\text{Ext}^{1}_{par}(F_{0},\mathcal{E}_{s}/F_{0}).

Proof.

Similar as Proposition 4.2. ∎

Corollary 4.8.

With same assumption as above, assuming that SS is smooth and rank=r\text{rank}\ \mathcal{E}=r we have

codim(Sus)\displaystyle\text{codim}(S^{us}) degD/K+(r1)(g1)\displaystyle\geq\text{deg}D/K+(r-1)(g-1)
codim(Suss)\displaystyle\text{codim}(S^{uss}) >degD/K+(r1)(g1).\displaystyle>\text{deg}D/K+(r-1)(g-1).
Proof.

As before, it suffice to estimate dimExtpar1(F0,s/F0)\text{dim}\text{Ext}^{1}_{par}(F_{0},\mathcal{E}_{s}/F_{0}). By [26], we have Extpar1(F0,s/F0)=H1(C,ompar(F0,s/F0))\text{Ext}^{1}_{par}(F_{0},\mathcal{E}_{s}/F_{0})=H^{1}(C,\mathcal{H}om_{par}(F_{0},\mathcal{E}_{s}/F_{0})), so

dimExtpar1(F0,s/F0)=dimHompar(F0,s/F0)degompar(F0,s/F0)+r(rr)(g1)\text{dim}\text{Ext}^{1}_{par}(F_{0},\mathcal{E}_{s}/F_{0})=\text{dim}\text{Hom}_{par}(F_{0},\mathcal{E}_{s}/F_{0})-\text{deg}\mathcal{H}om_{par}(F_{0},\mathcal{E}_{s}/F_{0})+r^{\prime}(r-r^{\prime})(g-1)

Since pardegompar(F0,s/F0)0\text{pardeg}\mathcal{H}om_{par}(F_{0},\mathcal{E}_{s}/F_{0})\leq 0. We see that degompar(F0,s/F0)degD/K-\text{deg}\mathcal{H}om_{par}(F_{0},\mathcal{E}_{s}/F_{0})\geq\text{deg}D/K. This would give our results. ∎

Remark 4.9.

Similar results have been given in [22] by a different way.

4.3. The case of parabolic symplectic/orthogonal bundle

The case of parabolic symplectic/orthogonal bundles is similar to those in former two sections, but we need define some notions first.

Let EE be a parabolic symplectic bundle over CC, and WW be the corresponding orbifold symplectic bundle over YY. By the constructions before, we have W(Ad)W(Ad) and W𝗌W_{\mathsf{s}} for 𝗌=𝗀/𝗉\mathsf{s}=\mathsf{g}/\mathsf{p}. WW is an orbifold symplectic bundle, so W(Ad)W(Ad) and W𝗌W_{\mathsf{s}} are both orbifold vector bundles over YY. We use E(Ad)E(Ad) and E𝗌E_{\mathsf{s}} to denote corresponding parabolic vector bundles over CC.

For any family of parabolic symplectic bundle \mathcal{E} over CC parametrized by a scheme SS, let 𝒲\mathcal{W} be the corresponding orbifold symplectic bundle on S×YS\times Y. For any tSt\in S, we have the Kodaira-Spencer map

TtSH1(Y,𝒲t(Ad))T_{t}S\longrightarrow\text{H}^{1}(Y,\mathcal{W}_{t}(Ad))

for 𝒲\mathcal{W}. This map is obviously Γ\Gamma-invariant, so we have

TtSH1(Y,𝒲t(Ad))Γ=H1(C,t(Ad))T_{t}S\longrightarrow\text{H}^{1}(Y,\mathcal{W}_{t}(Ad))^{\Gamma}=\text{H}^{1}(C,\mathcal{E}_{t}(Ad))
Definition 4.10.

The Kodaira-Spencer map for \mathcal{E} at tSt\in S is given by

TtSH1(C,t(Ad)).T_{t}S\longrightarrow\text{H}^{1}(C,\mathcal{E}_{t}(Ad)).
Proposition 4.11.

Let SS and \mathcal{E} be as before. Then there is a scheme ϕPS:QPSS\phi_{PS}:Q_{PS}\rightarrow S parametrizing all isotropic subbundles of \mathcal{E}, flat over SS with same fixed type τ\tau^{\prime}.

Moreover, for any sSs\in S and qϕPS1(s)q\in\phi_{PS}^{-1}(s), corresponding to an isotropic subbundle FsF\subset\mathcal{E}_{s}, which corresponds to a reduction to a parabolic subgroup PP of 𝒲s\mathcal{W}_{s}, we have an exact sequence:

0H0(C,s,𝗌)TqQPSTsSH1(C,s,𝗌)0\longrightarrow\text{H}^{0}(C,\mathcal{E}_{s,\mathsf{s}})\longrightarrow T_{q}Q_{PS}\longrightarrow T_{s}S\longrightarrow\text{H}^{1}(C,\mathcal{E}_{s,\mathsf{s}})
Proof.

Similar to Corollary 4.6. ∎

With similar method, we can show that:

Corollary 4.12.

With notations as before, assume that the Kodaira-Spencer map is surjective for any sSs\in S, then we have

codim(Sus)\displaystyle\text{codim}(S^{us}) degD/K+rank(s,𝗌)(g1)\displaystyle\geq\text{deg}D/K+\text{rank}(\mathcal{E}_{s,\mathsf{s}})(g-1)
codim(Suss)\displaystyle\text{codim}(S^{uss}) >degD/K+rank(s,𝗌)(g1).\displaystyle>\text{deg}D/K+\text{rank}(\mathcal{E}_{s,\mathsf{s}})(g-1).

5. Infinite Grassmannians and the theta line bundle

5.1. Infinite Grassmannians

In this subsection, we use GG to denote a connected simply connected simple affine algebraic group, and the parabolic GG bundle over CC we considered in this subsection is given by a principal GG bundle EE together with choices of one parameter subgroups in E(G)xE(G)_{x} for every xDx\in D; a quasi-parabolic GG bundle is just a choice of choices of parabolic subgroups of E(G)xE(G)_{x}, i.e. (quasi-)parabolic GG bundles in the sense of [3].

We fix a point pXp\in X, away from DD, let C=CpC^{*}=C-p, following [12], we define

𝒢\displaystyle\mathcal{G} =G(^p)\displaystyle=G(\hat{\mathbb{C}}_{p})
𝒫\displaystyle\mathcal{P} =G(𝒪^p)\displaystyle=G(\hat{\mathcal{O}}_{p})
Λ\displaystyle\Lambda =G([C])\displaystyle=G(\mathbb{C}[C^{*}])

where 𝒪^p\hat{\mathcal{O}}_{p} is the completion of local ring 𝒪p\mathcal{O}_{p} of pCp\in C; ^p\hat{\mathbb{C}}_{p} is the field of quotient of 𝒪^p\hat{\mathcal{O}}_{p}; [C]\mathbb{C}[C^{*}] is the coordinate ring of CC^{*}. Similarly in [12], we have

Proposition 5.1.

If we use 𝒳\mathcal{X} to denote the set of isomorphism classes of quasi-parabolic GG bundle with parabolic structure PxP_{x} at each xDx\in D, we have a bijection of sets:

α:Λ(𝒢/𝒫×xDG/Px)𝒳\alpha:\Lambda\setminus(\mathcal{G}/\mathcal{P}\times\prod_{x\in D}G/P_{x})\longrightarrow\mathcal{X}
Proof.

By proposition 1.5 of [12], there is a bijection between Λ𝒢/𝒫\Lambda\setminus\mathcal{G}/\mathcal{P} and the set of isomorphism classes of GG bundles. Notice that a quasi-parabolic GG bundle is nothing but a parabolic GG bundle plus a point in xDG/Px\prod_{x\in D}G/P_{x}, we have our bijection. ∎

Recall that in [12] the generalized flag variety X:=𝒢/𝒫X:=\mathcal{G}/\mathcal{P} has a structure of ind-variety, more precisely,

X=limXσX=\lim_{\rightarrow}X_{\sigma}

where XσX_{\sigma} are the generalised Schubert varieties they defined there. Moreover, there is an algebraic GG bundle 𝒰C×X\mathcal{U}\rightarrow C\times X such that 𝒰|C×X\mathcal{U}|_{C^{*}\times X} is trivial. So, for any xDx\in D, we have a trivial GG bundle 𝒰x\mathcal{U}_{x} over XX, then we define

XP=X×xDG/PxX_{P}=X\times\prod_{x\in D}G/P_{x}

to be the relative flag variety over XX defined by {𝒰x}xD\{\mathcal{U}_{x}\}_{x\in D}. Let π:XPX\pi:X_{P}\longrightarrow X be the natural projection.

Proposition 5.2.

There is a quasi-parabolic GG bundle 𝒰P\mathcal{U}_{P} over C×XPC\times X_{P} such that for any xXPx\in X_{P}, the quasi-parabolic bundle (𝒰P)x:=𝒰P|C×x(\mathcal{U}_{P})_{x}:=\mathcal{U}_{P}|_{C\times x} is exactly the parabolic GG bundle corresponds to xx though the bijection in Proposition 5.1. Moreover the bundle 𝒰|C×XP\mathcal{U}|_{C^{*}\times X_{P}} carries a trivialization ϵ:τ𝒰|C×XP\epsilon:\tau\rightarrow\mathcal{U}|_{C^{*}\times X_{P}} where τ\tau is a trivial quasi-parabolic GG bundle over C×XPC^{*}\times X_{P}.

For any scheme TT and any family of parabolic GG bundle \mathcal{F} over C×TC\times T, if |C×T\mathcal{F}|_{C^{*}\times T} and |Spec𝒪^p×T\mathcal{F}|_{Spec\hat{\mathcal{O}}_{p}\times T} are both trivial. Then if we choose a trivialization ε:τ|C×T\varepsilon:\tau^{\prime}\rightarrow\mathcal{F}|_{C^{*}\times T}, we would have a Schubert variety XσX_{\sigma}, and a morphism f:TXσ×xDG/Pxf:T\rightarrow X_{\sigma}\times\prod_{x\in D}G/P_{x} such that ε\varepsilon is exactly the trivialization pulled back from ϵ\epsilon by ff.

Proof.

This is just a parabolic analogy of proposition 2.8 in [12]. The quasi-parabolic GG bundle 𝒰P\mathcal{U}_{P} is given by π𝒰\pi^{*}\mathcal{U} with quasi-parabolic structure determined by universal property of flag variety.

To see the existence of the morphism ff, we firstly observe that by proposition 2.8 in [12], we have a morphism f:TXf^{\prime}:T\longrightarrow X. Now since |C×T\mathcal{F}|_{C^{*}\times T}is trivial, we would have a point in xDG/Px\prod_{x\in D}G/P_{x} determined by this trivial parabolic GG bundle. ∎

Corollary 5.3.

There is an open subset XPssXPX_{P}^{ss}\subset X_{P} and a morphism ϕ:XPssMG,P\phi:X_{P}^{ss}\longrightarrow M_{G,P} to the moduli space of semistable GG bundles.

By proposition 5.1, for any point mMG,Pm\in M_{G,P}, the fibre ϕ1(m)\phi^{-1}(m) is a union of certain Λ\Lambda-orbits. Next, we analyse the closure of these orbits.

Lemma 5.4.

Let EE be a semistable parabolic GG bundle on CC and we consider gr(E)gr(E) defined in Proposition 3.1 of [3]. Then there exists a family of parabolic GG bundle \mathcal{E} on C×𝔸1C\times\mathbb{A}^{1} such that:

  1. (a)

    |C×(𝔸1{0})pC(E)\mathcal{E}|_{C\times(\mathbb{A}^{1}\setminus\{0\})}\cong p_{C}^{*}(E) , |C×{0}gr(E)\mathcal{E}|_{C\times\{0\}}\cong gr(E) and

  2. (b)

    |C×𝔸1\mathcal{E}|_{C^{*}\times\mathbb{A}^{1}} and |Spec𝒪^p×𝔸1\mathcal{E}|_{Spec\hat{\mathcal{O}}_{p}\times\mathbb{A}^{1}} are both trivial.

Where pCp_{C} is the projection from C×𝔸1C\times\mathbb{A}^{1} to CC.

Proof.

This is the parabolic analogy of Proposition 3.7 of [12]. Proof is similar and we omit the it here. ∎

Now we have the following:

Proposition 5.5.

The morphism ϕ:Pic(MG,P)Pic(XPss)\phi^{*}:\text{Pic}(M_{G,P})\longrightarrow\text{Pic}(X_{P}^{ss}) is injective.

Proof.

By lemma 5.4 we know that the fibre ϕ1(m)\phi^{-1}(m) for any mMG,Pm\in M_{G,P} is a disjoint of Λ\Lambda orbits, and the closure of these orbits intersect with each other. Thus by a similar argument in the proof of lemma 2.1 in [11], we have our injection. ∎

5.2. The theta line bundle and the canonical line bundle of MG,PM_{G,P}

In this subsection, we fix

l:=1r(KχxDi=12lxdi(x)ri(x))l:=\dfrac{1}{r}(K\chi-\sum_{x\in D}\sum_{i=1}^{2l_{x}}d_{i}(x)r_{i}(x))

to be an integer, and Dl:=qlqzqD_{l}:=\sum_{q}l_{q}z_{q} to be an effective divisor of degree ll on CC.

Given a scheme SS and a flat family of parabolic principal GG bundle \mathcal{F} over S×CS\times C with parabolic type ({n(x)}xD,{a(x)}xD)\big{(}\{\overrightarrow{n}(x)\}_{x\in D},\{\overrightarrow{a}(x)\}_{x\in D}\big{)}, assuming that for each xDx\in D, the filtration is given by

0=F2lx+1(S×{x})Flx+1(S×{x})Flx(S×{x})F0(S×{x})=S×{x}0=F_{2l_{x}+1}(\mathcal{F}_{S\times\{x\}})\subseteq\cdots F_{l_{x}+1}(\mathcal{F}_{S\times\{x\}})\subseteq F_{l_{x}}(\mathcal{F}_{S\times\{x\}})\subseteq\cdots\subseteq F_{0}(\mathcal{F}_{S\times\{x\}})=\mathcal{F}_{S\times\{x\}}

which is equivalent to

S×{x}=Q2lx+1(S×{x})Qlx+1(S×{x})Qlx(S×{x})Q0(S×{x})=0\mathcal{F}_{S\times\{x\}}=Q_{2l_{x}+1}(\mathcal{F}_{S\times\{x\}})\twoheadrightarrow\cdots\twoheadrightarrow Q_{l_{x}+1}(\mathcal{F}_{S\times\{x\}})\twoheadrightarrow Q_{l_{x}}(\mathcal{F}_{S\times\{x\}})\twoheadrightarrow\cdots\twoheadrightarrow Q_{0}(\mathcal{F}_{S\times\{x\}})=0

then we can define a line bundle Θ,Dl\Theta_{\mathcal{F},D_{l}} on SS by

Θ,Dl:=(detRπS)KxD{i=12lxdet(Qi(S×{x}))di(x)}qdet(S×{zq})lq\Theta_{\mathcal{F},D_{l}}:=(\text{det}R\pi_{S}\mathcal{F})^{-K}\otimes\bigotimes_{x\in D}\big{\{}\bigotimes_{i=1}^{2l_{x}}\text{det}\big{(}Q_{i}(\mathcal{F}_{S\times\{x\}})\big{)}^{d_{i}(x)}\big{\}}\otimes\bigotimes_{q}\text{det}(\mathcal{F}_{S\times\{z_{q}\}})^{l_{q}}

where πS:S×CS\pi_{S}:S\times C\rightarrow S is the projection and detRπS\text{det}R_{\pi_{S}}\mathcal{F} is the determinant of cohomology: {detRπS}t=detH0(C,t)detH1(C,t)1\{\text{det}R_{\pi_{S}}\mathcal{F}\}_{t}=\text{det}\text{H}^{0}(C,\mathcal{F}_{t})\otimes\text{det}\text{H}^{1}(C,\mathcal{F}_{t})^{-1}. Notice that

det(Qi(S×{x}))det(Q2lx+1i(S×{x}))\text{det}\big{(}Q_{i}(\mathcal{F}_{S\times\{x\}})\big{)}\cong\text{det}\big{(}Q_{2l_{x}+1-i}(\mathcal{F}_{S\times\{x\}})\big{)}

for 1ilx1\leq i\leq l_{x}.

It is clear that for any morphism f:TSf:T\rightarrow S, we have fΘ,Dl=ΘfC,Dlf^{*}\Theta_{\mathcal{F},D_{l}}=\Theta_{f_{C}^{*}\mathcal{F},D_{l}}, where fC:T×CS×Cf_{C}:T\times C\rightarrow S\times C is the base change of ff. Moreover, we have:

Theorem 5.6.

There is a unique ample line bundle ΘDl\Theta_{D_{l}} over the moduli space MG,PM_{G,P}, such that:

  1. (1)

    For any scheme SS and any family of semistable parabolic GG bundle \mathcal{F} over S×CS\times C, let ϕ:SMG,P\phi_{\mathcal{F}}:S\rightarrow M_{G,P} be the induced map, then we have

    ϕΘDl=Θ,Dl.\phi_{\mathcal{F}}^{*}\Theta_{D_{l}}=\Theta_{\mathcal{F},D_{l}}.
  2. (2)

    Let DlD_{l} and DlD_{l}^{\prime} be two different effective divisor of degree ll on CC, then ΘDl\Theta_{D_{l}} and ΘDl\Theta_{D_{l}^{\prime}} are algebraically equivalent.

Proof.

ΘDl\Theta_{D_{l}} is the descent of Θ,Dl\Theta_{\mathcal{E},D_{l}} over ss\mathcal{R}^{ss} for the universal parabolic symplectic/orthogonal bundle. The reason of descent of Θ,Dl\Theta_{\mathcal{E},D_{l}} is the same as the parabolic bundle case as in [25], [17] once we see that the pull back of polarization over PHom(VV,H)\textbf{P}Hom(V\otimes V,H) to ss\mathcal{R}^{ss} is trivial. Similarly we can show ΘDl\Theta_{D_{l}} is ample and for different choice of DlD_{l}, the theta line bundles are algebraically equivalent. ∎

For any parabolic GG bundle EE, with parabolic structure txG/Px,xDt_{x}\in G/P_{x},\forall x\in D, we define 𝐃E\mathbf{D}_{E} to be the space of infinitesimal deformation of EE, i.e. the space of isomorphism classes of parabolic GG bundles E~\tilde{E} on C[ϵ]C[\epsilon], such that E~|CE\tilde{E}|_{C}\cong E, where C[ϵ]=C×Spec([ϵ]/(ϵ2))C[\epsilon]=C\times Spec(\mathbb{C}[\epsilon]/(\epsilon^{2})).

Proposition 5.7.

There is an exact sequence:

0xDTtx(G/Px)f𝐃EgH1(C,E(Ad))00\longrightarrow\prod_{x\in D}T_{t_{x}}(G/P_{x})\stackrel{{\scriptstyle f}}{{\longrightarrow}}\mathbf{D}_{E}\stackrel{{\scriptstyle g}}{{\longrightarrow}}H^{1}(C,E(Ad))\longrightarrow 0

where Ttx(G/Px)T_{t_{x}}(G/P_{x}) is the tangent space of G/PxG/P_{x} at txt_{x}.

Proof.

Recall that H1(C,E(Ad))H^{1}(C,E(Ad)) is the infinitesimal deformation space of EE as a twisted GG bundle, so the morphism gg is given by forgetting parabolic structures. Since every twisted GG bundle can be equipped with any parabolic structure, gg is an surjection.

To determine the kernel of gg, we need to figure out how many parabolic structures we can impose on a \mathcal{E} so that the restriction to CC are the parabolic structures {txG/Px}\{t_{x}\in G/P_{x}\}. The question is local, so it is equivalent to find a parabolic subgroups P~xG([ϵ]/(ϵ2))\tilde{P}_{x}\subset G(\mathbb{C}[\epsilon]/(\epsilon^{2})) such that P~x|0=txG/Px\tilde{P}_{x}|_{0}=t_{x}\in G/P_{x}. The space of such groups is exactly xDTtx(G/Px)\prod_{x\in D}T_{t_{x}}(G/P_{x}). ∎

Corollary 5.8.

For any family of stable parabolic GG bundle \mathcal{F} over S×CS\times C,let πS:S×CS\pi_{S}:S\times C\longrightarrow S be the projection and φS:SMG,P\varphi_{S}:S\longrightarrow M_{G,P} be the induced map, then

φS(ωMG,P1)=det(RπS(Ad))1xD{i=1lxdet(Qi(S×{x}))mi(x)}\varphi_{S}^{*}(\omega_{M_{G,P}}^{-1})=\text{det}(R\pi_{S}\mathcal{F}(Ad))^{-1}\otimes\bigotimes_{x\in D}\big{\{}\bigotimes_{i=1}^{l_{x}}\text{det}\big{(}Q_{i}(\mathcal{F}_{S\times\{x\}})\big{)}^{m_{i}(x)}\big{\}}

where mi(x)=ni(x)+ni+1(x)m_{i}(x)=n_{i}(x)+n_{i+1}(x) for 1xlx11\leq x\leq l_{x}-1; mlx=nlx+nlx+1+1m_{l_{x}}=n_{l_{x}}+n_{l_{x}+1}+1 for G=Sp(2n)G=Sp(2n); mlx=nlx+nlx+11m_{l_{x}}=n_{l_{x}}+n_{l_{x}+1}-1 for G=SO(2n)G=SO(2n) and mlx=nlx+nlx+1m_{l_{x}}=n_{l_{x}}+n_{l_{x}+1} for G=SO(2n+1)G=SO(2n+1).

The main results in this section is to under certain choices of weights, the moduli space of parabolic symplectic/orthogonal bundles are Fano varieties. A normal projective variety XX is call Fano if ωX1\omega_{X}^{-1} is an ample line bundle. Our method is to compare the pull back of anti-canonical line bundle over MG,PM_{G,P} to XPssX_{P}^{ss} with theta line bundle over XPssX_{P}^{ss}. It is known that the Picard group of moduli space of symplectic/orthogonal bundles has rank one, so there exists positive integer χG\chi_{G} such that det(RπS(Ad))(detRπS)χG\text{det}(R\pi_{S}\mathcal{F}(Ad))\cong(\text{det}R\pi_{S}\mathcal{F})^{\otimes\chi_{G}}. For G=Sp(2n)G=Sp(2n), χG=n+1\chi_{G}=n+1, for G=SO(2n)G=SO(2n), χG=2n2\chi_{G}=2n-2 and for G=SO(2n+1)G=SO(2n+1), χG=2n1\chi_{G}=2n-1.

We first deal with symplectic case, since symplectic groups are simply connected. Combine Proposition 5.5 and Theorem 5.6 together, we have:

Proposition 5.9.

Let G=Sp(2n)G=Sp(2n) , K=2χGK=2\chi_{G} and a(x)\overrightarrow{a}(x) satisfying ai+1(x)ai(x)=mi(x)a_{i+1}(x)-a_{i}(x)=m_{i}(x) for 1ilx1\leq i\leq l_{x} , the moduli space of parabolic symplectic bundles are Fano.

Proof.

We show that under the condition in the proposition, ΘDl\Theta_{D_{l}} is equal to ωMG,P2χG\omega_{M_{G,P}}^{-2\chi_{G}}. The problem here is that we do not know whether MG,PM_{G,P} is Gorenstein or not, i.e. whether ωMG,P\omega_{M_{G,P}} is a line bundle. But we do know that MG,PM_{G,P} is Cohen-Macaulay and normal. Let MMG,PM^{\circ}\subset M_{G,P} be the open subset where ωMG,P\omega_{M_{G,P}} is a line bundle and points in MM^{\circ} representing stable bundles, then we have codim(MG,PM)2\text{codim}(M_{G,P}\setminus M^{\circ})\geq 2. Apply Proposition 5.5 to MM^{\circ} we see that ωMG,P2χG\omega_{M_{G,P}}^{-2\chi_{G}} and ΘDl\Theta_{D_{l}} are coincide over MM^{\circ}. Now we use Lemma 2.7 of [11], and we see that ωMG,P\omega_{M_{G,P}} is a line bundle, moreover, MG,PM_{G,P} is a Fano variety. ∎

The special orthogonal group case is different, since SO(n)SO(n) is not simply connected, and its universal cover is Spin(n)Spin(n). For any one parameter subgroup of SO(n)SO(n), we choose a lift to be a one parameter subgroup of Spin(n)Spin(n). Then if we consider the moduli space of parabolic Spin(n)Spin(n) bundles with parabolic structure given by the lifts, by Lemma 1.4 of [3], we would have a natural map: t:MSpin(n),PMSO(n),Pt:M_{Spin(n),P}\rightarrow M_{SO(n),P} which identifies MSO(n),PM_{SO(n),P} as a quotient by a finite group of MSpin(n),PM_{Spin(n),P}. By discussion in the section 6 of [2], we have:

Proposition 5.10.

The map between Picard groups: t:Pic(MSO(n),P)Pic(MSpin(n),P)t^{*}:\text{Pic}(M_{SO(n),P})\rightarrow\text{Pic}(M_{Spin(n),P}) is injective on the subgroup of infinite order elements.

Similar as before, we have:

Proposition 5.11.

Let G=SO(n)G=SO(n) , K=2χGK=2\chi_{G} and a(x)\overrightarrow{a}(x) satisfying ai+1(x)ai(x)=mi(x)a_{i+1}(x)-a_{i}(x)=m_{i}(x) for 1ilx1\leq i\leq l_{x} , the moduli space of parabolic special orthogonal bundles are Fano.

6. Globally F regular type varieties and Main theorem

Let kk be a perfect field of char(k)=p>0char(k)=p>0 and XX be a normal variety over kk. Consider

F:XXF:X\longrightarrow X

to be the absolute Frobenius map and Fe:XXF^{e}:X\rightarrow X to be the ee-th iteration of FF.

For any Weil divisor DDiv(X)D\in Div(X), we have a reflexive sheaf

𝒪X(D)=j𝒪Xsm(D)\mathcal{O}_{X}(D)=j_{*}\mathcal{O}_{X^{sm}}(D)

where j:XsmXj:X^{sm}\hookrightarrow X is the inclusion of smooth locus, and 𝒪X(D)\mathcal{O}_{X}(D) is an invertible sheaf if and only if DD is a Cartier divisor.

Definition 6.1.

Let XX and DD be as above, XX is called stably Frobenius D-split if the natural homomorphism

𝒪XFe𝒪X(D)\mathcal{O}_{X}\longrightarrow F^{e}_{*}\mathcal{O}_{X}(D)

is split as an 𝒪X\mathcal{O}_{X} homomorphism for some e>0e>0. And XX is called globally F-regular if XX is stably Frobenius DD-split for any effective divisor DD.

We state the following lemma about globally F-regular varieties, for proof and more details, please refer to [25], [21].

Lemma 6.2 (Corollary 6.4 of [19]).

Let f:XYf:X\rightarrow Y be a morphism of normal varieties over kk. Assume that the natural map

f#:𝒪Yf𝒪Xf^{\#}:\mathcal{O}_{Y}\longrightarrow f_{*}\mathcal{O}_{X}

splits as an 𝒪Y\mathcal{O}_{Y} homomorphism, then if XX is globally F-regular, so is YY.

Now we let KK be a field of characteristic zero.

For any scheme XX over KK, there is a finitely generated \mathbb{Z}-algebra RKR\subset K such that XX is ”defined” over RR. That is, there is a flat RR-scheme

XRS=SpecRX_{R}\longrightarrow S=\text{Spec}R

such that XK:=XR×SSpecKXX_{K}:=X_{R}\times_{S}\text{Spec}K\cong X. XRSX_{R}\longrightarrow S is called an integral model of X/KX/K. For any closed point sSs\in S, Xs:=XR×SSpec(k(s)¯)X_{s}:=X_{R}\times_{S}\text{Spec}(\overline{k(s)}) is called the ”modulo pp reduction” of XX, where p=char(k(s))>0p=char(k(s))>0.

Definition 6.3.

A variety XX over KK is called of globally F-regular type if its ”modulo pp reduction” of XX are globally F-regular for a dense set of pp for some integral model XRSX_{R}\rightarrow S.

Globally F-regular type varieties have many nice properties, which we will state some of them as the following theorem. Again, for proof and more details, please refer to [25] and [21].

Theorem 6.4.

Let XX be a projective variety over KK, if XX is of globally F-regular type, then:

  1. (1)

    XX is normal, Cohen-Macaulay with rational singularities. If XX is \mathbb{Q}-Gorenstein, then XX has log terminal singularities.

  2. (2)

    For any nef line bundle \mathcal{L} over XX, we have Hi(X,)=0{\rm H}^{i}(X,\mathcal{L})=0, for any i>0i>0. In particular, Hi(X,𝒪X)=0{\rm H}^{i}(X,\mathcal{O}_{X})=0 for any i>0i>0.

Our main theorem of this paper is:

Theorem 6.5.

The moduli space of parabolic symplectic/orthogonal bundles MPM_{P} over a smooth projective curve CC over \mathbb{C} is of globally F-regular type.

Corollary 6.6.

Let ΘDl\Theta_{D_{l}} be the theta line bundle over MG,PM_{G,P} define before, then

Hi(MP,ΘDl)=0{\rm H}^{i}(M_{P},\Theta_{D_{l}})=0

for any i>0i>0.

Our beginning example of globally F-regular type variety is Fano variety.

Proposition 6.7 (Proposition 6.3 in [21]).

A Fano variety over KK with at most rational singularities is of globally F-regular type.

With our beginning example, the next step is to ask whether Lemma 6.2 holds in characteristic zero. To answer such question, in [25], they introduced the following:

Definition 6.8.

A morphism f:XYf:X\rightarrow Y of varieties over KK is called pp-compatible if there is an integral model

fR:XRYRf_{R}:X_{R}\longrightarrow Y_{R}

such that, if for any sS=SpecRs\in S=\text{Spec}R, we put Xs=XR×SSpeck(s)¯X_{s}=X_{R}\times_{S}\text{Spec}\overline{k(s)} , Ys=YR×SSpeck(s)¯Y_{s}=Y_{R}\times_{S}\text{Spec}\overline{k(s)} and consider

Xs\textstyle{X_{s}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}js\scriptstyle{j_{s}}fs\scriptstyle{f_{s}}XR\textstyle{X_{R}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}fR\scriptstyle{f_{R}}Ys\textstyle{Y_{s}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}is\scriptstyle{i_{s}}YR\textstyle{Y_{R}}

then we have that isfR𝒪XR=fsjs𝒪XRi_{s}^{*}f_{R*}\mathcal{O}_{X_{R}}=f_{s*}j_{s}^{*}\mathcal{O}_{X_{R}} holds for a dense set of ss.

It can be shown that if f:XYf:X\rightarrow Y is a flat proper morphisms such that Rif𝒪X=0\textbf{R}^{i}f_{*}\mathcal{O}_{X}=0 for all i1i\geq 1, then ff is pp-compatible.

To prove our main theorem, we need to introduce a key proposition from [25].

Let (,L)(\mathcal{R}^{\prime},L^{\prime}) and (,L)(\mathcal{R},L) be two polarized projective varieties over KK, with linear actions by a reductive group scheme GG over KK respectively. We use ()ss(L)(\mathcal{R}^{\prime})^{ss}(L^{\prime})\subseteq\mathcal{R}^{\prime} and ss(L)\mathcal{R}^{ss}(L)\subseteq\mathcal{R} to denote the GIT semistable locus, then there are projective GIT quotients:

ψ:ss(L)Y:=ss(L)//G,φ:()ss(L)Z:=()ss(L)//G\psi:\mathcal{R}^{ss}(L)\rightarrow Y:=\mathcal{R}^{ss}(L)//G\ ,\ \varphi:(\mathcal{R}^{\prime})^{ss}(L^{\prime})\rightarrow Z:=(\mathcal{R}^{\prime})^{ss}(L^{\prime})//G
Proposition 6.9 (Proposition 2.10 of [25]).

Let \mathcal{R}, \mathcal{R}^{\prime} as above. Considering the following diagram, assume

  1. (1)

    there is a GG-invariant pp-compatible morphism f^:\hat{f}:\mathcal{R}^{\prime}\rightarrow\mathcal{R} such that f^𝒪=𝒪\hat{f}_{*}\mathcal{O}_{\mathcal{R^{\prime}}}=\mathcal{O}_{\mathcal{R}};

  2. (2)

    there is a GG-invariant open subset W()ss(L)W\subset(\mathcal{R}^{\prime})^{ss}(L^{\prime}) such that

    Codim(W)2,X^=φ1φ(X^)\emph{Codim}(\mathcal{R}^{\prime}\setminus W)\geq 2,\ \hat{X}=\varphi^{-1}\varphi(\hat{X})

    where X^=Wf^1(ss(L))\hat{X}=W\cap\hat{f}^{-1}(\mathcal{R}^{ss}(L)). And we put X=φ(X^)X=\varphi(\hat{X}).

Then if ZZ is of globally F-regular type, so is YY.

()ss(L)\textstyle{(\mathcal{R}^{\prime})^{ss}(L^{\prime})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}φ\scriptstyle{\varphi}\textstyle{\mathcal{R}^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f^\scriptstyle{\hat{f}}\textstyle{\mathcal{R}}ss(L)\textstyle{\mathcal{R}^{ss}(L)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ\scriptstyle{\psi}W\textstyle{W\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}X^=Wf^1(ss(L))\textstyle{\hat{X}=W\cap\hat{f}^{-1}(\mathcal{R}^{ss}(L))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Z\textstyle{Z}X\textstyle{X\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f\scriptstyle{f}Y\textstyle{Y}

Finally, we will prove our main theorem:

Proof of Theorem 6.5.

We choose an effective divisor DD^{\prime} of CC such that DD=D^{\prime}\cap D=\emptyset, degDD^{\prime} being even and

deg(D)+deg(D)2χG+(r1)(g1)2\dfrac{\text{deg}(D)+\text{deg}(D^{\prime})}{2\chi_{G}}+(r-1)(g-1)\geq 2

and for each xDx\in D^{\prime}, we put n(x)=(1,,1)\overrightarrow{n}(x)=(1,\cdots,1). Let ZZ^{\prime} be the scheme parametrizing symplectic/orthogonal bundles (E,ω)(E,\omega) where ω:EE𝒪C(DD)\omega:E\otimes E\rightarrow\mathcal{O}_{C}(-D-D^{\prime}) as we constructed in section 3. We see that ZZZ^{\prime}\cong Z. Then we let

=×ZxDDFlagn(x)(x)=×Z(×ZxDFlagn(x)(x))f^.\mathcal{R}^{\prime}=\underset{x\in D\cup D^{\prime}}{\times_{Z}}Flag_{\vec{n}(x)}(\mathcal{F}_{x})=\mathcal{R}\times_{Z}\big{(}\underset{x\in D^{\prime}}{\times_{Z}}Flag_{\vec{n}(x)}(\mathcal{F}_{x})\big{)}\stackrel{{\scriptstyle\hat{f}}}{{\longrightarrow}}\mathcal{R}.

So f^:\hat{f}:\mathcal{R}^{\prime}\rightarrow\mathcal{R} is a flag bundle and hence pp-compatible with f^𝒪=𝒪\hat{f}_{*}\mathcal{O}_{\mathcal{R}^{\prime}}=\mathcal{O}_{\mathcal{R}}. We choose polarization for \mathcal{R}^{\prime} and \mathcal{R} as the ones given in Section 3, say LL^{\prime} and LL Clearly there are SL(V)SL(V) action on \mathcal{R}^{\prime} and \mathcal{R} and f^\hat{f} is SL(V)SL(V)-invariant.

Now we put K=2χGK=2\chi_{G} and give weights for \mathcal{R}^{\prime} by a(x)\overrightarrow{a}(x) satisfying ai+1(x)ai(x)=mi(x)a_{i+1}(x)-a_{i}(x)=m_{i}(x) and ai(x)+a2lx+2i(x)=Ka_{i}(x)+a_{2l_{x}+2-i}(x)=K for 1ilx1\leq i\leq l_{x} and any xDDx\in D\cup D^{\prime}. So by Proposition 5.9 and 5.11 we see that Z:=()ss(L)//SL(V)Z:=(\mathcal{R}^{\prime})^{ss}(L^{\prime})//SL(V) is a Fano variety. We use φ:()ssZ\varphi:(\mathcal{R}^{\prime})^{ss}\rightarrow Z to denote the quotient map.

Moreover, if one let W=()sW=(\mathcal{R}^{\prime})^{s}, X^=Wf^1(ss)\hat{X}=W\cap\hat{f}^{-1}(\mathcal{R}^{ss}) and X=φ(X^)X=\varphi(\hat{X}) then clearly X^=φ1(X)\hat{X}=\varphi^{-1}(X). By Corollary 4.12 and our assumption, we would have: Codim(W)2\text{Codim}(\mathcal{R}^{\prime}\setminus W)\geq 2. Now Proposition 6.9 shows that the moduli space of parabolic symplectic/orthogonal bundles Y:=(L)ss//SL(V)Y:=\mathcal{R}(L)^{ss}//SL(V) is of globally FF-regular type. ∎

References

  • [1] Vikraman Balaji, Indranil Biswas, and Donihakkalu S. Nagaraj. Principal bundles over projective manifolds with parabolic structure over a divisor. Tohoku Math. J. (2), 53(3):337–367, 2001.
  • [2] YVES Beauville, ARNAUD and Laszio and CHRISTOPH Sorger. The Picard group of the moduli of GG-bundles on a curve. Compositio Mathematica, 112(2):183–216, 1998.
  • [3] Usha Bhosle and A. Ramanathan. Moduli of parabolic G-bundles on curves. Mathematische Zeitschrift, 202(2):161–180, Jun 1989.
  • [4] Indranil Biswas. Parabolic bundles as orbifold bundles. Duke Math. J., 88(2):305–325, 06 1997.
  • [5] Indranil Biswas. The Atiyah bundle and connections on a principal bundle. Indian Academy of sciences, 120, 06 2010.
  • [6] Indranil Biswas, Souradeep Majumder, and Michael Lennox Wong. Orthogonal and symplectic parabolic bundles. Journal of Geometry and Physics, 61:1462–1475, 08 2011.
  • [7] Francesca Gavioli. Theta functions on the moduli space of parabolic bundles. International Journal of Mathematics, 15:259–287, 2004.
  • [8] Tomás L. Gómez and Ignacio Sols. Stable tensors and moduli space of orthogonal sheaves. 2003.
  • [9] Daniel Huybrechts and Manfred Lehn. The Geometry of Moduli Spaces of Sheaves. Cambridge Mathematical Library. Cambridge University Press, 2 edition, 2010.
  • [10] Verdier Jean-Louis and Le Potier. Joseph. Module Des Fibrés Stables Sur Les Courbes Algébriques. Progress in Mathematics 54. 1985.
  • [11] Shrawan Kumar and M. S. Narasimhan. Picard group of the moduli spaces of G-bundles. Mathematische Annalen, 308(1):155–173, May 1997.
  • [12] Shrawan Kumar, M. S. Narasimhan, and A. Ramanathan. Infinite Grassmannians and moduli spaces of G-bundles. Mathematische Annalen, 300(1):41–75, Sep 1994.
  • [13] V. B. Mehta and T. R. Ramadas. Moduli of vector bundles, Frobenius splitting, and invariant theory. Annals of Mathematics, 144(2):269–313, 1996.
  • [14] V. B. Mehta and A. Ramanathan. Frobenius splitting and cohomology vanishing for Schubert varieties. Annals of Mathematics, 122(1):27–40, 1985.
  • [15] V. B. Mehta and C. S. Seshadri. Moduli of vector bundles on curves with parabolic structures. Mathematische Annalen, 248(3):205–239, Oct 1980.
  • [16] Joel L.Roberts Melvin Hochster. Rings of invariants of reductive group acting on regular rings are Cohen-Macaulay. Advances in Mathematics, 13:115–175, 1974.
  • [17] Christian Pauly. Espaces de modules de fibrés paraboliques et blocs conformes. Duke Math. J., 84(1):217–235, 07 1996.
  • [18] A. Ramanathan. Moduli for principal bundles over algebraic curves: II. Proceedings of the Indian Academy of Sciences - Mathematical Sciences, 106(4):421–449, Nov 1996.
  • [19] Karl Schwede and Karen E. Smith. Globally F-regular and log Fano varieties. Advances in Mathematics, 224(3):863 – 894, 2010.
  • [20] C. S. Seshadri. Moduli of π\pi -Vector Bundles over an Algebraic Curve, pages 139–260. Springer Berlin Heidelberg, Berlin, Heidelberg, 2011.
  • [21] Karen E. Smith. Globally F-regular varieties: applications to vanishing theorems for quotients of Fano varieties. Michigan Math. J., 48(1):553–572, 2000.
  • [22] Xiaotao Sun. Degeneration of moduli spaces and generalized theta functions. Journal of Algebraic Geometry, 9:459–527, 2000.
  • [23] Xiaotao Sun. Factorization of generalized theta functions revisited. Algebra Colloquium, 24(01):1–52, 2017.
  • [24] Xiaotao Sun and Mingshuo Zhou. A finite dimensional proof of the verlinde formula. Science China Mathematics, 63:1935–1964, 2020.
  • [25] Xiaotao Sun and Mingshuo Zhou. Globally F-regular type of moduli spaces. Mathematische Annalen, 144:1245–1270, 2020.
  • [26] Kôji Yokogawa. Infinitesimal deformation of parabolic Higgs sheaves. International Journal of Mathematics, 6(1):125–148, 1995.