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

On subregular slices of the elliptic Grothendieck-Springer resolution

Dougal Davis School of Mathematics, University of Edinburgh, James Clerk Maxwell Building, The King’s Buildings, Peter Guthrie Tait Road, Edinburgh, EH9 3FD, UK [email protected]
Abstract.

In [davis19], the author constructed an elliptic version of the Grothendieck-Springer resolution for the stack BunG\mathrm{Bun}_{G} of principal bundles under a simply connected simple group GG on an elliptic curve EE. This is a simultaneous log resolution of a map from BunG\mathrm{Bun}_{G} to the union of the coarse moduli space of semistable GG-bundles and a single stacky point. In this paper, we study singularities, resolutions and deformations coming from subregular slices of this elliptic Grothendieck-Springer resolution. More precisely, we construct explicit slices of BunG\mathrm{Bun}_{G} through all subregular unstable bundles, for every GG. For GSL2G\neq SL_{2}, we describe the pullbacks of the elliptic Grothendieck-Springer resolution to these slices as concrete varieties, extending and refining earlier work of I. Grojnowski and N. Shepherd-Barron, who related these varieties to del Pezzo surfaces in type EE. We use the resolutions to identify the singularities of the unstable locus of the subregular slices, and prove that that the extended coarse moduli space map gives deformations that are miniversal among torus-equivariant deformations with appropriate weights.

1. Introduction

Since the work of E. Brieskorn [brieskorn70] and P. Slodowy [slodowy80a], it has been well known that du Val (aka simple, Kleinian, ADEADE, etc.) singularities of algebraic surfaces arise naturally in the geometry of simple algebraic groups and their Lie algebras. If GG is a simply connected simple algebraic group, say over an algebraically closed field kk of characteristic 0, then the cone 𝒩\mathcal{N} of nilpotent elements inside the Lie algebra 𝔤\mathfrak{g} of GG is a singular variety canonically associated to 𝔤\mathfrak{g}. The cone 𝒩\mathcal{N} has dimension dimGl\dim G-l, where ll is the rank of GG; to obtain a surface singularity, one first chooses a subregular nilpotent element x𝔤x\in\mathfrak{g} (i.e., one satisfying dimStabG(x)=l+2\dim\mathrm{Stab}_{G}(x)=l+2) and a transversal slice ZZ (a locally closed subvariety Z𝔤Z\subseteq\mathfrak{g} transverse to all GG-orbits for the adjoint representation) such that xx is the unique subregular nilpotent in ZZ. Then Z𝒩Z\cap\mathcal{N} is a surface, with a unique du Val singularity at xx whose Dynkin diagram is the same as that of GG when GG is of type AA, DD or EE.

The singular surfaces constructed in this way are also furnished with natural Lie-theoretic deformations and resolutions. The deformations arise from the (additive) adjoint quotient map

χadd:𝔤𝔤//G=Speck[𝔤]G,\chi^{add}\colon\mathfrak{g}\longrightarrow\mathfrak{g}{/\mkern-6.0mu/}G=\operatorname{\mathrm{Spec}}k[\mathfrak{g}]^{G}, (1.0.1)

where GG acts on 𝔤\mathfrak{g} via the adjoint representation. The morphism χadd\chi^{add} is a flat family of affine varieties with central fibre 𝒩=(χadd)1(0)\mathcal{N}=(\chi^{add})^{-1}(0); the restriction χZ=χ|Z:Z𝔤//G\chi_{Z}=\chi|_{Z}\colon Z\to\mathfrak{g}{/\mkern-6.0mu/}G gives a flat deformation of the singular surface Z𝒩=χZ1(0)Z\cap\mathcal{N}=\chi_{Z}^{-1}(0). In types ADEADE, it was proved by Brieskorn that this recovers the miniversal deformation of the singular surface Z𝒩Z\cap\mathcal{N}, while in types BCFGBCFG it was shown by Slodowy that the deformation is miniversal among those preserving a “folding symmetry” of the ADEADE du Val singularity.

The resolutions arise from a commutative diagram

𝔤~{\tilde{\mathfrak{g}}}𝔤{\mathfrak{g}}𝔱{\mathfrak{t}}𝔱//W𝔤//G,{\mathfrak{t}{/\mkern-6.0mu/}W\cong\mathfrak{g}{/\mkern-6.0mu/}G,}ψadd\scriptstyle{\psi^{add}}χ~add\scriptstyle{\tilde{\chi}^{add}}χadd\scriptstyle{\chi^{add}}q\scriptstyle{q} (1.0.2)

where 𝔤~=G×B𝔟\tilde{\mathfrak{g}}=G\times^{B}\mathfrak{b} for BGB\subseteq G a Borel subgroup with Lie algebra 𝔟\mathfrak{b}, 𝔱\mathfrak{t} is the Lie algebra of a maximal torus TGT\subseteq G and W=NG(T)/TW=N_{G}(T)/T is the Weyl group. The diagram (1.0.2) is called the (additive) Grothendieck-Springer resolution; it is a simultaneous resolution of singularities in the sense that χ~add\tilde{\chi}^{add} is smooth, ψadd\psi^{add} is proper, and (χ~add)1(t)(χadd)1(q(t))(\tilde{\chi}^{add})^{-1}(t)\to(\chi^{add})^{-1}(q(t)) is a resolution of singularities for all t𝔱t\in\mathfrak{t}. Setting Z~=𝔤~×𝔤Z\tilde{Z}=\tilde{\mathfrak{g}}\times_{\mathfrak{g}}Z, the induced diagram

Z~{\tilde{Z}}Z{Z}𝔱{\mathfrak{t}}𝔱//W{\mathfrak{t}{/\mkern-6.0mu/}W}ψZ\scriptstyle{\psi_{Z}}χ~Z\scriptstyle{\tilde{\chi}_{Z}}χZ\scriptstyle{\chi_{Z}}

is a simultaneous resolution for χZ\chi_{Z}. One pleasing way to identify the du Val singularity of χZ1(0)\chi_{Z}^{-1}(0) is to compute the fibre ψZ1(x)\psi_{Z}^{-1}(x), and to show that this gives the correct Dynkin configuration of (2)(-2)-curves on the resolution χ~Z1(0)\tilde{\chi}_{Z}^{-1}(0).

This paper is concerned with an elliptic version of the above additive story. (From now on, kk can be any algebraically closed field.) Building on earlier work [helmke-slodowy04][ben-zvi-nadler15][grojnowski-shep19], the author constructed in [davis19] a commutative diagram

Bun~G{\widetilde{\mathrm{Bun}}_{G}}BunG{\mathrm{Bun}_{G}}ΘY1/𝔾m{\Theta_{Y}^{-1}/\mathbb{G}_{m}}(Y^//W)/𝔾m,{(\widehat{Y}{/\mkern-6.0mu/}W)/\mathbb{G}_{m},}ψ\scriptstyle{\psi}χ~\scriptstyle{\tilde{\chi}}χ\scriptstyle{\chi}q\scriptstyle{q} (1.0.3)

where BunG\mathrm{Bun}_{G} is the stack of principal GG-bundles on an elliptic curve EE, Bun~G\widetilde{\mathrm{Bun}}_{G} is the Kontsevich-Mori compactification of the stack BunB0\mathrm{Bun}_{B}^{0} of degree 0 BB-bundles on EE, ΘY1\Theta_{Y}^{-1} is an anti-ample WW-linearised line bundle on the coarse moduli space Y=Hom(𝕏(T),Pic0(E))Y=\mathrm{Hom}(\mathbb{X}^{*}(T),\mathrm{Pic}^{0}(E)) of degree 0 TT-bundles on EE, Y^\widehat{Y} is the affine cone over YY obtained by contracting the zero section of ΘY1\Theta_{Y}^{-1} to a point, and /𝔾m/\mathbb{G}_{m} denotes the stack quotient by 𝔾m\mathbb{G}_{m}. Away from the image of the cone point of Y^\widehat{Y}, χ\chi agrees with the semistable coarse moduli space map χss:BunGssY//W\chi^{ss}\colon\mathrm{Bun}_{G}^{ss}\to Y{/\mkern-6.0mu/}W of R. Friedman and J. Morgan [friedman-morgan98], and the preimage of the (stacky) cone point is precisely the locus of unstable bundles.

The diagram (1.0.3) is called the elliptic Grothendieck-Springer resolution, and is closely analogous to a stacky version of (1.0.2) where the varieties 𝔤\mathfrak{g} and 𝔤~\tilde{\mathfrak{g}} are replaced by the stack quotients 𝔤/G\mathfrak{g}/G and 𝔤~/G\tilde{\mathfrak{g}}/G. It was shown in [davis19]*Corollary 4.4.7 that it is a simultaneous log resolution with respect to the zero section of ΘY1\Theta_{Y}^{-1} [davis19]*Definition 1.0.3; this means that the total space Bun~G\widetilde{\mathrm{Bun}}_{G} is smooth, χ~\tilde{\chi} is smooth away from the zero section, the preimage of the zero section is a divisor with normal crossings, the map ψ\psi is proper (with finite relative stabilisers) and for all yΘY1/𝔾my\in\Theta_{Y}^{-1}/\mathbb{G}_{m}, the map χ~1(y)χ1(q(y))\tilde{\chi}^{-1}(y)\to\chi^{-1}(q(y)) is an isomorphism over a dense open subset of the target. In particular, the restriction to semistable bundles is a genuine simultaneous resolution, and for yy in the zero section of ΘY1\Theta_{Y}^{-1}, each irreducible component of the locus χ1(q(y))=χ1(0)\chi^{-1}(q(y))=\chi^{-1}(0) of unstable bundles is resolved by some component of χ~1(y)\tilde{\chi}^{-1}(y).

Subregular slices in the elliptic setting have been studied by S. Helmke and P. Slodowy [helmke-slodowy01] [helmke-slodowy04] and I. Grojnowski and N. Shepherd-Barron [grojnowski-shep19]. In [helmke-slodowy01], Helmke and Slodowy classified the subregular unstable bundles (Definition 2.1.1) and gave simple descriptions of their coarse moduli spaces for all simply connected groups GG; these bundles play the role of subregular nilpotent elements in elliptic Springer theory. In [helmke-slodowy04], they constructed a version of the coarse quotient map χ\chi using loop groups, and briefly sketched the associated surface singularities arising from slices through subregular unstable bundles in types AA, DD and EE. In [grojnowski-shep19], Grojnowski and Shepherd-Barron considered certain subregular slices ZBunGZ\to\mathrm{Bun}_{G} for GG of types D5=E5D_{5}=E_{5}, E6E_{6}, E7E_{7} and E8E_{8} only, and studied simultaneous log resolutions

Z~{\tilde{Z}}Z{Z}ΘY1{\Theta_{Y}^{-1}}Y^//W{\widehat{Y}{/\mkern-6.0mu/}W}ψZ\scriptstyle{\psi_{Z}}χ~Z\scriptstyle{\tilde{\chi}_{Z}}χZ\scriptstyle{\chi_{Z}} (1.0.4)

deduced from (1.0.3), where Z~=Bun~G×BunGZ\tilde{Z}=\widetilde{\mathrm{Bun}}_{G}\times_{\mathrm{Bun}_{G}}Z. They showed that, in their examples, the preimage χ~Z1(0ΘY1)\tilde{\chi}_{Z}^{-1}(0_{\Theta_{Y}^{-1}}) of the zero section decomposes as a simple normal crossings divisor

χ~Z1(0ΘY1)=D0+D1+Q,\tilde{\chi}_{Z}^{-1}(0_{\Theta_{Y}^{-1}})=D_{0}+D_{1}+Q,

where D0YD_{0}\to Y is a family of resolutions of the singular surface χZ1(0)\chi_{Z}^{-1}(0), D1YD_{1}\to Y is some other family of projective surfaces, and QYQ\to Y is a 1×1\mathbb{P}^{1}\times\mathbb{P}^{1}-bundle. Moreover, they showed that contracting QQ along a ruling and flopping an unknown number of curves from D0D_{0} to D1D_{1} produces a birational modification Z~Z~\tilde{Z}\dashrightarrow\tilde{Z}^{-} such that the preimage of 0ΘY10_{\Theta_{Y}^{-1}} decomposes as D0+D1D_{0}^{-}+D_{1}^{-}, where D0D_{0}^{-} is a line bundle over Y×EY\times E and D1YD_{1}^{-}\to Y is a family of del Pezzo surfaces of degree 9l9-l, from which they deduced that χZ1(0)\chi_{Z}^{-1}(0) has a simply elliptic singularity of the same degree. Their results show that the elliptic Grothendieck-Springer resolution in some sense “contains” the well-known combinatorial correspondence between exceptional groups, del Pezzo surfaces, and simply elliptic singularities.

Remark 1.0.1.

One of the nice features of Grojnowski and Shepherd-Barron’s construction is that the stack quotients by 𝔾m\mathbb{G}_{m} in the bottom row of (1.0.3) are exchanged for a global action of 𝔾m\mathbb{G}_{m} on the sliced diagram (1.0.4). This desirable behaviour is axiomatised by the notion of equivariant slices in [davis19]*Definition 4.1.9; these are stacks ZZ equipped with an action of a torus HH (the equivariance group), a morphism Z/HBunGZ/H\to\mathrm{Bun}_{G} (or to the rigidification BunG,rig\mathrm{Bun}_{G,rig} [davis19]*§2.2), and a lift of Z(Y^//W)/𝔾mZ\to(\widehat{Y}{/\mkern-6.0mu/}W)/\mathbb{G}_{m} to an HH-equivariant morphism ZY^//WZ\to\widehat{Y}{/\mkern-6.0mu/}W, where HH acts on Y^//W\widehat{Y}{/\mkern-6.0mu/}W through some fixed weight H𝔾mH\to\mathbb{G}_{m}, such that the morphism ZBunGZ\to\mathrm{Bun}_{G} (or BunG,rig\mathrm{Bun}_{G,rig}) is smooth modulo translations.

The goal of the present work is to describe all the singularities and log resolutions obtained from the elliptic Grothendieck-Springer resolution by taking equivariant slices through subregular unstable bundles, for all simply connected groups GG.

Our first main result gives the existence of an equivariant slice with particularly nice properties through any subregular bundle (when GSL2G\neq SL_{2}). In order to to ensure the existence of slices ZZ with generically trivial inertia, we have chosen to work with the rigidified stack BunG,rig\mathrm{Bun}_{G,rig} (cf. [davis19]*§2.2) obtained by taking the quotient of all automorphism groups in BunG\mathrm{Bun}_{G} by the centre Z(G)Z(G) of GG.

Theorem 1.0.2.

Let ξGE\xi_{G}\to E be a subregular unstable GG-bundle, and assume that GSL2G\neq SL_{2}. Then there exists an equivariant slice ZBunG,rigZ\to\mathrm{Bun}_{G,rig} with equivariance group

H={𝔾m×𝔾m,in typeA,𝔾m,otherwise,H=\begin{cases}\mathbb{G}_{m}\times\mathbb{G}_{m},&\text{in type}\;A,\\ \mathbb{G}_{m},&\text{otherwise,}\end{cases}

with the following properties.

  1. (1)

    The HH-fixed locus Z0=ZHZ_{0}=Z^{H} is a proper Artin stack with finite and generically trivial inertia.

  2. (2)

    The set of points zZz\in Z such that the associated GG-bundle is subregular unstable is equal to Z0Z_{0}, and the given family identifies the coarse moduli space of Z0Z_{0} with the connected component of the coarse moduli space of subregular unstable GG-bundles up to translation containing ξG\xi_{G}.

  3. (3)

    All nonempty geometric fibres of the morphism Z0BunG,rig/EZ_{0}\to\mathrm{Bun}_{G,rig}/E are connected.

There are no essentially new ideas in the proof of Theorem 1.0.2: following a suggestion of Helmke and Slodowy [helmke-slodowy01]*Remark 5.14, the slices ZBunG,rigZ\to\mathrm{Bun}_{G,rig} are constructed by parabolic induction from regular slices Z0BunL,rigZ_{0}\to\mathrm{Bun}_{L,rig}, for LL the Harder-Narasimhan Levi of ξG\xi_{G}, which are either obvious or in turn constructed by parabolic induction from a single unstable LL-bundle according to the recipe of Friedman and Morgan [friedman-morgan00]. The only new thing we do in Theorem 1.0.2 is to check by hand that the morphisms ZBunG,rigZ\to\mathrm{Bun}_{G,rig} constructed in this way are actually equivariant slices with the desired properties.

For each of the equivariant slices ZZ constructed in Theorem 1.0.2, we get a morphism

χ~Z:Z~=Bun~G,rig×BunG,rigZΘY1.\tilde{\chi}_{Z}\colon\tilde{Z}=\widetilde{\mathrm{Bun}}_{G,rig}\times_{\mathrm{Bun}_{G,rig}}Z\longrightarrow\Theta_{Y}^{-1}.

Our second main result gives very explicit descriptions of the unstable fibres χ~Z1(y)\tilde{\chi}_{Z}^{-1}(y) for y0ΘY1y\in 0_{\Theta_{Y}^{-1}}. This result is really the core content of the paper.

Theorem 1.0.3.

Assume that GSL2G\neq SL_{2}, let ξGE\xi_{G}\to E be a subregular unstable GG-bundle, and let ZBunG,rigZ\to\mathrm{Bun}_{G,rig} be the equivariant slice of Theorem 1.0.2. Then we have the following.

  1. (1)

    The preimage of the zero section of ΘY1\Theta_{Y}^{-1} decomposes as a divisor with normal crossings

    χ~Z1(0ΘY1)=dDαi(Z)+Dαj(Z)+Dαi+αj(Z),\tilde{\chi}_{Z}^{-1}(0_{\Theta_{Y}^{-1}})=dD_{\alpha_{i}^{\vee}}(Z)+D_{\alpha_{j}^{\vee}}(Z)+D_{\alpha_{i}^{\vee}+\alpha_{j}^{\vee}}(Z),

    where each component Dλ(Z)D_{\lambda}(Z) is smooth over YY, and

    d={1,ifξGis of typeA,B,DorE,2,ifξGis of typeCorF,3,ifξGis of typeG.d=\begin{cases}1,&\text{if}\;\;\xi_{G}\;\text{is of type}\;A,B,D\;\text{or}\;E,\\ 2,&\text{if}\;\;\xi_{G}\;\text{is of type}\;C\;\text{or}\;F,\\ 3,&\text{if}\;\;\xi_{G}\;\text{is of type}\;G.\end{cases}
  2. (2)

    The divisor Dαj(Z)D_{\alpha_{j}^{\vee}}(Z) is isomorphic to the iterated blowup of a line bundle D1D_{1} on Y×EY\times E at a specified sequence of sections (given in Proposition 3.4.1) over YY.

  3. (3)

    Each fibre of the morphism Dαi+αj(Z)YD_{\alpha_{i}^{\vee}+\alpha_{j}^{\vee}}(Z)\to Y is isomorphic to the Hirzebruch surface 𝔽d1\mathbb{F}_{d-1}.

  4. (4)

    The divisor Dαi(Z)D_{\alpha_{i}^{\vee}}(Z) is the iterated blowup of a smooth family of surfaces D1YD_{1}^{\prime}\to Y at a specified sequence of sections (given in Proposition 3.6.1) over YY, where each fibre of D1YD_{1}^{\prime}\to Y is isomorphic to

    • a line bundle on EE, if ξG\xi_{G} is of type AA,

    • one of the Hirzebruch surfaces 𝔽0\mathbb{F}_{0} or 𝔽2\mathbb{F}_{2}, if ξG\xi_{G} is of type CC, DD or FF,

    • one of the stacky Hirzebruch surfaces (1,2)(𝒪𝒪(1))\mathbb{P}_{\mathbb{P}(1,2)}(\mathcal{O}\oplus\mathcal{O}(1)) or (1,2)(𝒪𝒪(3))\mathbb{P}_{\mathbb{P}(1,2)}(\mathcal{O}\oplus\mathcal{O}(3)), if ξG\xi_{G} is of type BB, or

    • the projective plane 2\mathbb{P}^{2}, if ξG\xi_{G} is of type EE or GG.

Remark 1.0.4.

In Theorem 1.0.3, we have referred to the type of the subregular unstable GG-bundle ξG\xi_{G}, rather than to the type of the group GG. This follows the terminology introduced in §2.1. The idea is that a given algebraic group GG may belong to multiple series in the classification (the relevant examples here being D5=E5D_{5}=E_{5} and B3=F3B_{3}=F_{3}); in these cases, there are connected components of the locus of subregular unstable bundles corresponding to each of the different series.

Remark 1.0.5.

In type ElE_{l}, Theorem 1.0.3 recovers Grojnowski and Shepherd-Barron’s result discussed above, with D0=Dαj(Z)D_{0}=D_{\alpha_{j}^{\vee}}(Z), D1=Dαi(Z)D_{1}=D_{\alpha_{i}^{\vee}}(Z) and Q=Dαi+αj(Z)Q=D_{\alpha_{i}^{\vee}+\alpha_{j}^{\vee}}(Z). Moreover, the slightly mysterious flopping curves are made manifest in our description as the exceptional fibres of the blowups of the line bundle D1D_{1} (excluding the last one, which undoes the contraction of QQ). In particular, the detailed statement Proposition 3.6.1 specifies the exact number (n0=l4n_{0}=l-4) and configuration of these curves, which was not accessible using the Grojnowski and Shepherd-Barron’s proof. The del Pezzo surfaces also appear very concretely as blowups of D1=2D_{1}^{\prime}=\mathbb{P}^{2} at ll points; the first 44 are the blowups in (4) giving Dαi(Z)D_{\alpha_{i}^{\vee}}(Z), and the remaining l4l-4 are the result of the flops.

As an application of Theorem 1.0.3, we deduce the following descriptions of the singular surfaces χZ1(0)\chi_{Z}^{-1}(0) and their deformations. For completeness, we have also included the case G=SL2G=SL_{2} with the subregular slice Z=IndTG(Z0)Z=\mathrm{Ind}_{T}^{G}(Z_{0}) with equivariance group 𝔾m\mathbb{G}_{m} of Remark 2.2.10, although this slice does not satisfy the hypotheses of Theorem 1.0.2.

Theorem 1.0.6 (Theorems 4.1.3 and 4.2.9).

If the characteristic of kk is not 22 or 33, then the surface χZ1(0)\chi_{Z}^{-1}(0) can be constructed explicitly as follows.

  1. (1)

    In type AlA_{l}, l>1l>1, χZ1(0)\chi_{Z}^{-1}(0) is obtained by gluing together two line bundles on EE along their zero sections.

  2. (2)

    In types CC and DD (resp., BB), χZ1(0)\chi_{Z}^{-1}(0) is obtained by identifying points in the fibres of a degree 22 map E1E\to\mathbb{P}^{1} (resp., E(1,2)E\to\mathbb{P}(1,2)) inside the zero section of a line bundle on EE.

  3. (3)

    In types A1A_{1}, EE, FF and GG, χZ1(0)\chi_{Z}^{-1}(0) is a cone obtained by contracting the zero section of a line bundle on EE to a point.

In each case, the deformation χZ:ZY^//W\chi_{Z}\colon Z\to\widehat{Y}{/\mkern-6.0mu/}W is miniversal among HH-equivariant deformations with weights in >0λ\mathbb{Z}_{>0}\lambda, where λ𝕏(H)\lambda\in\mathbb{X}^{*}(H) is the weight of the equivariant slice ZBunG,rigZ\to\mathrm{Bun}_{G,rig}.

Remark 1.0.7.

The description of the singularities in types AA, DD and EE was given without proof in [helmke-slodowy04]. As far as we know, the description for types BB, CC, FF and GG is new.

Remark 1.0.8.

It follows from the explicit degrees and weights given in Theorems 4.1.3 and 4.2.9 and in Table 4 that the deformations of types A1A_{1}, CC, FF and GG are related to those of types DD and EE by a curious twist on the usual folding story for du Val singularities. For each pair (A1,E5)(A_{1},E_{5}), (Cl,Dl+4)(C_{l},D_{l+4}), (Fl,El+3)(F_{l},E_{l+3}) and (G2,E8)(G_{2},E_{8}), the surfaces χZ1(0)\chi_{Z}^{-1}(0) are isomorphic in both cases, and the deformation for the first case is naturally identified with the subspace preserving the action of μd𝔾m\mu_{d}\subseteq\mathbb{G}_{m} inside the deformation for the second, where d=2d=2 or 33. Note that this links different pairs of groups to the usual folding, i.e., the du Val singularities are not the same in these cases.

1.1. Plan of the paper

The paper consists of 4 sections, including this introduction.

The main purpose of §2 is to prove Theorem 1.0.2. We lay the groundwork in §2.1 by reviewing Helmke and Slodowy’s classification of subregular unstable bundles (Theorem 2.1.2). In §2.2, we review the theory of parabolic induction for equivariant slices, and use it to reduce Theorem 1.0.2 to a statement about existence of slices for Levi subgroups of GG (Theorem 2.2.6). We prove this theorem in §2.4 using a detailed study of the structure of the relevant Levis in §2.3.

In §3, we prove Theorem 1.0.3. The theorem is broken into four parts, Proposition 3.1.1, 3.4.1, 3.5.1 and 3.6.1, concerning the decomposition of χ~Z1(0ΘY1)\tilde{\chi}_{Z}^{-1}(0_{\Theta_{Y}^{-1}}) into irreducible components and the detailed structure of each of the three components respectively, which are proved in subsections 3.1, 3.4, 3.5 and 3.6. This section also features a brief review of the construction of “Bruhat cells” for principal bundles in §3.2 and an important auxiliary calculation of certain Bruhat cells G=GLnG=GL_{n} in §3.3.

In §4, we give the application to the identification of the singular surfaces χZ1(0)\chi_{Z}^{-1}(0) and their deformations. We give the identification of the surfaces (Theorem 4.1.3) in §4.1. In §4.2, we briefly discuss deformation theory with weights, and prove (Theorem 4.2.9) that the deformations χZ:ZY^//W\chi_{Z}\colon Z\to\widehat{Y}{/\mkern-6.0mu/}W have the miniversality properties asserted in Theorem 1.0.6.

1.2. Notation and conventions

Our notations and conventions are all consistent with [davis19].

Unless otherwise specified, by a reductive group we will mean a split connected reductive group scheme over Spec\operatorname{\mathrm{Spec}}\mathbb{Z}.

Throughout the paper, we will fix a connected regular stack SS, a smooth elliptic curve ESE\to S with origin OE:SEO_{E}\colon S\to E, and a simply connected simple reductive group GG (over Spec\operatorname{\mathrm{Spec}}\mathbb{Z}) with maximal torus and Borel subgroup TBGT\subseteq B\subseteq G.

We will write (𝕏(T),Φ,𝕏(T),Φ)(\mathbb{X}^{*}(T),\Phi,\mathbb{X}_{*}(T),\Phi^{\vee}) for the root datum of GG, where

𝕏(T)=Hom(T,𝔾m)and𝕏(T)=Hom(𝔾m,T)\mathbb{X}^{*}(T)=\mathrm{Hom}(T,\mathbb{G}_{m})\quad\text{and}\quad\mathbb{X}_{*}(T)=\mathrm{Hom}(\mathbb{G}_{m},T)

are the groups of characters and cocharacters of the split torus TT. The set of roots Φ\Phi is by definition the set of weights of TT acting on the Lie algebra 𝔤=Lie(G)\mathfrak{g}=\mathrm{Lie}(G); we will adopt the convention that the set ΦΦ\Phi_{-}\subseteq\Phi of negative roots is the set of nonzero weights of TT acting on Lie(B)\mathrm{Lie}(B), and let Φ+=Φ\Phi_{+}=-\Phi_{-} be the corresponding set of positive roots. Note that this convention means that for λ𝕏(T)\lambda\in\mathbb{X}^{*}(T), the line bundle λ=G×Bλ\mathcal{L}_{\lambda}=G\times^{B}\mathbb{Z}_{\lambda} on the flag variety G/BG/B is nef if and only if λ\lambda is dominant (i.e., λ,α0\langle\lambda,\alpha^{\vee}\rangle\geq 0 for all αΦ+\alpha^{\vee}\in\Phi_{+}^{\vee}). We will write Δ={α1,,αl}Φ+\Delta=\{\alpha_{1},\ldots,\alpha_{l}\}\subseteq\Phi_{+} and Δ={α1,,αl}Φ+\Delta^{\vee}=\{\alpha_{1}^{\vee},\ldots,\alpha_{l}^{\vee}\}\subseteq\Phi_{+}^{\vee} for the sets of positive simple roots and coroots respectively, and {ϖ1,,ϖl}\{\varpi_{1},\ldots,\varpi_{l}\} and {ϖ1,,ϖl}\{\varpi_{1}^{\vee},\ldots,\varpi_{l}^{\vee}\} for the bases of (Φ)(\mathbb{Z}\Phi^{\vee})^{\vee} and (Φ)(\mathbb{Z}\Phi)^{\vee} dual to Δ\Delta and Δ\Delta^{\vee} respectively. Note that Φ=𝕏(T)\mathbb{Z}\Phi^{\vee}=\mathbb{X}_{*}(T) since GG is simply connected, so {α1,,αl}\{\alpha_{1}^{\vee},\ldots,\alpha_{l}^{\vee}\} is a basis for 𝕏(T)\mathbb{X}_{*}(T) and {ϖ1,,ϖl}\{\varpi_{1},\ldots,\varpi_{l}\} is a basis for 𝕏(T)\mathbb{X}^{*}(T). We will also write WNG(T)/TW\cong N_{G}(T)/T for the Weyl group of GG generated by the reflections siWs_{i}\in W in the simple roots αiΔ\alpha_{i}\in\Delta.

We will also use the notation

𝕏(T)+={λ𝕏(T)ϖi,λ0for allαiΔ}=0Φ+\mathbb{X}_{*}(T)_{+}=\{\lambda\in\mathbb{X}_{*}(T)\mid\langle\varpi_{i},\lambda\rangle\geq 0\;\text{for all}\;\alpha_{i}\in\Delta\}=\mathbb{Z}_{\geq 0}\Phi_{+}^{\vee}

and set 𝕏(T)=𝕏(T)+\mathbb{X}_{*}(T)_{-}=-\mathbb{X}_{*}(T)_{+}. We have a related partial ordering on 𝕏(T)\mathbb{X}_{*}(T) defined by λμ\lambda\leq\mu if μλ𝕏(T)+\mu-\lambda\in\mathbb{X}_{*}(T)+. Similarly, for any reductive group and coweights λ\lambda and μ\mu, we defined λμ\lambda\leq\mu if μλ\mu-\lambda is an integer linear combination of positive coroots with nonnegative coefficients.

If PGP\subseteq G is a parabolic subgroup, we will say that PP is standard if BPB\subseteq P, and that a Levi factor LPL\subseteq P is standard if TLT\subseteq L. Every parabolic subgroup is conjugate to a unique standard one, and every standard parabolic has a unique standard Levi. If PP is standard, the type of PP is the set

t(P)={αiΔαiis not a root ofP}Δ.t(P)=\{\alpha_{i}\in\Delta\mid\alpha_{i}\;\text{is not a root of}\;P\}\subseteq\Delta.

More generally, one defines the type of a parabolic subgroup for any reductive group with a choice of Borel as a subset of the positive simple roots in the same way. The construction Pt(P)P\mapsto t(P) defines a bijection between (proper) parabolic subgroups of GG and (nonempty) subsets of Δ\Delta. For any parabolic subgroup PP, we will often write TP=P/[P,P]T_{P}=P/[P,P] for the torus with character group 𝕏(TP)=𝕏(P)\mathbb{X}^{*}(T_{P})=\mathbb{X}^{*}(P).

We also fix the following notation for the root datum and parabolic subgroups of GLnGL_{n}. Define parabolic subgroups

Qkn={(ap,q)1p,qnGLnap,q=0forp<min(q,k)}Q^{n}_{k}=\{(a_{p,q})_{1\leq p,q\leq n}\in GL_{n}\mid a_{p,q}=0\;\text{for}\;p<\mathrm{min}(q,k)\}

for 1kn1\leq k\leq n. Note that QnnGLnQ^{n}_{n}\subseteq GL_{n} is the Borel subgroup of lower triangular matrices, so TQnn:=Qnn/[Qnn,Qnn]T_{Q^{n}_{n}}:=Q^{n}_{n}/[Q^{n}_{n},Q^{n}_{n}] is naturally identified with the maximal torus of diagonal matrices in GLnGL_{n}. We will write e1,,en𝕏(TQnn)e_{1},\ldots,e_{n}\in\mathbb{X}^{*}(T_{Q^{n}_{n}}) for the basis given by ei(aj,k)=ai,ie_{i}(a_{j,k})=a_{i,i}, and e1,,en𝕏(TQnn)e_{1}^{*},\ldots,e_{n}^{*}\in\mathbb{X}_{*}(T_{Q^{n}_{n}}) for the dual basis. We label the simple roots of GLnGL_{n} as βi=eiei+1\beta_{i}=e_{i}-e_{i+1} for 1i<n1\leq i<n, so QknGLnQ^{n}_{k}\subseteq GL_{n} is the standard parabolic subgroup of type {β1,,βk1}\{\beta_{1},\ldots,\beta_{k-1}\}.

For any group scheme HH over Spec\operatorname{\mathrm{Spec}}\mathbb{Z}, we will write BunH\mathrm{Bun}_{H} for the relative stack of HH-bundles on EE over SS. If the quotient H/Ru(H)H/R_{u}(H) of HH by its unipotent radical Ru(H)R_{u}(H) is split reductive and ξHX\xi_{H}\to X is an HH-bundle on a curve XX, then we write degξH𝕏(H/Ru(H)[H,H])\deg\xi_{H}\in\mathbb{X}_{*}(H/R_{u}(H)[H,H]) and μ(ξH)𝕏(Z(H/Ru(H)))\mu(\xi_{H})\in\mathbb{X}_{*}(Z(H/R_{u}(H))^{\circ})_{\mathbb{Q}} for degree and slope of ξH\xi_{H} in the sense of [davis19, §1.2]. Note that these are related by

λ,deg(ξH)=λ,μ(ξH)\langle\lambda,\deg(\xi_{H})\rangle=\langle\lambda,\mu(\xi_{H})\rangle

for all λ𝕏(H)\lambda\in\mathbb{X}^{*}(H) under the obvious pairings, so in fact there is a canonical bijection between degrees and slopes. We write BunHdBunH\mathrm{Bun}_{H}^{d}\subseteq\mathrm{Bun}_{H} and BunHμ\mathrm{Bun}_{H}^{\mu} for the open and closed substacks of HH-bundles of degree dd and slope μ\mu respectively. A superscript ()ss(-)^{ss} denotes the open substack of semistable bundles.

For any split torus TT^{\prime} and λ𝕏(T)\lambda\in\mathbb{X}^{*}(T^{\prime}), we write YTλY_{T^{\prime}}^{\lambda} for the coarse moduli space of BunTλ\mathrm{Bun}_{T^{\prime}}^{\lambda} over SS. This can also be described as the quotient by the natural 𝔹T\mathbb{B}T^{\prime}-action, and the fine moduli space of TT^{\prime}-bundles of degree λ\lambda on EE trivialised at OEO_{E}. For the sake of brevity, we will drop the subscript ()T(-)_{T^{\prime}} when T=TT^{\prime}=T is the maximal torus of GG, and drop the superscript ()λ(-)^{\lambda} when λ=0\lambda=0. So, in particular, YY denotes the coarse moduli space of TT-bundles on EE of degree 0. We will also write YP=YTPY_{P}=Y_{T_{P}} when TP=P/[P,P]T_{P}=P/[P,P] for some parabolic subgroup PP of a reductive group.

For any reductive group HH and parabolic subgroup PHP\subseteq H with Levi subgroup LP/Ru(P)L\cong P/R_{u}(P), and a degree d𝕏(L/[L,L])d\in\mathbb{X}_{*}(L/[L,L]) (resp., slope d𝕏(Z(L))d\in\mathbb{X}_{*}(Z(L)^{\circ})_{\mathbb{Q}}) we will write KMP,Hd\mathrm{KM}_{P,H}^{d} for the Kontsevich-Mori compactification of BunPd\mathrm{Bun}_{P}^{d} over BunG\mathrm{Bun}_{G}. This is a smooth stack, proper over BunH\mathrm{Bun}_{H}, containing BunPd\mathrm{Bun}_{P}^{d} as a dense open substack whose complement is a divisor with normal crossings, such that the natural map BunPdYPd\mathrm{Bun}_{P}^{d}\to Y_{P}^{d} extends to KMP,HdYPd\mathrm{KM}_{P,H}^{d}\to Y_{P}^{d}. It parametrises tuples (ξH,C,σ)(\xi_{H},C,\sigma) where ξHE\xi_{H}\to E is an HH-bundle and σ:CξH/P\sigma\colon C\to\xi_{H}/P is a stable map from a nodal curve of genus 11 such that CEC\to E is degree 11 and degσ(ξH×Pλ)=λ,d\deg\sigma^{*}(\xi_{H}\times^{P}\mathbb{Z}_{\lambda})=\langle\lambda,d\rangle for all λ𝕏(P)\lambda\in\mathbb{X}^{*}(P). As in the introduction, we will write Bun~G=KMB,G0\widetilde{\mathrm{Bun}}_{G}=\mathrm{KM}_{B,G}^{0}. For a detailed discussion of these compactifications, see [davis19a, Chapter 3] or [campbell16].

If XX is any stack equipped with an injective action of the classifying stack 𝔹Z(G)\mathbb{B}Z(G) of the centre of GG, then we write XrigX_{rig} for the rigidification of XX with respect to Z(G)Z(G) obtained by taking the quotient of all automorphism groups in XX by Z(G)Z(G) [davis19, Definition 2.2.2]. For example, if HH is any group scheme with Z(G)Z(H)Z(G)\subseteq Z(H), then 𝔹Z(G)\mathbb{B}Z(G) acts injectively on BunH\mathrm{Bun}_{H}, so we have a rigidification BunH,rig\mathrm{Bun}_{H,rig}.

If XSX\to S is a morphism of Artin stacks, we will write 𝕃X/S\mathbb{L}_{X/S} for the relative cotangent complex [olsson07]*§8 and 𝕋X/S=(𝕃X/S)\mathbb{T}_{X/S}=(\mathbb{L}_{X/S})^{\vee} for the relative tangent complex.

If VV is a vector space or a vector bundle on a scheme, we adopt the convention that the projectivisation (V)\mathbb{P}(V) parametrises 11-dimensional subspaces or subbundles (rather than quotients).

1.3. Acknowledgements

The author would like to thank Ian Grojnowski, Travis Schedler, Nicholas Shepherd-Barron and Richard Thomas for many helpful conversations.

The majority of this research was done while the author was a PhD student at King’s College London, and completed at the University of Edinburgh. This work was supported by the Engineering and Physical Sciences Research Council grants [EP/L015234/1] (The EPSRC Centre for Doctoral Training in Geometry and Number Theory (The London School of Geometry and Number Theory), University College London), and [EP/R034826/1].

2. Subregular slices

The purpose of this section is to prove Theorem 1.0.2. We prepare the ground in §2.1, where we review the classification of subregular unstable bundles, and §2.2, where we review the parabolic induction construction for slices and use it to reduce Theorem 1.0.2 to a statement for Levi subgroups (Theorem 2.2.6). We give very explicit descriptions of all the relevant Levi subgroups in §2.3, and use these descriptions to give a case-by-case proof of Theorem 2.2.6 in §2.4.

2.1. Subregular unstable bundles

In this subsection, we review Helmke and Slodowy’s classification of subregular unstable bundles [helmke-slodowy01].

Definition 2.1.1.

Let s:SpeckSs\colon\operatorname{\mathrm{Spec}}k\to S be a geometric point and let ξGEs\xi_{G}\to E_{s} be an unstable GG-bundle. We say that ξG\xi_{G} is regular (resp., subregular) if dimAut(ξG)=l+2\dim\mathrm{Aut}(\xi_{G})=l+2 (resp. l+4l+4).

In the following theorem, if s:SpeckSs\colon\operatorname{\mathrm{Spec}}k\to S is a geometric point, LGL\subseteq G is a Levi subgroup, and ξL\xi_{L} is a semistable LL-bundle on EsE_{s} of slope μ𝕏(Z(L))\mu\in\mathbb{X}_{*}(Z(L)^{\circ})_{\mathbb{Q}}, then we say that ξL\xi_{L} is regular if its automorphism group has minimal dimension among all automorphism groups of semistable LL-bundles on EsE_{s} of slope μ\mu.

Theorem 2.1.2.

Let s:SpeckSs\colon\operatorname{\mathrm{Spec}}k\to S be a geometric point and let ξGEs\xi_{G}\to E_{s} be an unstable GG-bundle. Then either ξG\xi_{G} is regular and dimAut(ξG)=l+2\dim\mathrm{Aut}(\xi_{G})=l+2, or dimAut(ξG)l+4\dim\mathrm{Aut}(\xi_{G})\geq l+4. If ξG\xi_{G} has Harder-Narasimhan reduction ξP\xi_{P} to a standard parabolic PP with Levi factor LL, and associated LL-bundle ξL\xi_{L} of slope μ\mu, then ξG\xi_{G} is subregular if and only if ξL\xi_{L} is regular semistable and (G,P,μ)(G,P,\mu) satisfies one of the following conditions.

(Type A1A_{1}):

GG is of type A1A_{1}, t(P)={α1}t(P)=\{\alpha_{1}\} and ϖ1,μ=2\langle\varpi_{1},\mu\rangle=-2.

(Type AlA_{l}):

GG is of type AlA_{l} for l>1l>1, t(P)={αi,αi+1}t(P)=\{\alpha_{i},\alpha_{i+1}\} for some ii with 1i<l1\leq i<l, and ϖi,μ=ϖi+1,μ=1\langle\varpi_{i},\mu\rangle=\langle\varpi_{i+1},\mu\rangle=-1.

(Type BlB_{l}):

GG is of type BlB_{l} for l3l\geq 3, t(P)={αl2}t(P)=\{\alpha_{l-2}\} and ϖl2,μ=1\langle\varpi_{l-2},\mu\rangle=-1.

(Type ClC_{l}):

GG is of type ClC_{l} for l2l\geq 2, t(P)={αl1}t(P)=\{\alpha_{l-1}\} and ϖl1,μ=1\langle\varpi_{l-1},\mu\rangle=-1.

(Type DlD_{l}):

GG is of type DlD_{l} for l4l\geq 4, t(P)={αi}t(P)=\{\alpha_{i}\} and ϖi,μ=1\langle\varpi_{i},\mu\rangle=-1, where i{1,3,4}i\in\{1,3,4\} if l=4l=4 and i=l3i=l-3 otherwise.

(Type ElE_{l}):

GG is of type D5D_{5}, E6E_{6}, E7E_{7} or E8E_{8}, t(P)={αi}t(P)=\{\alpha_{i}\} and ϖi,μ=1\langle\varpi_{i},\mu\rangle=-1, where i{4,5}i\in\{4,5\} if GG is of type D5D_{5}, i{2,5}i\in\{2,5\} if GG is of type E6E_{6}, and i=5i=5 if GG is of type E7E_{7} or E8E_{8}.

(Type FlF_{l}):

GG is of type B3B_{3} or F4F_{4}, t(P)={α3}t(P)=\{\alpha_{3}\} and ϖ3,μ=1\langle\varpi_{3},\mu\rangle=-1.

(Type GlG_{l}):

GG is of type G2G_{2}, t(P)={α1}t(P)=\{\alpha_{1}\} and ϖ1,μ=1\langle\varpi_{1},\mu\rangle=-1.

Here the labelling of the Dynkin diagrams is as in Table 1.

Al:A_{l}: 1\scriptstyle{1}2\scriptstyle{2}3\scriptstyle{3}l1\scriptstyle{l-1}l\scriptstyle{l} El:E_{l}: 1\scriptstyle{1}2\scriptstyle{2}3\scriptstyle{3}4\scriptstyle{4}5\scriptstyle{5}l\scriptstyle{l}
Bl:B_{l}: 1\scriptstyle{1}2\scriptstyle{2}l2\scriptstyle{l-2}l1\scriptstyle{l-1}l\scriptstyle{l}>> F4:F_{4}: 1\scriptstyle{1}2\scriptstyle{2}3\scriptstyle{3}4\scriptstyle{4}>>
Cl:C_{l}: 1\scriptstyle{1}2\scriptstyle{2}l2\scriptstyle{l-2}l1\scriptstyle{l-1}l\scriptstyle{l}<< G2:G_{2}: 1\scriptstyle{1}2\scriptstyle{2}<<
Dl:D_{l}: 1\scriptstyle{1}2\scriptstyle{2}l3\scriptstyle{l-3}l2\scriptstyle{l-2}l\scriptstyle{l}l1\scriptstyle{l-1}
Table 1. Labelling of the Dynkin diagrams
Proof.

The theorem is a selection of statements from [helmke-slodowy01, Theorems 5.1 and 5.12], which are proved there when S=SpecS=\operatorname{\mathrm{Spec}}\mathbb{C}. To deduce the theorem in general, note that by specialisation (and [helmke-slodowy01, Proposition 2.4, (c)], whose proof works over any field) we have

dimAut(ξG)=2ρ,μ+dimAut(ξL)2ρ,μ+d(L,μ),\dim\mathrm{Aut}(\xi_{G})=-\langle 2\rho,\mu\rangle+\dim\mathrm{Aut}(\xi_{L})\geq-\langle 2\rho,\mu\rangle+d(L,\mu),

where d(L,μ)d(L,\mu) is the dimension of the automorphism group of a regular semistable LL-bundle with slope μ\mu on an elliptic curve over \mathbb{C}. So [davis19, Proposition 4.2.3] and the statement of the theorem over \mathbb{C} imply that there are no unstable bundles with dimAut(ξG)=l+3\dim\mathrm{Aut}(\xi_{G})=l+3 and that the Harder-Narasimhan reduction of any subregular unstable bundle must appear on the list above. A priori, there may be an elliptic curve EsE_{s} over a field of positive characteristic such that regular semistable LL-bundles ξL\xi_{L} on EsE_{s} of slope μ\mu have dimAut(ξL)>d(L,μ)\dim\mathrm{Aut}(\xi_{L})>d(L,\mu), and hence GG-bundles with Harder-Narasimhan reductions on the list above that are not subregular. However, in case (Type A1A_{1}): this cannot happen since L=TL=T, and the proof of Theorem 2.2.6 shows that this does not happen for the other Levis and slopes on the list (see Remark 2.2.7). So the theorem holds in all characteristics. ∎

Definition 2.1.3.

We will say that a tuple (G,P,μ)(G,P,\mu) consisting of a simply connected simple group GG, a standard parabolic PP with Levi factor LL, and a Harder-Narasimhan vector μ\mu for PP is a subregular Harder-Narasimhan class if ξL×LG\xi_{L}\times^{L}G is subregular unstable for ξL\xi_{L} a regular semistable LL-bundle of slope μ\mu. (Recall from [davis19, Definition 2.3.3] that μ\mu is a Harder-Narasimhan vector if, for every root αΦ\alpha\in\Phi of GG, α\alpha is a root of PP if and only if α,μ0\langle\alpha,\mu\rangle\geq 0. By definition, the slope of a Harder-Narasimhan reduction is always a Harder-Narasimhan vector.) We will say that (G,P,μ)(G,P,\mu) is of type A1A_{1} (resp., type AlA_{l}, type BlB_{l}, etc.) if it satisfies (Type A1A_{1}): (resp., (Type AlA_{l}): , (Type BlB_{l}): , etc.) of Theorem 2.1.2.

Remark 2.1.4.

We stress that the type of a subregular Harder-Narasimhan class (G,P,μ)(G,P,\mu) is often, but not always, the type of the group GG. For example, for GG of type B3B_{3}, there are subregular Harder-Narasimhan classes of types B3B_{3} and F3F_{3}, and for GG of type D5D_{5}, there are subregular Harder-Narasimhan classes of types D5D_{5} and E5E_{5}.

2.2. Slicing by parabolic induction

In this subsection, we explain how the proof of Theorem 1.0.2 can be reduced to the construction of well-behaved slices of BunL,rigss,μ\mathrm{Bun}_{L,rig}^{ss,\mu} for each subregular Harder-Narasimhan class. We first recall the definitions.

Definition 2.2.1.

Let LGL\subseteq G be a Levi subgroup. A slice of BunL,rig\mathrm{Bun}_{L,rig} is stack ZZ equipped with a map ZBunL,rigZ\to\mathrm{Bun}_{L,rig} such that the map ZBunL,rig/EZ\to\mathrm{Bun}_{L,rig}/E is smooth, where the quotient is taken with respect to the natural action of EE on BunL,rig\mathrm{Bun}_{L,rig} by translations. If HH is a torus and λ𝕏(H)\lambda\in\mathbb{X}^{*}(H) is an equivariant slice (of BunG,rig\mathrm{Bun}_{G,rig}) with equivariance group HH and weight λ\lambda is a stack ZZ equipped with an action of HH, a slice Z/HBunG,rigZ/H\to\mathrm{Bun}_{G,rig}, and an HH-equivariant lift ZY^//WZ\to\widehat{Y}{/\mkern-6.0mu/}W of the coarse quotient map ZBunG,rig(Y^//W)/𝔾mZ\to\mathrm{Bun}_{G,rig}\to(\widehat{Y}{/\mkern-6.0mu/}W)/\mathbb{G}_{m}, where HH acts on Y^//W\widehat{Y}{/\mkern-6.0mu/}W through the 𝔾m\mathbb{G}_{m}-action and the homomorphism λ:H𝔾m\lambda\colon H\to\mathbb{G}_{m}.

We also recall a few elements of the theory of parabolic induction for slices, the idea of which goes back to R. Friedman and J. Morgan [friedman-morgan00]. A more detailed exposition can be found in [davis19, §4.1] or [davis19a, §5.1-5.2].

Definition 2.2.2 ([davis19, Definition 4.1.1] or [davis19a, Definition 5.2.1]).

Let LLGL\subseteq L^{\prime}\subseteq G be Levi subgroups, let μ𝕏(Z(L))\mu\in\mathbb{X}_{*}(Z(L)^{\circ})_{\mathbb{Q}}, and let P+LP^{+}\subseteq L^{\prime} be the unique parabolic subgroup with Levi factor LL for which μ-\mu is a Harder-Narasimhan vector for P+P^{+} [davis19, Definition 2.3.3]. If Z0BunL,rigss,μZ_{0}\to\mathrm{Bun}_{L,rig}^{ss,\mu} is a slice, then the parabolic induction of Z0Z_{0} to LL^{\prime} is the slice

IndLL(Z0)=BunP+,rig×BunL,rigZ0BunP+,rigμBunL,rigμ,\mathrm{Ind}_{L}^{L^{\prime}}(Z_{0})=\mathrm{Bun}_{P^{+},rig}\times_{\mathrm{Bun}_{L,rig}}Z_{0}\longrightarrow\mathrm{Bun}_{P^{+},rig}^{\mu^{\prime}}\longrightarrow\mathrm{Bun}_{L^{\prime},rig}^{\mu^{\prime}},

where μ𝕏(Z(L))\mu^{\prime}\in\mathbb{X}_{*}(Z(L^{\prime})^{\circ})_{\mathbb{Q}} is the image of μ\mu.

In the following proposition, we write

Φμ={αΦLα,μ<0}\Phi_{\mu}=\{\alpha\in\Phi_{L^{\prime}}\mid\langle\alpha,\mu\rangle<0\}

for the set of roots in the unipotent radical Ru(P+)LR_{u}(P^{+})\subseteq L^{\prime}, and

2ρP+=αΦμα.2\rho_{P^{+}}=-\sum_{\alpha\in\Phi_{\mu}}\alpha.
Proposition 2.2.3.

In the situation of Definition 2.2.2, the natural morphism IndLL(Z0)Z0\mathrm{Ind}_{L}^{L^{\prime}}(Z_{0})\to Z_{0} is an affine space bundle, with fibres of dimension 2ρP+,μ\langle 2\rho_{P^{+}},\mu\rangle. Moreover, the torus Z(L)rig:=Z(L)/Z(G)Z(L)_{rig}:=Z(L)/Z(G) naturally acts on the fibres of this bundle with weights in Φμ-\Phi_{\mu}. The morphism IndLL(Z0)BunL,rig\mathrm{Ind}_{L}^{L^{\prime}}(Z_{0})\to\mathrm{Bun}_{L^{\prime},rig} is equivariant with respect to this action.

Proof.

This is an immediate consequence of [davis19, Propositions 4.1.6 and 4.1.8]. ∎

From now on, we will take L=GL^{\prime}=G.

Recall from [davis19, §3.2] that there is a unique positive generator ΘBunG,rigPic(BunG,rig)\Theta_{\mathrm{Bun}_{G,rig}}\in\mathrm{Pic}(\mathrm{Bun}_{G,rig}); this theta bundle is nothing but the inverse of the pullback of the universal line bundle under

BunG,rig(Y^//W)/𝔾m𝔹𝔾m.\mathrm{Bun}_{G,rig}\longrightarrow(\widehat{Y}{/\mkern-6.0mu/}W)/\mathbb{G}_{m}\longrightarrow\mathbb{B}\mathbb{G}_{m}.
Definition 2.2.4 ([davis19, Definition 4.1.15]).

Let LL and μ\mu be as in Definition 2.2.2. A Θ\Theta-trivial slice of BunL,rigss,μ\mathrm{Bun}_{L,rig}^{ss,\mu} is a slice Z0BunL,rigss,μZ_{0}\to\mathrm{Bun}_{L,rig}^{ss,\mu} equipped with a trivialisation of the pullback of ΘBunG,rig\Theta_{\mathrm{Bun}_{G,rig}} along

Z0BunL,rigss,μBunG,rig.Z_{0}\longrightarrow\mathrm{Bun}_{L,rig}^{ss,\mu}\longrightarrow\mathrm{Bun}_{G,rig}.

The point of Θ\Theta-trivial slices is that they naturally give equivariant slices after parabolic induction. In the following proposition, we write (|):𝕏(T)𝕏(T)(\,{\,|\,}\,)\colon\mathbb{X}_{*}(T)\otimes\mathbb{X}_{*}(T)\to\mathbb{Z} for the Killing form normalised so that (α|α)=2(\alpha^{\vee}{\,|\,}\alpha^{\vee})=2 for αΦ\alpha^{\vee}\in\Phi^{\vee} a short coroot.

Proposition 2.2.5 ([davis19, Proposition 4.1.12]).

Let Z0BunL,rigss,μZ_{0}\to\mathrm{Bun}_{L,rig}^{ss,\mu} be a Θ\Theta-trivial slice. Then IndLG(Z0)BunG,rig\mathrm{Ind}_{L}^{G}(Z_{0})\to\mathrm{Bun}_{G,rig} is naturally endowed with the structure of an equivariant slice with equivariance group Z(L)rigZ(L)_{rig} and weight (μ|)(\mu{\,|\,}-).

The parabolic induction construction allows us to deduce Theorem 1.0.2 from the following statement.

Theorem 2.2.6.

Let (G,P,μ)(G,P,\mu) be a subregular Harder-Narasimhan class not of type A1A_{1}, and let d{1,2,3}d\in\{1,2,3\} be as in Theorem 1.0.3. Then there is a μd\mu_{d}-gerbe 𝔊uni\mathfrak{G}^{uni} on the stack M1,1M_{1,1} of elliptic curves such that if the pullback 𝔊\mathfrak{G} of 𝔊uni\mathfrak{G}^{uni} to SS is trivial then there exists a Θ\Theta-trivial slice Z0BunL,rigss,μZ_{0}\to\mathrm{Bun}_{L,rig}^{ss,\mu} with the following properties.

  1. (1)

    The morphism Z0SZ_{0}\to S is smooth and proper with finite and generically trivial relative stabilisers.

  2. (2)

    The morphism Z0BunL,rigss,μ/EZ_{0}\to\mathrm{Bun}_{L,rig}^{ss,\mu}/E is smooth with connected fibres.

  3. (3)

    The image of Z0BunL,rigss,μ/EZ_{0}\to\mathrm{Bun}_{L,rig}^{ss,\mu}/E is equal to the locus of regular semistable bundles.

  4. (4)

    The induced equivariant slice Z=IndLG(Z0)BunG,rigZ=\mathrm{Ind}_{L}^{G}(Z_{0})\to\mathrm{Bun}_{G,rig} has relative dimension l+3l+3 over SS.

We will prove Theorem 2.2.6 in §2.4 by writing down explicit slices in each case of Theorem 2.1.2. Although a classification-free proof is probably possible, the explicit slices are also useful in the proof of Theorem 1.0.3.

Remark 2.2.7.

The proof will show that Theorem 2.2.6 holds for every tuple (G,P,μ)(G,P,\mu) on the list of Theorem 2.1.2, excluding (Type A1A_{1}): . In the notation of the proof of Theorem 2.1.2, this shows that in each case we have a slice Z0BunL,rigss,μZ_{0}\to\mathrm{Bun}_{L,rig}^{ss,\mu} with relative dimension l+3+2ρ,μ=d(L,μ)1l+3+\langle 2\rho,\mu\rangle=d(L,\mu)-1 over SS, and hence relative dimension d(L,μ)d(L,\mu) over BunL,rigss,μ/E\mathrm{Bun}_{L,rig}^{ss,\mu}/E. Since Z0SZ_{0}\to S has finite relative stabilisers, this shows that dimAut(ξL)d(L,μ)\dim\mathrm{Aut}(\xi_{L})\leq d(L,\mu) for a regular semistable LL-bundle in all characteristics.

Corollary 2.2.8.

Theorem 1.0.2 is true (with S=SpeckS=\operatorname{\mathrm{Spec}}k for kk an algebraically closed field as in the introduction).

Proof.

Let (G,P,μ)(G,P,\mu) be the subregular Harder-Narasimhan class of ξG\xi_{G}. Since S=SpeckS=\operatorname{\mathrm{Spec}}k, the μd\mu_{d}-gerbe 𝔊\mathfrak{G} on SS of Theorem 2.2.6 is necessarily trivial, so there exists a Θ\Theta-trivial slice Z0BunL,rigss,μZ_{0}\to\mathrm{Bun}_{L,rig}^{ss,\mu} satisfying conditions (1)–(4). We let Z=IndLG(Z0)BunG,rigZ=\mathrm{Ind}_{L}^{G}(Z_{0})\to\mathrm{Bun}_{G,rig} be the parabolic induction of Z0Z_{0} to GG, endowed with the equivariant slice structure of Proposition 2.2.5. Note that the equivariance group H=Z(L)rigH=Z(L)_{rig} is isomorphic to 𝔾m×𝔾m\mathbb{G}_{m}\times\mathbb{G}_{m} for (G,P,μ)(G,P,\mu) of type AA and 𝔾m\mathbb{G}_{m} otherwise, as required for the statement of Theorem 1.0.2.

Condition (1) of Theorem 1.0.2 follows immediately from Proposition 2.2.3 and (1) of Theorem 2.2.6 (note that Z0=IndLG(Z0)Z(L)rigZ_{0}=\mathrm{Ind}_{L}^{G}(Z_{0})^{Z(L)_{rig}}). Condition (3) of Theorem 1.0.2 follows from (2) of Theorem 2.2.6.

To prove that Theorem 1.0.2 (2) is satisfied, first note that for any zZZ0z\in Z\setminus Z_{0}, comparing the codimensions of Z(L)rigzZ(L)_{rig}\cdot z in ZZ and the corresponding GG-bundle ξG,z\xi_{G,z} in BunG,rig/E\mathrm{Bun}_{G,rig}/E shows that dimAut(ξG,z)l+3\dim\mathrm{Aut}(\xi_{G,z})\leq l+3, so ξG,z\xi_{G,z} is not subregular. Moreover, we claim that the smooth morphism Z0BunL,rigss,reg,μ/EZ_{0}\to\mathrm{Bun}_{L,rig}^{ss,reg,\mu}/E is a bijection on KK-points for any algebraically closed field KK, from which it follows that it is an isomorphism on coarse moduli spaces. This proves (2), modulo the claim.

To prove the claim, note that Proposition 2.2.3 shows that the dimension of the fibre (Z0)x(Z_{0})_{x} of Z0BunL,rigss,μ/EZ_{0}\to\mathrm{Bun}_{L,rig}^{ss,\mu}/E over the image xx of ξL=ξP×PL\xi_{L}=\xi_{P}\times^{P}L is given by

l+42ρP+,μ=l+4+2ρ,μ.l+4-\langle 2\rho_{P^{+}},\mu\rangle=l+4+\langle 2\rho,\mu\rangle.

But from Remark 2.2.7 and the proof of Theorem 2.1.2, this is equal to the dimension d(L,μ)d(L,\mu) of Aut(ξL)\mathrm{Aut}(\xi_{L}). So (Z0)x/Aut(x)(Z0)sZs(Z_{0})_{x}/\mathrm{Aut}(x)\subseteq(Z_{0})_{s}\subseteq Z_{s} is a closed connected substack of dimension 0, where ss is the image of xx in SS, and hence has a single point over any algebraically closed field since (Z0)s(Z_{0})_{s} has finite stabilisers. ∎

Remark 2.2.9.

From the proof, we can also read off the weights of the equivariant slices in Theorem 1.0.2: abstractly, they are the characters (μ|)𝕏(Z(L)rig)(\mu{\,|\,}-)\in\mathbb{X}^{*}(Z(L)_{rig}) by Proposition 2.2.5. More explicitly, if we identify Z(L)rigZ(L)_{rig} with 𝔾m\mathbb{G}_{m} via the cocharacter ϖi𝕏(Z(L)rig)-\varpi_{i}^{\vee}\in\mathbb{X}_{*}(Z(L)_{rig}) where t(P)={αi}t(P)=\{\alpha_{i}\} (resp., with 𝔾m×𝔾m\mathbb{G}_{m}\times\mathbb{G}_{m} via (ϖi,ϖi+1)(-\varpi_{i}^{\vee},-\varpi_{i+1}^{\vee}) in type AA), then the weight is identified with (1,1)(1,1) in type AA and with d{1,2,3}d\in\{1,2,3\} in the other types.

Remark 2.2.10.

We have deliberately excluded the subregular Harder-Narasimhan class of type A1A_{1} from Theorem 2.2.6. In this case, we have L=T𝔾mL=T\cong\mathbb{G}_{m} and BunL,rigss,μ=Bun𝔾m,rig2\mathrm{Bun}_{L,rig}^{ss,\mu}=\mathrm{Bun}_{\mathbb{G}_{m},rig}^{-2}, and one can try to construct the desired slice Z0=SBunLss,μZ_{0}=S\to\mathrm{Bun}_{L}^{ss,\mu} by lifting the natural section 𝒪(2OE):Z0=SPicS2(E)\mathcal{O}(-2O_{E})\colon Z_{0}=S\to\mathrm{Pic}^{-2}_{S}(E). It follows from [davis19, Proposition 4.1.15] that the fibre of the map

Bun𝔾m,rig2=PicS2(E)×𝔹𝔾m𝔹𝔾m\mathrm{Bun}_{\mathbb{G}_{m},rig}^{-2}=\mathrm{Pic}_{S}^{-2}(E)\times\mathbb{B}\mathbb{G}_{m}\longrightarrow\mathbb{B}\mathbb{G}_{m}

classifying the pullback of the theta bundle is a μ2\mu_{2}-gerbe on PicS2(E)\mathrm{Pic}_{S}^{-2}(E), which is trivial if and only if Z0PicS2(E)Z_{0}\to\mathrm{Pic}^{-2}_{S}(E) lifts to a Θ\Theta-trivial map Z0BunL,rigss,μZ_{0}\to\mathrm{Bun}_{L,rig}^{ss,\mu}. This map will be a slice as long as 22 is invertible in 𝒪S\mathcal{O}_{S} (so that the stabiliser E[2]E[2] of a point in PicS2(E)\mathrm{Pic}^{-2}_{S}(E) is smooth). This slice satisfies (1), (3) and (4), but the map Z0BunL,rigss,2/EZ_{0}\to\mathrm{Bun}_{L,rig}^{ss,-2}/E is a torsor under an extension of E[2]E[2] by 𝔾m\mathbb{G}_{m} and hence has disconnected fibres. See also, however, Proposition 4.1.9.

2.3. The structure of Levi subgroups

In this subsection, we explicitly describe the Levi subgroups LPL\subseteq P for each subregular Harder-Narasimhan class, i.e. for each (G,P,μ)(G,P,\mu) on the list of Theorem 2.1.2.

We begin with a general description of Levi subgroups LGL\subseteq G whose Dynkin diagrams are of type AA. Suppose that LL is the Levi subgroup of a standard parabolic of type tΔt\subseteq\Delta. Then the Dynkin diagram of LL is obtained from the Dynkin diagram of GG by deleting the nodes labelled by elements of tt. We will assume that the Dynkin diagram of LL is a union of connected components of type AA.

The reductive group LL can be described directly in terms of the following data. First, write π0=π0(Δt)\pi_{0}=\pi_{0}(\Delta\setminus t) for the set of connected components of the Dynkin diagram of LL. For each component cπ0c\in\pi_{0}, write ncn_{c} for the number of nodes in cc, and choose a labelling αc,1,,αc,nc\alpha_{c,1},\ldots,\alpha_{c,n_{c}} of the nodes of cc so that αc,i\alpha_{c,i} is adjacent to αc,i+1\alpha_{c,i+1} for 1inc11\leq i\leq n_{c}-1. For each αkt\alpha_{k}\in t adjacent to a node of cc, let αc,ic,k\alpha_{c,i_{c,k}} be the unique node adjacent to αk\alpha_{k}, and for each αkt\alpha_{k}\in t not adjacent to any node of cc, set ic,k=nc+1i_{c,k}=n_{c}+1. Finally, write

mc,k=i=1ncαc,i,αk={αc,ic,k,αk,ific,knc,0,ific,k=nc+1,m_{c,k}=-\sum_{i=1}^{n_{c}}\langle\alpha_{c,i},\alpha_{k}^{\vee}\rangle=\begin{cases}-\langle\alpha_{c,i_{c,k}},\alpha_{k}^{\vee}\rangle,&\text{if}\;\;i_{c,k}\leq n_{c},\\ 0,&\text{if}\;\;i_{c,k}=n_{c}+1,\end{cases}

for cπ0c\in\pi_{0} and αkt\alpha_{k}\in t.

Proposition 2.3.1.

Assume we are in the setup above. Then there is an isomorphism

L{((Ac)cπ0,(λk)αkt)cπ0GLnc+1×αkt𝔾m|detAc=αktλkmc,k(nc+1ic,k)}L\overset{\sim}{\longrightarrow}\left\{((A_{c})_{c\in\pi_{0}},(\lambda_{k})_{\alpha_{k}\in t})\in\prod_{c\in\pi_{0}}GL_{n_{c}+1}\times\prod_{\alpha_{k}\in t}\mathbb{G}_{m}\,\left|\,\det A_{c}=\prod_{\alpha_{k}\in t}\lambda_{k}^{m_{c,k}(n_{c}+1-i_{c,k})}\right.\right\} (2.3.1)

with the property that for each αkt\alpha_{k}\in t, the character ϖk\varpi_{k} of LL is given by (2.3.1) composed with the projection ((Ac)cπ0,(λj)αjt)λk((A_{c})_{c\in\pi_{0}},(\lambda_{j})_{\alpha_{j}\in t})\mapsto\lambda_{k}.

Proof.

Since both sides of (2.3.1) are split reductive groups over Spec\operatorname{\mathrm{Spec}}\mathbb{Z}, it is enough to specify an isomorphism between their root data.

The root datum (M0,Ψ0,M0,Ψ0)(M_{0},\Psi_{0},M_{0}^{\vee},\Psi_{0}^{\vee}) of cπ0GLnc+1×αkt𝔾m\prod_{c\in\pi_{0}}GL_{n_{c}+1}\times\prod_{\alpha_{k}\in t}\mathbb{G}_{m} is specified as follows. The weight lattice is

M0=cπ0nc+1αktωk.M_{0}=\bigoplus_{c\in\pi_{0}}\mathbb{Z}^{n_{c}+1}\oplus\bigoplus_{\alpha_{k}\in t}\mathbb{Z}\omega_{k}.

The roots and coroots Ψ0\Psi_{0} and Ψ0\Psi_{0}^{\vee} are determined by requiring that

{βc,j=ec,jec,j+1cπ0and 1jnc}M0\{\beta_{c,j}=e_{c,j}-e_{c,j+1}\mid c\in\pi_{0}\;\text{and}\;1\leq j\leq n_{c}\}\subseteq M_{0}

be a set of positive simple roots for Ψ0\Psi_{0}, and that

βc,j=ec,jec,j+1\beta_{c,j}^{\vee}=e_{c,j}^{*}-e_{c,j+1}^{*}

where {ec,j1jnc+1}\{e_{c,j}\mid 1\leq j\leq n_{c}+1\} is the standard basis for nc+1\mathbb{Z}^{n_{c}+1}, and ec,jM0e_{c,j}^{*}\in M_{0}^{\vee} satisfies

ec,j,ec,j={1,if(c,j)=(c,j),0,otherwise,andωk,ec,j=0.\langle e_{c^{\prime},j^{\prime}},e_{c,j}^{*}\rangle=\begin{cases}1,&\text{if}\;\;(c^{\prime},j^{\prime})=(c,j),\\ 0,&\text{otherwise},\end{cases}\quad\text{and}\quad\langle\omega_{k},e_{c,j}^{*}\rangle=0.

The root datum (M,Ψ,M,Ψ)(M,\Psi,M^{\vee},\Psi^{\vee}) is given by setting

M=M0-span{j=1nc+1ec,jαktmc,k(nc+1ic,k)ωk|cπ0},M=\frac{M_{0}}{\mathbb{Z}\textnormal{-span}\left\{\left.\sum_{j=1}^{n_{c}+1}e_{c,j}-\sum_{\alpha_{k}\in t}m_{c,k}(n_{c}+1-i_{c,k})\omega_{k}\,\right|\,c\in\pi_{0}\right\}},

setting Ψ\Psi to be the image of Ψ0\Psi_{0} in MM, and setting ΨM\Psi^{\vee}\subseteq M^{\vee} to be the preimage of Ψ0\Psi_{0}^{\vee} under the injection MM0M^{\vee}\hookrightarrow M_{0}^{\vee}. Note that MM is indeed a lattice, so this is the root datum of a connected reductive group.

We define an isomorphism of (M,Ψ,M,Ψ)(M,\Psi,M^{\vee},\Psi^{\vee}) with the root datum (𝕏(T),Φt,𝕏(T),Φt)(\mathbb{X}^{*}(T),\Phi_{t},\mathbb{X}_{*}(T),\Phi_{t}^{\vee}) of LL via the isomorphism

ϕ:𝕏(T)\displaystyle\phi\colon\mathbb{X}_{*}(T) M\displaystyle\overset{\sim}{\longrightarrow}M^{\vee}
αc,j\displaystyle\alpha_{c,j}^{\vee} ec,jec,j+1\displaystyle\longmapsto e_{c,j}^{*}-e_{c,j+1}^{*}
αk\displaystyle\alpha_{k}^{\vee} ωk+cπ0j=ic,k+1nc+1mc,kec,j,\displaystyle\longmapsto\omega_{k}^{*}+\sum_{c\in\pi_{0}}\sum_{j=i_{c,k}+1}^{n_{c}+1}m_{c,k}e_{c,j}^{*},

for cπ0c\in\pi_{0}, 1jnc1\leq j\leq n_{c} and αkt\alpha_{k}\in t, where ωkM0\omega_{k}^{*}\in M_{0}^{\vee} satisfies ec,j,ωk=0\langle e_{c,j},\omega_{k}^{*}\rangle=0 and ωk,ωk=δk,k\langle\omega_{k^{\prime}},\omega_{k}^{*}\rangle=\delta_{k,k^{\prime}}. It is clear by inspection that ϕ\phi is a well-defined homomorphism of free abelian groups such that the dual is surjective. Since MM^{\vee} and 𝕏(T)\mathbb{X}_{*}(T) have the same rank, ϕ\phi is therefore an isomorphism. To prove that ϕ\phi defines an isomorphism of root data, it is enough to show that ϕ:𝕏(T)M\phi\colon\mathbb{X}_{*}(T)\to M^{\vee} sends αc,j\alpha_{c,j}^{\vee} to βc,j\beta_{c,j}^{\vee} and that ϕ:M𝕏(T)\phi^{*}\colon M\to\mathbb{X}^{*}(T) sends βc,j\beta_{c,j} to αc,j\alpha_{c,j} for all cπ0c\in\pi_{0} and 1jnc1\leq j\leq n_{c}. This is easily checked by direct calculation, so we are done. ∎

Now fix a subregular Harder-Narasimhan class (G,P,μ)(G,P,\mu) not of type A1A_{1}. It will be convenient to decompose the Dynkin diagram of GG as follows.

Notation 2.3.2.

If (G,P,μ)(G,P,\mu) is of type AA, then let {αi,αj}={αi,αi+1}=t(P)\{\alpha_{i},\alpha_{j}\}=\{\alpha_{i},\alpha_{i+1}\}=t(P). Otherwise, we let {αi}=t(P)\{\alpha_{i}\}=t(P) and let αjΔ\alpha_{j}\in\Delta be the unique special root. (Recall [friedman-morgan00, Definition 3.1.1] [davis19, Definition 4.2.1] that a simple root αΔ\alpha\in\Delta is special if it is a long root such that the Dynkin diagram Δ{α}\Delta\setminus\{\alpha\} is a union of components of type AA each meeting α\alpha at a single end.) Theorem 2.1.2 shows that in each case, αi\alpha_{i} is adjacent to αj\alpha_{j}. Deleting the edge joining αi\alpha_{i} and αj\alpha_{j} breaks the Dynkin diagram of GG into two connected components; we write c0c_{0} (resp., c1c_{1}) for the component containing αi\alpha_{i} (resp., αj\alpha_{j}) and n0n_{0} (resp., n1n_{1}) for the number of vertices in c0c_{0} (resp., c1c_{1}). Since αj\alpha_{j} is special, the Dynkin diagram of c0c_{0} is of type An0A_{n_{0}}. We write {αc0,1,,αc0,n0}Δ\{\alpha_{c_{0},1},\ldots,\alpha_{c_{0},n_{0}}\}\subseteq\Delta for the vertices of c0c_{0}, labelled so that αc0,k\alpha_{c_{0},k} is adjacent to αc0,k+1\alpha_{c_{0},k+1} for all k<n0k<n_{0} and αc0,n0=αi\alpha_{c_{0},n_{0}}=\alpha_{i}. For c{c0,c1}c\in\{c_{0},c_{1}\} and αc,k\alpha_{c,k} a root of cc, we also write ϖc,k𝕏(T)\varpi_{c,k}\in\mathbb{X}^{*}(T) for the corresponding fundamental dominant weight.

Our descriptions of the Levi subgroup LPL\subseteq P fall into four distinct cases.

Case 1: (G,P,μ)(G,P,\mu) is of type AA. In this case, we have the following elementary description of the Levi LL.

Lemma 2.3.3.

In the setup above, there is an isomorphism

LGLi×GLli=GLn0×GLn1L\cong GL_{i}\times GL_{l-i}=GL_{n_{0}}\times GL_{n_{1}}

so that the characters ϖi\varpi_{i} and ϖi+1\varpi_{i+1} are identified with the determinants of the first and second factors respectively.

Proof.

The desired isomorphism is given by applying Proposition 2.3.1 with an appropriate labelling. Explicitly, it is given by

GLi×GLli\displaystyle GL_{i}\times GL_{l-i} LSLl+1\displaystyle\overset{\sim}{\longrightarrow}L\subseteq{SL_{l+1}}
(A,B)\displaystyle(A,B) (A000(detA)1detB000v(Bt)1v1),\displaystyle\longmapsto\left(\begin{matrix}A&0&0\\ 0&(\det A)^{-1}\det B&0\\ 0&0&v(B^{t})^{-1}v^{-1}\end{matrix}\right),

where vSliv\in S_{l-i} is the matrix of the permutation of {1,,li}\{1,\ldots,l-i\} sending jj to lij+1l-i-j+1. ∎

Case 2: (G,P,μ)(G,P,\mu) is of type CC, DD, EE or FF. In this case, the connected component c1c_{1} containing αj\alpha_{j} of the Dynkin diagram with the edge joining αi\alpha_{i} and αj\alpha_{j} deleted is of type An1A_{n_{1}}, and we can choose a labelling αc1,1,,αc1,n1\alpha_{c_{1},1},\ldots,\alpha_{c_{1},n_{1}} such that αc1,p\alpha_{c_{1},p} is adjacent to αc1,p+1\alpha_{c_{1},p+1} for each pp and αj\alpha_{j} is either αc1,n1\alpha_{c_{1},n_{1}} (in types CC and FF) or αc1,n11\alpha_{c_{1},n_{1}-1} (in types DD and EE). We have the following description of LL.

Lemma 2.3.4.

In the setup above, there is an isomorphism

L{(A,B)GLn0×GLn1+1detB=(detA)2},L\cong\{(A,B)\in GL_{n_{0}}\times GL_{n_{1}+1}\mid\det B=(\det A)^{2}\},

such that ϖi\varpi_{i} is identified with the character (A,B)detA(A,B)\mapsto\det A and LBL\cap B is the preimage of the Borel subgroup Qn0n0×Qn1+1n1+1GLn0×GLn1+1Q_{n_{0}}^{n_{0}}\times Q_{n_{1}+1}^{n_{1}+1}\subseteq GL_{n_{0}}\times GL_{n_{1}+1}. Moreover, the induced map

𝕏(Qn1+1n1+1)𝕏(LB)=𝕏(T)\mathbb{X}^{*}(Q_{n_{1}+1}^{n_{1}+1})\longrightarrow\mathbb{X}^{*}(L\cap B)=\mathbb{X}^{*}(T)

is given in types DD and EE by

ek{ϖc1,1,ifk=1,ϖc1,kϖc1,k1,if  1<k<n1,ϖc1,n1ϖc1,n11+ϖi,ifk=n1,ϖc1,n1+ϖi,ifk=n1+1,e_{k}\longmapsto\begin{cases}\varpi_{c_{1},1},&\text{if}\;\;k=1,\\ \varpi_{c_{1},k}-\varpi_{c_{1},k-1},&\text{if}\;\;1<k<n_{1},\\ \varpi_{c_{1},n_{1}}-\varpi_{c_{1},n_{1}-1}+\varpi_{i},&\text{if}\;\;k=n_{1},\\ -\varpi_{c_{1},n_{1}}+\varpi_{i},&\text{if}\;\;k=n_{1}+1,\end{cases}

and in types CC and FF by

ek{ϖc1,1,ifk=1,ϖc1,kϖc1,k1,if  1<kn1,ϖc1,n1+2ϖi,ifk=n1+1.e_{k}\longmapsto\begin{cases}\varpi_{c_{1},1},&\text{if}\;\;k=1,\\ \varpi_{c_{1},k}-\varpi_{c_{1},k-1},&\text{if}\;\;1<k\leq n_{1},\\ -\varpi_{c_{1},n_{1}}+2\varpi_{i},&\text{if}\;\;k=n_{1}+1.\end{cases}
Proof.

Apply Proposition 2.3.1; the expressions for 𝕏(Qn1+1n1+1)𝕏(T)\mathbb{X}^{*}(Q_{n_{1}+1}^{n_{1}+1})\to\mathbb{X}^{*}(T) follow by examining the specific isomorphism given in the proof. ∎

Case 3: (G,P,μ)(G,P,\mu) is of type GG. In type GG, the Levi LL has a similarly simple form.

Lemma 2.3.5.

Assume that (G,P,μ)(G,P,\mu) is of type GG. Then there is an isomorphism

L{(λ,A)𝔾m×GL2detA=λ3}L\overset{\sim}{\longrightarrow}\{(\lambda,A)\in\mathbb{G}_{m}\times GL_{2}\mid\det A=\lambda^{3}\} (2.3.2)

such that ϖ1\varpi_{1} is identified with the character (λ,A)λ(\lambda,A)\mapsto\lambda and LBL\cap B is the preimage of the Borel subgroup 𝔾m×Q22𝔾m×GL2\mathbb{G}_{m}\times Q^{2}_{2}\subseteq\mathbb{G}_{m}\times GL_{2}. Moreover, the induced morphism

𝕏(Q22)𝕏(LB)=𝕏(T)\mathbb{X}^{*}(Q^{2}_{2})\longrightarrow\mathbb{X}^{*}(L\cap B)=\mathbb{X}^{*}(T)

sends the characters e1e_{1} and e2e_{2} to ϖ2\varpi_{2} and 3ϖ1ϖ23\varpi_{1}-\varpi_{2} respectively.

Proof.

Apply Proposition 2.3.1 again and inspect the explicit isomorphism given in the proof to compute 𝕏(Q22)𝕏(T)\mathbb{X}^{*}(Q^{2}_{2})\to\mathbb{X}^{*}(T). ∎

Case 4: (G,P,μ)(G,P,\mu) is of type BB. This case is somewhat more complicated, as the Levi subgroup LL is not of type AA. In what follows, we let

GSp4={BGL4BtJB=χ(B)Jfor someχ(B)𝔾m},GSp_{4}=\{B\in GL_{4}\mid B^{t}JB=\chi(B)J\;\text{for some}\;\chi(B)\in\mathbb{G}_{m}\},

where

J=(0001001001001000).J=\left(\begin{matrix}0&0&0&1\\ 0&0&1&0\\ 0&-1&0&0\\ -1&0&0&0\end{matrix}\right).

Note that GSp4GSp_{4} is a reductive group with weight lattice 𝕏(GSp4Q44)=k=14fk/(f1f2f3+f4)\mathbb{X}^{*}(GSp_{4}\cap Q^{4}_{4})=\bigoplus_{k=1}^{4}\mathbb{Z}f_{k}/\mathbb{Z}(f_{1}-f_{2}-f_{3}+f_{4}), where fkf_{k} is the character sending a matrix to its kkth diagonal entry, simple roots β1=f2f3\beta_{1}=f_{2}-f_{3} and β2=f1f2\beta_{2}=f_{1}-f_{2}, and simple coroots β1=f2f3\beta_{1}^{\vee}=f_{2}^{*}-f_{3}^{*} and β2=f1f2+f3f4\beta_{2}^{\vee}=f_{1}^{*}-f_{2}^{*}+f_{3}^{*}-f_{4}^{*}. In this description, χ\chi is the character χ=f1+f4=f2+f3\chi=f_{1}+f_{4}=f_{2}+f_{3}.

Lemma 2.3.6.

If (G,P,μ)(G,P,\mu) is of type BB, then there is an isomorphism

L{(A,B)GLl2×GSp4det(A)=χ(B)},L\overset{\sim}{\longrightarrow}\{(A,B)\in GL_{l-2}\times GSp_{4}\mid\det(A)=\chi(B)\},

such that ϖi=ϖl2\varpi_{i}=\varpi_{l-2} is identified with the character (A,B)det(A)=χ(B)(A,B)\mapsto\det(A)=\chi(B) and LBL\cap B is the preimage of the Borel subgroup Ql2l2×(GSp4Q44)GLl2×GSp4Q^{l-2}_{l-2}\times(GSp_{4}\cap Q^{4}_{4})\subseteq GL_{l-2}\times GSp_{4}. Moreover, the induced morphism

𝕏(GSp4Q44)=k=14fk𝕏(LB)=𝕏(T)\mathbb{X}^{*}(GSp_{4}\cap Q^{4}_{4})=\bigoplus_{k=1}^{4}\mathbb{Z}f_{k}\longrightarrow\mathbb{X}^{*}(L\cap B)=\mathbb{X}^{*}(T)

sends f1f_{1}, f2f_{2}, f3f_{3} and f4f_{4} to ϖl\varpi_{l}, ϖl1ϖl\varpi_{l-1}-\varpi_{l}, ϖl2ϖl1+ϖl\varpi_{l-2}-\varpi_{l-1}+\varpi_{l} and ϖl2ϖl\varpi_{l-2}-\varpi_{l} respectively.

Proof.

We describe the isomorphism at the level of root data.

Write

L0={(A,B)GLl2×GSp4det(A)=χ(B)}GLl2×GSp4.L_{0}=\{(A,B)\in GL_{l-2}\times GSp_{4}\mid\det(A)=\chi(B)\}\subseteq GL_{l-2}\times GSp_{4}.

The root datum (M,Ψ,M,Ψ)(M,\Psi,M^{\vee},\Psi^{\vee}) of L0L_{0} is specified as follows. The weight lattice is

M=i=1l2eij=14fjf1f2f3+f4,f1+f4i=1l2ei,M=\frac{\bigoplus_{i=1}^{l-2}\mathbb{Z}e_{i}\oplus\bigoplus_{j=1}^{4}\mathbb{Z}f_{j}}{\langle f_{1}-f_{2}-f_{3}+f_{4},f_{1}+f_{4}-\sum_{i=1}^{l-2}e_{i}\rangle},

and the coweight lattice is therefore

M={λi=1l2eij=14fj|f1+f4,λ=f2+f3,λ=i=1l2ei,λ}.M^{\vee}=\left\{\left.\lambda\in\bigoplus_{i=1}^{l-2}\mathbb{Z}e_{i}^{*}\oplus\bigoplus_{j=1}^{4}\mathbb{Z}f_{j}^{*}\,\right|\,\langle f_{1}+f_{4},\lambda\rangle=\langle f_{2}+f_{3},\lambda\rangle=\sum_{i=1}^{l-2}\langle e_{i},\lambda\rangle\right\}.

The roots Ψ\Psi and coroots Ψ\Psi^{\vee} and the bijection ΨΨ\Psi\to\Psi^{\vee} are determined by requiring that

{γi1il,il2}\{\gamma_{i}\mid 1\leq i\leq l,\,i\neq l-2\}

be a set of simple roots, where

γi={eiei+1,ifi<l2,f2f3,ifi=l1,f1f2,ifi=l,andγi={eiei+1,ifi<l2,f2f3,ifi=l1,f1f2+f3f4,ifi=l.\gamma_{i}=\begin{cases}e_{i}-e_{i+1},&\text{if}\;i<l-2,\\ f_{2}-f_{3},&\text{if}\;i=l-1,\\ f_{1}-f_{2},&\text{if}\;i=l,\end{cases}\qquad\text{and}\qquad\gamma_{i}^{\vee}=\begin{cases}e_{i}^{*}-e_{i+1}^{*},&\text{if}\;i<l-2,\\ f_{2}^{*}-f_{3}^{*},&\text{if}\;i=l-1,\\ f_{1}^{*}-f_{2}^{*}+f_{3}^{*}-f_{4}^{*},&\text{if}\;i=l.\end{cases}

There is an isomorphism

ϕ:𝕏(T)=i=1lαiM\phi\colon\mathbb{X}_{*}(T)=\bigoplus_{i=1}^{l}\mathbb{Z}\alpha_{i}^{\vee}\overset{\sim}{\longrightarrow}M^{\vee}

sending αi\alpha_{i}^{\vee} to γi\gamma_{i}^{\vee} for il2i\neq l-2, and αl2\alpha_{l-2}^{\vee} to el2+f3+f4e_{l-2}^{*}+f_{3}^{*}+f_{4}^{*}, such that the dual ϕ:M𝕏(T)\phi^{*}\colon M\to\mathbb{X}^{*}(T) sends βi\beta_{i} to αi\alpha_{i} for il2i\neq l-2. So ϕ\phi defines an isomorphism of root data, which has the desired properties by inspection. ∎

2.4. Existence of slices

In this section, we give the proof of Theorem 2.2.6. The proof we give here is somewhat ad hoc, and relies on the explicit descriptions of the Levi subgroups given in §2.3.

We first give the construction in type AA.

Proof of Theorem 2.2.6 in type AA.

First note that since d=1d=1 in this case, the μd=μ1\mu_{d}=\mu_{1}-gerbe 𝔊uni\mathfrak{G}^{uni} must be the trivial one.

Using the identification LGLi×GLliL\cong GL_{i}\times GL_{l-i}, Atiyah’s classification of stable vector bundles (in the form [davis19, Theorem 4.2.6]) implies that the morphism

(ϖi,ϖi+1):BunL,rigss,μPicS1(E)×SPicS1(E)(\varpi_{i},\varpi_{i+1})\colon\mathrm{Bun}_{L,rig}^{ss,\mu}\longrightarrow\mathrm{Pic}^{-1}_{S}(E)\times_{S}\mathrm{Pic}^{-1}_{S}(E)

is a trivial Z(L)rigZ(L)_{rig}-gerbe. Note that in particular, all semistable LL-bundles of slope μ\mu are regular in this case.

By [davis19, Proposition 4.1.15], the pullback of Θ\Theta to BunL,rigss,μ\mathrm{Bun}_{L,rig}^{ss,\mu} has Z(L)rigZ(L)_{rig}-weight (μ|)𝕏(Z(L)rig)(-\mu{\,|\,}-)\in\mathbb{X}^{*}(Z(L)_{rig}). (This means that tensoring a map XBunL,rigX\to\mathrm{Bun}_{L,rig} with a Z(L)rigZ(L)_{rig}-torsor η\eta on XX tensors the pullback of ΘBunG,rig\Theta_{\mathrm{Bun}_{G,rig}} to XX with the line bundle (μ|η)(-\mu{\,|\,}\eta).) Since the corresponding homomorphism 𝕏(Z(L)rig)\mathbb{X}_{*}(Z(L)_{rig})\to\mathbb{Z} is surjective, it follows that there exists a section

PicS1(E)×SPicS1(E)BunL,rigss,μ\mathrm{Pic}^{-1}_{S}(E)\times_{S}\mathrm{Pic}^{-1}_{S}(E)\longrightarrow\mathrm{Bun}_{L,rig}^{ss,\mu}

such that the pullback of ΘBunG,rig\Theta_{\mathrm{Bun}_{G,rig}} is trivial. Since such a section is necessarily smooth, composing it with any choice of section of

PicS1(E)×SPicS1(E)PicS1(E)×SPicS1(E)/EE\mathrm{Pic}^{-1}_{S}(E)\times_{S}\mathrm{Pic}^{-1}_{S}(E)\longrightarrow\mathrm{Pic}^{-1}_{S}(E)\times_{S}\mathrm{Pic}^{-1}_{S}(E)/E\cong E

gives a Θ\Theta-trivial slice Z0BunL,rigss,μZ_{0}\to\mathrm{Bun}_{L,rig}^{ss,\mu} with Z0=EZ_{0}=E, such that Z0BunL,rigss,μ/EZ_{0}\to\mathrm{Bun}_{L,rig}^{ss,\mu}/E is surjective with fibres isomorphic to Z(L)rigZ(L)_{rig}, hence connected. So (1), (2) and (3) are satisfied. A simple root-theoretic calculation shows that 2ρ,μ=l+2-\langle 2\rho,\mu\rangle=l+2, so (4) follows from Proposition 2.2.3. So this proves the theorem in this case. ∎

The construction in the exceptional types EE, FF and GG is also fairly straightforward.

Proof of Theorem 2.2.6 in types EE, FF and GG.

In these cases, Proposition 2.3.1 and Atiyah’s theorem show that the morphism

ϖi:BunL,rigss,μPicS1(E)\varpi_{i}\colon\mathrm{Bun}_{L,rig}^{ss,\mu}\longrightarrow\mathrm{Pic}^{-1}_{S}(E) (2.4.1)

is a 𝔾m=Z(L)rig\mathbb{G}_{m}=Z(L)_{rig}-gerbe. Let Z0=SZ_{0}=S, and let 𝔊\mathfrak{G}^{\prime} be the Z(L)rigZ(L)_{rig}-gerbe given by the pullback along 𝒪(OE):Z0PicS1(E)\mathcal{O}(-O_{E})\colon Z_{0}\to\mathrm{Pic}^{-1}_{S}(E). By [davis19, Proposition 4.1.15], the pullback of the theta bundle defines a 𝔹Z(L)rig\mathbb{B}Z(L)_{rig}-equivariant morphism 𝔊𝔹𝔾m\mathfrak{G}^{\prime}\to\mathbb{B}\mathbb{G}_{m}, where 𝔹Z(L)rig\mathbb{B}Z(L)_{rig} acts on 𝔹𝔾m\mathbb{B}\mathbb{G}_{m} through the homomorphism

(μ|):Z(L)rig𝔾m.-(\mu{\,|\,}-)\colon Z(L)_{rig}\longrightarrow\mathbb{G}_{m}.

So a section of 𝔊\mathfrak{G}^{\prime} such that the pullback of ΘBunG,rig\Theta_{\mathrm{Bun}_{G,rig}} is trivial is the same thing as a section of the μd=ker(μ|)\mu_{d}=\ker(\mu{\,|\,}-)-gerbe 𝔊=𝔊×𝔹𝔾mSpec\mathfrak{G}=\mathfrak{G}^{\prime}\times_{\mathbb{B}\mathbb{G}_{m}}\operatorname{\mathrm{Spec}}\mathbb{Z}. The μd\mu_{d}-gerbe is by construction pulled back from one 𝔊uni\mathfrak{G}^{uni} on M1,1M_{1,1}, defined in the same way, and if it is trivial then there is a morphism Z0BunL,rigss,μZ_{0}\to\mathrm{Bun}_{L,rig}^{ss,\mu} lifting the section 𝒪(OE):Z0PicS1(E)\mathcal{O}(-O_{E})\colon Z_{0}\to\mathrm{Pic}^{-1}_{S}(E) such that the pullback of ΘBunG,rig\Theta_{\mathrm{Bun}_{G,rig}} is trivial.

It is immediately clear that (1) is satisfied. Letting (BunL,rigss,μ)0(\mathrm{Bun}_{L,rig}^{ss,\mu})_{0} be the fibre of (2.4.1) over 𝒪(OE):SPicS1(E)\mathcal{O}(-O_{E})\colon S\to\mathrm{Pic}^{-1}_{S}(E), we have that (BunL,rigss,μ)0BunL,rigss,μ/E(\mathrm{Bun}_{L,rig}^{ss,\mu})_{0}\cong\mathrm{Bun}_{L,rig}^{ss,\mu}/E is a Z(L)rigZ(L)_{rig}-gerbe over S=Z0S=Z_{0} and the map Z0BunL,rigss,μ/EZ_{0}\to\mathrm{Bun}_{L,rig}^{ss,\mu}/E is a section. In particular, it is smooth with connected fibres (isomorphic to Z(L)rigZ(L)_{rig}), so (2) is satisfied, and surjective, so (3) is also satisfied. Finally, to prove (4), simply note that Proposition 2.2.3 and a simple root-theoretic calculation show that Z=IndLG(Z0)Z0=SZ=\mathrm{Ind}_{L}^{G}(Z_{0})\to Z_{0}=S is an affine space bundle of relative dimension l+3l+3. ∎

The proof of Theorem 2.2.6 in types BB, CC and DD will require a few more preliminaries. First, we remark on the following realisation of GSp4GSp_{4}-bundles in terms of vector bundles.

Definition 2.4.1.

A conformally symplectic vector bundle is a tuple (W,M,ω)(W,M,\omega), where WW is a vector bundle, MM is a line bundle, and ω:2WM\omega\colon\wedge^{2}W\to M is a morphism such that the induced morphism WWMW\to W^{\vee}\otimes M is an isomorphism.

Lemma 2.4.2.

There is an isomorphism of BunGSp4\mathrm{Bun}_{GSp_{4}} with the relative stack of conformally symplectic vector bundles (W,M,ω)(W,M,\omega) on EE over SS, which identifies χ:BunGSp4Bun𝔾m\chi\colon\mathrm{Bun}_{GSp_{4}}\to\mathrm{Bun}_{\mathbb{G}_{m}} with the map (W,M,ω)M(W,M,\omega)\mapsto M.

Proof.

Let VV be the standard representation of GSp4GSp_{4} coming from the inclusion GSp4GL4GSp_{4}\subseteq GL_{4}. Then JJ defines a homomorphism of GSp4GSp_{4}-representations J:2VχJ\colon\wedge^{2}V\to\mathbb{Z}_{\chi}. The isomorphism from BunGSp4\mathrm{Bun}_{GSp_{4}} to the stack of conformally symplectic vector bundles sends a GSp4GSp_{4}-bundle ξ\xi to (ξ×GSp4V,ξ×GSp4χ,ξ×GSp4J)(\xi\times^{GSp_{4}}V,\xi\times^{GSp_{4}}\mathbb{Z}_{\chi},\xi\times^{GSp_{4}}J). ∎

Now assume that (G,P,μ)(G,P,\mu) is a subregular Harder-Narasimhan class of type BB, CC or DD with corresponding Levi LPL\subseteq P. Let PLP^{\prime}\subseteq L denote the standard parabolic of type t(P)={αl}t(P^{\prime})=\{\alpha_{l}\}, and LPL^{\prime}\subseteq P^{\prime} its standard Levi subgroup. In types CC and DD, let ρL:LGLn1+1\rho_{L}\colon L\to GL_{n_{1}+1} be the composition of the isomorphism of Lemma 2.3.4 with the projection to the second factor (where for concreteness we choose the labelling so that αc1,n1=αl\alpha_{c_{1},n_{1}}=\alpha_{l}), and in type BB let ρL:LGL4\rho_{L}\colon L\to GL_{4} be the composition of the isomorphism of Lemma 2.3.6 with the projection to the second factor and the inclusion GSp4GL4GSp_{4}\subseteq GL_{4}.

Lemma 2.4.3.

Assume we are in types BB, CC or DD. Then there is an isomorphism of BunP\mathrm{Bun}_{P^{\prime}} with the stack of pairs (ξL,MW)(\xi_{L},M\subseteq W), where ξLBunL\xi_{L}\in\mathrm{Bun}_{L} and MWM\subseteq W is a line subbundle of the vector bundle WW associated to ξL\xi_{L} under the representation ρL\rho_{L}, such that the morphism

ϖl:BunPBun𝔾m\varpi_{l}\colon\mathrm{Bun}_{P^{\prime}}\longrightarrow\mathrm{Bun}_{\mathbb{G}_{m}}

is identified with the morphism

(ξL,MW){ϖi(ξL)M1,in types B and D,ϖi(ξL)2M1,in type C.(\xi_{L},M\subseteq W)\longmapsto\begin{cases}\varpi_{i}(\xi_{L})\otimes M^{-1},&\text{in types }B\text{ and }D,\\ \varpi_{i}(\xi_{L})^{\otimes 2}\otimes M^{-1},&\text{in type }C.\end{cases}

In types CC and DD (resp., type BB), if ξP\xi_{P^{\prime}} corresponds to (ξL,MW)(\xi_{L},M\subseteq W) and VV is the vector bundle induced by ξL\xi_{L} under the projection LGLn0L\to GL_{n_{0}} coming from Lemma 2.3.4 (resp., 2.3.6), then the LL^{\prime}-bundle ξP×PL\xi_{P^{\prime}}\times^{P^{\prime}}L^{\prime} is semistable if and only if the vector bundles VV and W/MW/M (resp., ker(ω:W/MdetVM)\ker(\omega\colon W/M\to\det V\otimes M^{\vee})) are semistable.

Proof.

In types CC and DD, the isomorphism of Lemma 2.3.4 identifies PP^{\prime} with the parabolic GLn0×Rn1+1GL_{n_{0}}\times R_{n_{1}+1}, where Rn1+1GLn1+1R_{n_{1}+1}\subseteq GL_{n_{1}+1} is the maximal parabolic of type {βn1}\{\beta_{n_{1}}\}, and the result follows routinely. In type BB, using Lemma 2.3.6 we have an LL-equivariant identification L/PGSp4/(GSp4R4)GL4/R44L/P^{\prime}\cong GSp_{4}/(GSp_{4}\cap R_{4})\cong GL_{4}/R_{4}\cong\mathbb{P}^{4} with the space of lines in the representation ρL\rho_{L}, where

R4={(000)}GL4.R_{4}=\left\{\left(\begin{matrix}*&*&*&0\\ *&*&*&0\\ *&*&*&0\\ *&*&*&*\end{matrix}\right)\right\}\subseteq GL_{4}.

The claimed isomorphism in this case now follows. To get the desired identification of the semistable bundles, notice that the Levi factor of GSp4R4GSp_{4}\cap R_{4} is

{(λ1detA000000A0000λ)|AGL2,λ𝔾m}GL2×𝔾m,\left\{\left.\left(\begin{array}[]{c|cc|c}\lambda^{-1}\det A&0&0&0\\ \hline\cr 0&&&0\\ 0&\lx@intercol\hfil\smash{\raisebox{6.0pt}{$A$}}\hfil\lx@intercol\vrule\lx@intercol&0\\ \hline\cr\\[-12.0pt] 0&0&0&\lambda\end{array}\right)\right|A\in GL_{2},\lambda\in\mathbb{G}_{m}\right\}\cong GL_{2}\times\mathbb{G}_{m},

so we have an isomorphism

BunLBunGLn0×Bun𝔾mBunGL2×SBun𝔾m,\mathrm{Bun}_{L^{\prime}}\cong\mathrm{Bun}_{GL_{n_{0}}}\times_{\mathrm{Bun}_{\mathbb{G}_{m}}}\mathrm{Bun}_{GL_{2}}\times_{S}\mathrm{Bun}_{\mathbb{G}_{m}},

such that the map BunPBunL\mathrm{Bun}_{P^{\prime}}\to\mathrm{Bun}_{L^{\prime}} is identified with

(ξL,MW)(V,ker(W/MdetVM),M).(\xi_{L},M\subseteq W)\longmapsto(V,\ker(W/M\to\det V\otimes M^{\vee}),M).

This now implies the claim. ∎

Lemma 2.4.4.

Let (G,P,μ)(G,P,\mu) be of type BB, CC or DD, and assume that ξLEs\xi_{L}\to E_{s} is a semistable LL-bundle of slope μ\mu over a geometric fibre of ESE\to S. Then dimAut(ξL)2\dim\mathrm{Aut}(\xi_{L})\geq 2.

Proof.

By Lemmas 2.3.4, 2.3.6 and 2.4.2 and [davis19, Theorem 4.2.6], it suffices to show that if WW is a semistable vector bundle of degree 2-2 and rank 2r2r (resp., (W,M,ω)(W,M,\omega) is a conformally symplectic vector bundle with WW semistable and degM=1\deg M=-1), then dimAut(W)2\dim\mathrm{Aut}(W)\geq 2 (resp., dimAut(W,M,ω)2\dim\mathrm{Aut}(W,M,\omega)\geq 2).

In the first case, observe that if UU and UU^{\prime} are nonisomorphic semistable vector bundles of degree 1-1 and rank rr, then U(U)U\otimes(U^{\prime})^{\vee} is a vector bundle of degree 0 with H0(E,U(U))=0H^{0}(E,U\otimes(U^{\prime})^{\vee})=0, and hence H1(E,U(U))=0H^{1}(E,U\otimes(U^{\prime})^{\vee})=0 also. It follows that the morphism

BunGLrss,1×BunGLrss,1\displaystyle\mathrm{Bun}_{GL_{r}}^{ss,-1}\times\mathrm{Bun}_{GL_{r}}^{ss,-1} BunGL2rss,2\displaystyle\longrightarrow\mathrm{Bun}_{GL_{2r}}^{ss,-2}
(U,U)\displaystyle(U,U^{\prime}) UU\displaystyle\longmapsto U\oplus U^{\prime}

is étale at (U,U)(U,U^{\prime}) if U≇UU\not\cong U^{\prime}. Since the locus of vector bundles WW in BunGL2dss,2\mathrm{Bun}_{GL_{2d}}^{ss,-2} with dimAut(W)<2\dim\mathrm{Aut}(W)<2 is open, it is either empty or dense. So by openness of étale morphisms, if it is nonempty, then there exists such a bundle W=UUW=U\oplus U^{\prime} with U≇UU\not\cong U^{\prime}. But Aut(W)=Aut(U)×Aut(U)=𝔾m×𝔾m\mathrm{Aut}(W)=\mathrm{Aut}(U)\times\mathrm{Aut}(U^{\prime})=\mathbb{G}_{m}\times\mathbb{G}_{m} for such bundles, so this is a contradiction and we are done in this case.

The proof for conformally symplectic bundles is similar. Consider the Levi subgroup

GL2×𝔾mL′′={(λJ0(At)1J000λA)|AGL2,λ𝔾m},GL_{2}\times\mathbb{G}_{m}\cong L^{\prime\prime}=\left\{\left.\left(\begin{array}[]{c|c}\lambda J_{0}(A^{t})^{-1}J_{0}&0\\ \hline\cr 0&\lambda A\end{array}\right)\right|A\in GL_{2},\lambda\in\mathbb{G}_{m}\right\},

where

J0=(0110).J_{0}=\left(\begin{matrix}0&1\\ 1&0\end{matrix}\right).

Given (U,M)BunGL2ss,1×SBun𝔾m1(U,M)\in\mathrm{Bun}_{GL_{2}}^{ss,-1}\times_{S}\mathrm{Bun}_{\mathbb{G}_{m}}^{-1} corresponding to an L′′L^{\prime\prime}-bundle ξL′′\xi_{L^{\prime\prime}}, with U≇UMU\not\cong U^{\vee}\otimes M, we have that

ξL′′×L′′𝔤𝔰𝔭4/𝔩′′U(UM)U(UM)\xi_{L^{\prime\prime}}\times^{L^{\prime\prime}}\mathfrak{gsp}_{4}/\mathfrak{l}^{\prime\prime}\subseteq U^{\vee}\otimes(U^{\vee}\otimes M)\oplus U\otimes(U^{\vee}\otimes M)^{\vee}

is a degree 0 vector bundle on EsE_{s} with H0(Es,ξL′′×L′′𝔤𝔰𝔭4/𝔩′′)=0H^{0}(E_{s},\xi_{L^{\prime\prime}}\times^{L^{\prime\prime}}\mathfrak{gsp}_{4}/\mathfrak{l}^{\prime\prime})=0 and hence H1(Es,ξL′′×L′′𝔤𝔰𝔭4/𝔩′′)=0H^{1}(E_{s},\xi_{L^{\prime\prime}}\times^{L^{\prime\prime}}\mathfrak{gsp}_{4}/\mathfrak{l}^{\prime\prime})=0 also, where 𝔤𝔰𝔭4=Lie(GSp4)\mathfrak{gsp}_{4}=\mathrm{Lie}(GSp_{4}) and 𝔩′′=Lie(L′′)\mathfrak{l}^{\prime\prime}=\mathrm{Lie}(L^{\prime\prime}). So we conclude that the morphism

BunL′′BunGSp41\mathrm{Bun}_{L^{\prime\prime}}\longrightarrow\mathrm{Bun}_{GSp_{4}}^{-1}

is étale at (U,M)(U,M).

Since the locus of conformally symplectic vector bundles (W,M,ω)(W,M,\omega) in BunGSp4ss,1\mathrm{Bun}_{GSp_{4}}^{ss,-1} with automorphism group of dimension <2<2 is open, it is either empty or dense. If it is nonempty, then by openness of étale morphisms we can find such a bundle of the form W=UUMW=U\oplus U^{\vee}\otimes M as above. But dimAutGSp4(W)=dimAut(U)×dimAut(M)=2\dim\mathrm{Aut}_{GSp_{4}}(W)=\dim\mathrm{Aut}(U)\times\dim\mathrm{Aut}(M)=2, so this is a contradiction and the lemma is proved. ∎

Proof of Theorem 2.2.6 in types BB, CC and DD.

Let μ𝕏(Z(L))\mu^{\prime}\in\mathbb{X}_{*}(Z(L^{\prime}))_{\mathbb{Q}} be the unique vector with ϖi,μ=1\langle\varpi_{i},\mu^{\prime}\rangle=-1 and ϖl,μ=0\langle\varpi_{l},\mu^{\prime}\rangle=0. Then Proposition 2.3.1 and Atiyah’s classification show that the morphism

(ϖi,ϖl):BunL,rigss,μPicS1(E)×SPicS0(E)(\varpi_{i},\varpi_{l})\colon\mathrm{Bun}_{L^{\prime},rig}^{ss,\mu^{\prime}}\longrightarrow\mathrm{Pic}^{-1}_{S}(E)\times_{S}\mathrm{Pic}^{0}_{S}(E)

is a Z(L)rigZ(L^{\prime})_{rig}-gerbe. Let 𝔊′′\mathfrak{G}^{\prime\prime} be the Z(L)rigZ(L^{\prime})_{rig}-gerbe on SS given by pulling back along the section

(𝒪(OE),𝒪):SPicS1(E)×SPicS0(E).(\mathcal{O}(-O_{E}),\mathcal{O})\colon S\to\mathrm{Pic}^{-1}_{S}(E)\times_{S}\mathrm{Pic}^{0}_{S}(E).

The pullback of the theta bundle gives a 𝔹Z(L)rig\mathbb{B}Z(L^{\prime})_{rig}-equivariant morphism 𝔊′′𝔹𝔾m\mathfrak{G}^{\prime\prime}\to\mathbb{B}\mathbb{G}_{m} where 𝔹Z(L)rig\mathbb{B}Z(L^{\prime})_{rig} acts through the homomorphism (μ|)(-\mu^{\prime}{\,|\,}-), so we get a ker(μ|)\ker(\mu^{\prime}{\,|\,}-)-gerbe 𝔊=𝔊′′×𝔹𝔾mSpec\mathfrak{G}^{\prime}=\mathfrak{G}^{\prime\prime}\times_{\mathbb{B}\mathbb{G}_{m}}\operatorname{\mathrm{Spec}}\mathbb{Z}. Let 𝔊\mathfrak{G} be the rigidification of 𝔊\mathfrak{G}^{\prime} with respect to ϖl(𝔾m)\varpi_{l}^{\vee}(\mathbb{G}_{m}). Then 𝔊\mathfrak{G} is a ker(μ|)/ϖl(𝔾m)μd\ker(\mu^{\prime}{\,|\,}-)/\varpi_{l}^{\vee}(\mathbb{G}_{m})\cong\mu_{d}-gerbe, pulled back from a gerbe 𝔊uni\mathfrak{G}^{uni} on M1,1M_{1,1}, and if it is trivial then we have a 𝔹𝔾m\mathbb{B}\mathbb{G}_{m}-equivariant morphism 𝔹S𝔾mBunL,rigss,μ\mathbb{B}_{S}\mathbb{G}_{m}\to\mathrm{Bun}_{L^{\prime},rig}^{ss,\mu^{\prime}} (with 𝔹𝔾m\mathbb{B}\mathbb{G}_{m} acting through ϖl\varpi_{l}^{\vee}) lifting the section (𝒪(OE),𝒪)(\mathcal{O}(-O_{E}),\mathcal{O}) such that the pullback of the theta bundle is trivial. Define

Z0=IndLL(𝔹S𝔾m)𝔹S𝔾mBunL,rigμ,Z_{0}=\mathrm{Ind}_{L^{\prime}}^{L}(\mathbb{B}_{S}\mathbb{G}_{m})\setminus\mathbb{B}_{S}\mathbb{G}_{m}\longrightarrow\mathrm{Bun}_{L,rig}^{\mu},

and observe that the pullback of ΘBunG,rig\Theta_{\mathrm{Bun}_{G,rig}} to Z0Z_{0} is also trivial since Z0𝔹S𝔾mZ_{0}\to\mathbb{B}_{S}\mathbb{G}_{m} is an affine space bundle.

Type αΦL\alpha\in\Phi_{L} with α,μ<0\langle\alpha,\mu^{\prime}\rangle<0 α,μ\langle\alpha,\mu^{\prime}\rangle α,ϖl\langle\alpha,\varpi_{l}^{\vee}\rangle
BB αl-\alpha_{l} 12-\frac{1}{2} 1-1
αl1αl-\alpha_{l-1}-\alpha_{l} 12-\frac{1}{2} 1-1
αl12αl-\alpha_{l-1}-2\alpha_{l} 1-1 2-2
CC αl-\alpha_{l} 2-2 1-1
DD αl-\alpha_{l} 23-\frac{2}{3} 1-1
αl2αl-\alpha_{l-2}-\alpha_{l} 23-\frac{2}{3} 1-1
αl2αl1αl-\alpha_{l-2}-\alpha_{l-1}-\alpha_{l} 23-\frac{2}{3} 1-1
Table 2. Roots of LL with α,μ<0\langle\alpha,\mu^{\prime}\rangle<0

We now check that Z0Z_{0} satisfies the conditions of Theorem 2.2.6. Since the claims are local on SS, we can assume for convenience that the section 𝔹S𝔾mBunL,rigss,μ\mathbb{B}_{S}\mathbb{G}_{m}\to\mathrm{Bun}_{L^{\prime},rig}^{ss,\mu^{\prime}} lifts to a morphism SBunLss,μS\to\mathrm{Bun}_{L^{\prime}}^{ss,\mu^{\prime}} and that the line bundle on EE associated to this section via the character ϖl\varpi_{l} is trivial. Note that in this case, we have a natural identification

Z0(IndLL(S)S)/𝔾m.Z_{0}\cong(\mathrm{Ind}_{L^{\prime}}^{L}(S)\setminus S)/\mathbb{G}_{m}.

First, the roots αΦL\alpha\in\Phi_{L} with α,μ<0\langle\alpha,\mu^{\prime}\rangle<0 are given in Table 2, along with the values of α,μ\langle\alpha,\mu^{\prime}\rangle and α,ϖl\langle\alpha,\varpi_{l}^{\vee}\rangle. Using Proposition 2.2.3 and [davis19a, Proposition 5.2.7], it follows that IndLL(S)S\mathrm{Ind}_{L^{\prime}}^{L}(S)\to S is an 𝔸2\mathbb{A}^{2}-bundle on which 𝔾m\mathbb{G}_{m} acts with weight 11 in types CC and DD, and weights 11 and 22 in type BB. So Z0SZ_{0}\to S is a (1,2)\mathbb{P}(1,2)-bundle in type BB and a 1\mathbb{P}^{1}-bundle in types CC and DD. In particular, (1) is satisfied.

We next show that Z0BunL,rigμZ_{0}\to\mathrm{Bun}_{L,rig}^{\mu} factors through BunL,rigss,μ\mathrm{Bun}_{L,rig}^{ss,\mu}. Note that Table 2 shows that μ-\mu^{\prime} is a Harder-Narasimhan vector for PLP^{\prime}\subseteq L, so IndLL(S)=BunP,rigss,μ×BunL,rigss,μS\mathrm{Ind}_{L^{\prime}}^{L}(S)=\mathrm{Bun}_{P^{\prime},rig}^{ss,\mu^{\prime}}\times_{\mathrm{Bun}_{L^{\prime},rig}^{ss,\mu^{\prime}}}S. So Lemma 2.4.3 shows that ξL\xi_{L} is in the image of IndLL(S)\mathrm{Ind}_{L^{\prime}}^{L}(S) if and only if VV is semistable of determinant 𝒪(OE)\mathcal{O}(-O_{E}) and there exists a nonvanishing section of W𝒪(dOE)W\otimes\mathcal{O}(dO_{E}) such that the vector bundle

U={W/𝒪(dOE),in types C and D,ker(W/𝒪((dOE)𝒪),in type B,U=\begin{cases}W/\mathcal{O}(-dO_{E}),&\text{in types }C\text{ and }D,\\ \ker(W/\mathcal{O}(-(dO_{E})\to\mathcal{O}),&\text{in type }B,\end{cases}

is semistable. Here VV and WW are as in the statement of Lemma 2.4.3, and

d={1,in typesBandD,2,in typeCd=\begin{cases}1,&\text{in types}\;B\;\text{and}\;D,\\ 2,&\text{in type}\;C\end{cases}

is as in the statement of the theorem. The bundle ξL\xi_{L} is in the image of IndLL(S)S\mathrm{Ind}_{L^{\prime}}^{L}(S)\setminus S if and only if 𝒪(dOE)W\mathcal{O}(-dO_{E})\to W can be chosen not to admit a retraction. Suppose that ξL\xi_{L} is such a bundle and that ξL\xi_{L} is unstable; we deduce a contradiction in each type.

In type BB, WW is an unstable conformally symplectic vector bundle of rank 44 and degree 2-2, so there exists a quotient WNW\to N where NN has slope <1/2<-1/2. Replacing NN with coker(N𝒪(OE)W)\operatorname{\mathrm{coker}}(N^{\vee}\otimes\mathcal{O}(-O_{E})\to W) if necessary, we can assume that NN has rank 2\leq 2. Since any vector bundle of rank 22 and slope <1/2<-1/2 surjects onto some line bundle of negative degree, we can therefore assume without loss of generality that NN is a line bundle. Examining slopes, we see from semistability of UU that WNW\to N does not factor through W/𝒪(OE)W/\mathcal{O}(-O_{E}), and hence that 𝒪(OE)N\mathcal{O}(-O_{E})\to N is nonzero. So 𝒪(OE)N\mathcal{O}(-O_{E})\to N must be an isomorphism since degNdeg𝒪(OE)\deg N\leq\deg\mathcal{O}(-O_{E}), and we therefore have a retraction W𝒪(OE)=NW\to\mathcal{O}(-O_{E})=N. Since this is a contradiction, we are done in this case.

In type CC, WW is an unstable vector bundle of rank 22 and degree 2-2, so there exists a quotient WNW\to N where NN is a a line bundle of degree <1<-1. Examining slopes, we see that WNW\to N does not factor through W/𝒪(2OE)W/\mathcal{O}(-2O_{E}) and hence that 𝒪(2OE)N\mathcal{O}(-2O_{E})\to N is nonzero. So 𝒪(2OE)N\mathcal{O}(-2O_{E})\to N must be an isomorphism since degNdeg𝒪(2OE)\deg N\leq\deg\mathcal{O}(-2O_{E}), and we therefore have a retraction W𝒪(2OE)=NW\to\mathcal{O}(-2O_{E})=N. Since this is a contradiction, we are done in this case as well.

Finally, in type DD, WW is an unstable vector bundle of rank 44 and degree 2-2, so there exists a quotient WNW\to N where NN is a semistable vector bundle of slope <1/2<-1/2. Examining slopes and using semistability of W/𝒪(OE)W/\mathcal{O}(-O_{E}) and of NN, we see that WNW\to N does not factor through W/𝒪(OE)W/\mathcal{O}(-O_{E}) and we again get a retraction WN𝒪(OE)W\to N\cong\mathcal{O}(-O_{E}). So we have shown that ξL\xi_{L} must be semistable in all cases.

We next show that the morphism Z0BunL,rigss,μ/EZ_{0}\to\mathrm{Bun}_{L,rig}^{ss,\mu}/E is smooth with connected fibres, which proves (2) and that Z0BunL,rigss,μZ_{0}\to\mathrm{Bun}_{L,rig}^{ss,\mu} is a Θ\Theta-trivial slice. Write (BunLss,μ)0(\mathrm{Bun}_{L}^{ss,\mu})_{0} for the fibre of ϖi:BunLss,μPicS1(E)\varpi_{i}\colon\mathrm{Bun}_{L}^{ss,\mu}\to\mathrm{Pic}^{-1}_{S}(E) over 𝒪(OE)\mathcal{O}(-O_{E}) and (BunPμ)0ss=BunPμ×BunLμ(BunLss,μ)0(\mathrm{Bun}_{P^{\prime}}^{\mu^{\prime}})^{ss}_{0}=\mathrm{Bun}_{P^{\prime}}^{\mu^{\prime}}\times_{\mathrm{Bun}_{L}^{\mu}}(\mathrm{Bun}_{L}^{ss,\mu})_{0}. Then Lemma 2.4.3 gives an open immersion

(BunPμ)0ss(BunLss,μ)0π(Wuni𝒪(dOE)),(\mathrm{Bun}_{P^{\prime}}^{\mu^{\prime}})_{0}^{ss}\subseteq\mathbb{P}_{(\mathrm{Bun}_{L}^{ss,\mu})_{0}}\pi_{*}(W^{uni}\otimes\mathcal{O}(dO_{E})),

where we write WuniW^{uni} for the universal bundle on (BunLss,μ)0×SE(\mathrm{Bun}_{L}^{ss,\mu})_{0}\times_{S}E induced by the representation ρL\rho_{L} of LL and π:(BunLss,μ)0×SE(BunLss,μ)0\pi\colon(\mathrm{Bun}_{L}^{ss,\mu})_{0}\times_{S}E\to(\mathrm{Bun}_{L}^{ss,\mu})_{0} for the natural projection. Moreover,

Z0×(BunL,rigss,μ)0(BunLss,μ)0(BunPμ)0ssZ_{0}\times_{(\mathrm{Bun}_{L,rig}^{ss,\mu})_{0}}(\mathrm{Bun}_{L}^{ss,\mu})_{0}\longrightarrow(\mathrm{Bun}_{P^{\prime}}^{\mu^{\prime}})^{ss}_{0}

is a 𝔾m=Z(L)rig/ϖl(𝔾m)\mathbb{G}_{m}=Z(L)_{rig}/\varpi_{l}^{\vee}(\mathbb{G}_{m})-torsor over the open substack where the associated LL^{\prime}-bundle is semistable. This shows in particular that Z0×(BunL,rigss,μ)0(BunLss,μ)0(BunLss,μ)0Z_{0}\times_{(\mathrm{Bun}_{L,rig}^{ss,\mu})_{0}}(\mathrm{Bun}_{L}^{ss,\mu})_{0}\to(\mathrm{Bun}_{L}^{ss,\mu})_{0} is smooth with connected fibres of dimension 22, and hence that the same is true for Z0(BunL,rigss,μ)0BunL,rigss,μ/EZ_{0}\to(\mathrm{Bun}_{L,rig}^{ss,\mu})_{0}\cong\mathrm{Bun}_{L,rig}^{ss,\mu}/E as claimed.

To prove (3), first observe that since Z0SZ_{0}\to S has finite relative stabilisers, any LL-bundle in the image of Z0(BunL,rigss,μ)0BunL,rigss,μZ_{0}\to(\mathrm{Bun}_{L,rig}^{ss,\mu})_{0}\subseteq\mathrm{Bun}_{L,rig}^{ss,\mu} can have automorphism group of dimension at most 22, and is hence regular by Lemma 2.4.4. For the converse, note that since every regular semistable LL-bundle is a translate of one in (BunLss,μ)0(\mathrm{Bun}_{L}^{ss,\mu})_{0}, it suffices to show that any regular semistable bundle in (BunLss,μ)0(\mathrm{Bun}_{L}^{ss,\mu})_{0} is in the image of (BunPss,μ)0BunLμ(\mathrm{Bun}_{P^{\prime}}^{ss,\mu})_{0}\to\mathrm{Bun}_{L}^{\mu}, and hence in the image of Z0BunL,rigss,μZ_{0}\to\mathrm{Bun}_{L,rig}^{ss,\mu}.

Suppose then that ξLEs\xi_{L}\to E_{s} is a semistable LL-bundle in (BunLss,μ)0(\mathrm{Bun}_{L}^{ss,\mu})_{0} over s:SpeckSs\colon\operatorname{\mathrm{Spec}}k\to S that is not in the image of (BunPss,μ)0(\mathrm{Bun}_{P^{\prime}}^{ss,\mu^{\prime}})_{0}. We show in each type that dimAut(ξL)>2\dim\mathrm{Aut}(\xi_{L})>2 so ξL\xi_{L} is not regular.

In type BB, in the notation of Lemma 2.4.3, we have that for every nonzero morphism γ:𝒪(OE)W\gamma\colon\mathcal{O}(-O_{E})\to W, the vector bundle Uγ=ker(W/𝒪(OE)𝒪)U_{\gamma}=\ker(W/\mathcal{O}(-O_{E})\to\mathcal{O}) is unstable. (Note that WW is semistable of rank 44 and degree 2-2, so any such morphism is a subbundle.) Using semistability of WW, the Harder-Narasimhan decomposition of UγU_{\gamma} must be of the form Uγ=NγNγ𝒪(OE)U_{\gamma}=N_{\gamma}\oplus N_{\gamma}^{\vee}\otimes\mathcal{O}(-O_{E}), where NγN_{\gamma} is a line bundle of degree 0 on EsE_{s} and the preimage of NγN_{\gamma} in WW is the unique non-split extension NγN_{\gamma}^{\prime} of NγN_{\gamma} by 𝒪(OE)\mathcal{O}(-O_{E}). By uniqueness of Harder-Narasimhan filtrations, it follows that we have a morphism k1=H0(Es,W𝒪(OE))Pic0(Es)\mathbb{P}^{1}_{k}=\mathbb{P}H^{0}(E_{s},W\otimes\mathcal{O}(O_{E}))\to\mathrm{Pic}^{0}(E_{s}) sending γ\gamma to the isomorphism class of NγN_{\gamma}. Since there are no non-constant morphisms from k1\mathbb{P}^{1}_{k} to any elliptic curve over kk, we deduce that Nγ=NN_{\gamma}=N and Nγ=NN_{\gamma}^{\prime}=N^{\prime} are independent of γ\gamma. So every nonzero morphism 𝒪(OE)W\mathcal{O}(-O_{E})\to W factors through some Lagrangian inclusion NWN^{\prime}\hookrightarrow W. Choosing any such morphism gives an exact sequence

0NW(N)𝒪(OE)0.0\longrightarrow N^{\prime}\longrightarrow W\longrightarrow(N^{\prime})^{\vee}\otimes\mathcal{O}(-O_{E})\longrightarrow 0.

Since dimHom(𝒪(OE),N)=1\dim\mathrm{Hom}(\mathcal{O}(-O_{E}),N^{\prime})=1 and dimHom(𝒪(OE),W)=2\dim\mathrm{Hom}(\mathcal{O}(-O_{E}),W)=2, we can choose another homomorphism 𝒪(OE)\mathcal{O}(-O_{E}) not factoring through the given copy of NN^{\prime}, and hence get another Lagrangian inclusion NWN^{\prime}\hookrightarrow W, which must map NN^{\prime} isomorphically onto (N)𝒪(OE)(N^{\prime})^{\vee}\otimes\mathcal{O}(-O_{E}). So the above exact sequence splits, and we have

WNN,W\cong N^{\prime}\oplus N^{\prime},

where both summands are Lagrangian. In particular, WW and hence ξL\xi_{L} carries a faithful action of Sp2Sp_{2}, so dimAut(ξL)>2\dim\mathrm{Aut}(\xi_{L})>2 as claimed.

In type CC, we have that every nonzero morphism γ:𝒪(2OE)W\gamma\colon\mathcal{O}(-2O_{E})\to W must vanish at some unique point xγEsx_{\gamma}\in E_{s}. So again we have a morphism k1=H0(Es,W𝒪(2OE))Es\mathbb{P}^{1}_{k}=\mathbb{P}H^{0}(E_{s},W\otimes\mathcal{O}(-2O_{E}))\to E_{s} sending γ\gamma to xγx_{\gamma}, which must be constant. So xγ=xx_{\gamma}=x is independent of γ\gamma, and every morphism 𝒪(2OE)W\mathcal{O}(-2O_{E})\to W therefore factors through a subbundle 𝒪(x2OE)W\mathcal{O}(x-2O_{E})\subseteq W. Since WW is semistable of trivial determinant, choosing any two linearly independent morphisms gives an isomorphism W𝒪(x2OE)𝒪(x2OE)W\cong\mathcal{O}(x-2O_{E})\oplus\mathcal{O}(x-2O_{E}). So SL2SL_{2} acts faithfully on WW and hence on ξL\xi_{L} and dimAut(ξL)>2\dim\mathrm{Aut}(\xi_{L})>2 as claimed.

In type DD, we have that Uγ=W/𝒪(OE)U_{\gamma}=W/\mathcal{O}(-O_{E}) is unstable for every nonzero morphism γ:𝒪(OE)W\gamma\colon\mathcal{O}(-O_{E})\to W. (Note that again any such γ\gamma must be a subbundle since WW is semistable of slope 1/2-1/2.) Since WW is semistable, one sees that the Harder-Narasimhan decomposition of UγU_{\gamma} must be of the form Uγ=Nγdet(Nγ)𝒪(OE)U_{\gamma}=N_{\gamma}\oplus\det(N_{\gamma})^{\vee}\otimes\mathcal{O}(-OE), where NγN_{\gamma} is a rank 22 semistable vector bundle of degree 1-1. Again we get a morphism k1=H0(Es,W𝒪(OE))Pic1(Es)\mathbb{P}^{1}_{k}=\mathbb{P}H^{0}(E_{s},W\otimes\mathcal{O}(O_{E}))\to\mathrm{Pic}^{-1}(E_{s}) sending γ\gamma to the isomorphism class of det(Nγ)\det(N_{\gamma}), which again must be constant. So det(Nγ)\det(N_{\gamma}), and hence Nγ=NN_{\gamma}=N are independent of γ\gamma, and every nonzero morphism 𝒪(OE)W\mathcal{O}(-O_{E})\to W factors through the kernel of some surjection WNW\to N. Choosing two linearly independent morphisms 𝒪(OE)W\mathcal{O}(-O_{E})\to W therefore gives a map WNNW\to N\oplus N, which one easily sees must be an isomorphism. So again SL2SL_{2} acts faithfully on WW fixing the determinant, and hence on ξL\xi_{L}, which proves that dimAut(ξL)>2\dim\mathrm{Aut}(\xi_{L})>2 in this case as well.

Finally, to prove (4), simply note that Proposition 2.2.3 implies that ZZ0Z\to Z_{0} is an affine space bundle of relative dimension 2ρ,μ=l+2-\langle 2\rho,\mu\rangle=l+2, so ZSZ\to S has relative dimension l+3l+3 as required. ∎

3. Computing resolutions

The purpose of this section is to give the proof of Theorem 1.0.3. We prove (1), (2), (3) and (4) separately (as Propositions 3.1.1, 3.4.1, 3.5.1 and 3.6.1) in §3.1, §3.4, §3.5 and §3.6 respectively.

The proofs of Propositions 3.4.1 and 3.6.1 make use of the idea that sections of a flag variety bundle decompose naturally according to which Bruhat cells they meet. This is used to give decompositions of the divisor Dαi(Z)D_{\alpha_{i}^{\vee}}(Z) and Dαj(Z)D_{\alpha_{j}^{\vee}}(Z) into locally closed subsets, each of which can be identified in terms of an analogous set of sections of a flag variety for some copy of GLnGL_{n} inside GG. We manage to show that these “Bruhat cells” fit together into the blowups in Theorem 1.0.3 by explicitly constructing the blow downs as spaces of (stable) sections of partial flag variety bundles. The Bruhat cells are discussed in general in §3.2, and the specific cells of interest for GLnGL_{n} are studied in §3.3.

3.1. Decomposition of χ~Z1(0ΘY1)\tilde{\chi}_{Z}^{-1}(0_{\Theta_{Y}^{-1}})

In this subsection, we prove the following proposition, which is a slightly more general version of part (1) of Theorem 1.0.3.

Proposition 3.1.1.

Let (G,P,μ)(G,P,\mu) be a subregular Harder-Narasimhan class not of type A1A_{1}. Assume that the μd\mu_{d}-gerbe 𝔊\mathfrak{G} of Theorem 2.2.6 is trivial, let Z0BunL,rigss,μZ_{0}\to\mathrm{Bun}_{L,rig}^{ss,\mu} be the corresponding Θ\Theta-trivial slice, and let Z=IndLG(Z0)BunG,rigZ=\mathrm{Ind}_{L}^{G}(Z_{0})\to\mathrm{Bun}_{G,rig} be the induced equivariant slice. Then the preimage of the zero section of ΘY1\Theta_{Y}^{-1} in Z~=Bun~G,rig×BunGZ\tilde{Z}=\widetilde{\mathrm{Bun}}_{G,rig}\times_{\mathrm{Bun}_{G}}Z decomposes as a divisor with normal crossings

χ~Z1(0ΘY1)=dDαi(Z)+Dαj(Z)+Dαi+αj(Z)\tilde{\chi}_{Z}^{-1}(0_{\Theta_{Y}^{-1}})=dD_{\alpha_{i}^{\vee}}(Z)+D_{\alpha_{j}^{\vee}}(Z)+D_{\alpha_{i}^{\vee}+\alpha_{j}^{\vee}}(Z) (3.1.1)

such that each summand is smooth over YY, where Dλ(Z)D_{\lambda}(Z) denotes the closure of the locus of stable maps with a single rational component of degree λ𝕏(T)\lambda\in\mathbb{X}_{*}(T).

Proof.

Since ZBunG,rigZ\to\mathrm{Bun}_{G,rig} is a slice, [davis19, Proposition 2.1.10 and Corollary 3.3.8] imply that the preimage of the zero section decomposes as a divisor with normal crossings

χ~Z1(0ΘY1)=λ𝕏(T)+12(λ|λ)Dλ(Z),\tilde{\chi}_{Z}^{-1}(0_{\Theta_{Y}^{-1}})=\sum_{\lambda\in\mathbb{X}_{*}(T)_{+}}\frac{1}{2}(\lambda{\,|\,}\lambda)D_{\lambda}(Z),

where (|)(\,{\,|\,}\,) is the normalised Killing form on 𝕏(T)\mathbb{X}_{*}(T). By Lemma 3.1.2 below, Dλ(Z)=D_{\lambda}(Z)=\emptyset unless λ{αi,αj,αi+αj}\lambda\in\{\alpha_{i}^{\vee},\alpha_{j}^{\vee},\alpha_{i}^{\vee}+\alpha_{j}^{\vee}\}, so this simplifies to (3.1.1) as required, since αj\alpha_{j}^{\vee} and αi+αj\alpha_{i}^{\vee}+\alpha_{j}^{\vee} are short coroots and (αi|αi)=2d(\alpha_{i}^{\vee}{\,|\,}\alpha_{i}^{\vee})=2d in each case.

It remains to show that each Dλ(Z)D_{\lambda}(Z) is smooth over YY. Note that χ~Z1(0ΘY1)\tilde{\chi}_{Z}^{-1}(0_{\Theta_{Y}^{-1}}) is in fact a divisor with normal crossings relative to YY: this follows from [davis19a, Proposition 3.5.3], the definition [davis19, Definition 2.1.14] of the blow down morphism Bun~GY\widetilde{\mathrm{Bun}}_{G}\to Y, and the fact that the boundary of the stack 𝔇egS(E){\mathfrak{D}\mathrm{eg}}_{S}(E) of prestable degenerations of EE is a divisor with normal crossings relative to SS [davis19, Proposition 2.1.7]. So it is enough to show that each Dλ(Z)D_{\lambda}(Z) has no self-intersections. But a point in such a self-intersection would have to be given by a stable map with at least two rational components both of degree αi\geq\alpha_{i}^{\vee} or both of degree αj\geq\alpha_{j}^{\vee}. But this is forbidden by Lemma 3.1.2, so we are done. ∎

Lemma 3.1.2.

For λ𝕏(T)+\lambda\in\mathbb{X}_{*}(T)_{+}, we have Dλ(Z)D_{\lambda}(Z)\neq\emptyset if and only if λ{αi,αj,αi+αj}\lambda\in\{\alpha_{i}^{\vee},\alpha_{j}^{\vee},\alpha_{i}^{\vee}+\alpha_{j}^{\vee}\}.

Proof.

For simplicity, we can assume without loss of generality that S=SpeckS=\operatorname{\mathrm{Spec}}k for kk an algebraically closed field. We first show that Dαi(Z)D_{\alpha_{i}^{\vee}}(Z)\neq\emptyset and Dαj(Z)D_{\alpha_{j}^{\vee}}(Z)\neq\emptyset.

If (G,P,μ)(G,P,\mu) is of type AA, then μ\mu is the image of αiαj-\alpha_{i}^{\vee}-\alpha_{j}^{\vee} under the homomorphism 𝕏(T)𝕏(Z(L))\mathbb{X}_{*}(T)\to\mathbb{X}_{*}(Z(L)^{\circ})_{\mathbb{Q}} and α,αi+αj0\langle\alpha,\alpha_{i}^{\vee}+\alpha_{j}^{\vee}\rangle\leq 0 for all αΦ+\alpha\in\Phi_{+} a root of PP. So by [davis19a, Proposition 3.6.4], the morphism

KMB,GαiαjKMP,Gμ\mathrm{KM}_{B,G}^{-\alpha_{i}^{\vee}-\alpha_{j}^{\vee}}\longrightarrow\mathrm{KM}_{P,G}^{\mu}

is surjective. In particular, for every zZ0z\in Z_{0}, there exists a section of ξL,z×LP/BξL,z×LG/B\xi_{L,z}\times^{L}P/B\subseteq\xi_{L,z}\times^{L}G/B with degree λ0αiαj-\lambda_{0}\leq-\alpha_{i}^{\vee}-\alpha_{j}^{\vee}. So we must have Dλ0(Z)D_{\lambda_{0}}(Z)\neq\emptyset, and hence Dαi(Z)D_{\alpha_{i}^{\vee}}(Z)\neq\emptyset and Dαj(Z)D_{\alpha_{j}^{\vee}}(Z)\neq\emptyset, since we can always add rational tails to such a section to produce a stable map in each of these divisors.

On the other hand, if (G,P,μ)(G,P,\mu) is not of type AA, then μ\mu is the image of αi-\alpha_{i}^{\vee} in 𝕏(Z(L))\mathbb{X}_{*}(Z(L)^{\circ})_{\mathbb{Q}}, and α,αi0\langle\alpha,\alpha_{i}^{\vee}\rangle\leq 0 for αΦ+\alpha\in\Phi_{+} a root of PP. So

KMB,GαiKMP,Gμ\mathrm{KM}_{B,G}^{-\alpha_{i}^{\vee}}\longrightarrow\mathrm{KM}_{P,G}^{\mu}

is surjective by [davis19a, Proposition 3.6.4], so we deduce that Dαi(Z)D_{\alpha_{i}^{\vee}}(Z)\neq\emptyset. For Dαj(Z)D_{\alpha_{j}^{\vee}}(Z), note that since αjΔ\alpha_{j}\in\Delta is the unique special root, [davis19, Proposition 4.2.3] implies that the Harder-Narasimhan locus BunQss,αjBunG\mathrm{Bun}_{Q}^{ss,-\alpha_{j}^{\vee}}\subseteq\mathrm{Bun}_{G} is dense in the locus of unstable GG-bundles, where QQ is the standard parabolic with t(Q)={αj}t(Q)=\{\alpha_{j}^{\vee}\}. So BunQ,rigss,αj×BunG,rigZ\mathrm{Bun}_{Q,rig}^{ss,-\alpha_{j}^{\vee}}\times_{\mathrm{Bun}_{G,rig}}Z\neq\emptyset, and hence Dαj(Z)D_{\alpha_{j}^{\vee}}(Z)\neq\emptyset by [davis19, Proposition 4.3.8].

Conversely, suppose that λ𝕏(T)\lambda\in\mathbb{X}_{*}(T) and that Dλ(Z)D_{\lambda}(Z)\neq\emptyset. Then for any αkΔ\alpha_{k}\in\Delta with corresponding maximal parabolic PkP_{k}, there exists a point in ZZ and a section of the corresponding G/PkG/P_{k}-bundle with degree νk=ϖk,λ/ϖk,ϖkϖk\nu_{k}=-\langle\varpi_{k},\lambda\rangle/\langle\varpi_{k},\varpi_{k}^{\vee}\rangle\varpi_{k}^{\vee} (the image of λ\lambda in 𝕏(TPk)\mathbb{X}_{*}(T_{P_{k}})). So by Lemma 3.1.3 and [friedman-morgan00, Lemma 3.3.2], we must have

(l+1)ϖk,λ2ρ,ϖkϖk,ϖkϖk,λ=2ρ,νk2ρ,μl+3.(l+1)\langle\varpi_{k},\lambda\rangle\leq\frac{\langle 2\rho,\varpi_{k}^{\vee}\rangle}{\langle\varpi_{k},\varpi_{k}^{\vee}\rangle}\langle\varpi_{k},\lambda\rangle=-\langle 2\rho,\nu_{k}\rangle\leq-\langle 2\rho,\mu\rangle\leq l+3.

So

ϖk,λl+3l+1<2,\langle\varpi_{k},\lambda\rangle\leq\frac{l+3}{l+1}<2,

since l>1l>1. So ϖk,λ=0\langle\varpi_{k},\lambda\rangle=0 or 11 for all kk.

Now assume for a contradiction that there exists λ𝕏(T)+{αi,αj,αi+αj}\lambda\in\mathbb{X}_{*}(T)_{+}\setminus\{\alpha_{i}^{\vee},\alpha_{j}^{\vee},\alpha_{i}^{\vee}+\alpha_{j}^{\vee}\} such that Dλ(Z)D_{\lambda}(Z)\neq\emptyset. Since the divisor D(Z)=χ~Z1(0ΘY1)D(Z)=\tilde{\chi}_{Z}^{-1}(0_{\Theta_{Y}^{-1}}) is connected by Lemma 3.1.4 below, we can choose λ\lambda so that Dλ(Z)D_{\lambda}(Z) has nonempty intersection with one of Dαi(Z)D_{\alpha_{i}^{\vee}}(Z), Dαj(Z)D_{\alpha_{j}^{\vee}}(Z) or Dαi+αj(Z)D_{\alpha_{i}^{\vee}+\alpha_{j}^{\vee}}(Z). Choose a point in such an intersection over zZz\in Z, and let λ𝕏(T)-\lambda^{\prime}\in\mathbb{X}_{*}(T)_{-} denote the degree of the corresponding stable map restricted to the irreducible component of genus 11. Then we have Dλ(Z)D_{\lambda^{\prime}}(Z)\neq\emptyset, λλ\lambda^{\prime}\geq\lambda and λαr\lambda^{\prime}\geq\alpha_{r}^{\vee} for some αr{αi,αj}\alpha_{r}\in\{\alpha_{i},\alpha_{j}\}. By the bound proved above, we must have ϖk,λ=1\langle\varpi_{k},\lambda\rangle=1 for some αkΔ{αi,αj}\alpha_{k}\in\Delta\setminus\{\alpha_{i},\alpha_{j}\}, and hence λαr+αk\lambda^{\prime}\geq\alpha_{r}^{\vee}+\alpha_{k}^{\vee}. So adding rational tails to the degree λ-\lambda^{\prime} section if necessary, we deduce Dαr+αk(Z)D_{\alpha_{r}^{\vee}+\alpha_{k}^{\vee}}(Z)\neq\emptyset and Dαk(Z)D_{\alpha_{k}^{\vee}}(Z)\neq\emptyset.

Assume first that GG is not of type AA. Since Dαk(Z)D_{\alpha_{k}^{\vee}}(Z)\neq\emptyset, there exists zZz\in Z and a section of ξG,z/B\xi_{G,z}/B with degree αk-\alpha_{k}^{\vee}, and hence a section of ξG,z/Pk\xi_{G,z}/P_{k} with slope ϖk/ϖk,ϖk-\varpi_{k}^{\vee}/\langle\varpi_{k},\varpi_{k}^{\vee}\rangle. So by Lemma 3.1.3, there exists zZz^{\prime}\in Z such that ξG,z\xi_{G,z^{\prime}} has Harder-Narasimhan reduction to PkP_{k} with slope ϖk/ϖk,ϖk-\varpi_{k}^{\vee}/\langle\varpi_{k},\varpi_{k}^{\vee}\rangle. Since PkPP_{k}\neq P, we have zZZ0z^{\prime}\in Z\setminus Z_{0} so in particular zz is not fixed under the Z(L)rigZ(L)_{rig}-action. Comparing codimensions in BunG,rig/E\mathrm{Bun}_{G,rig}/E and in ZZ, we deduce that ξG,z\xi_{G,z^{\prime}} must be regular unstable, which is a contradiction since it has the wrong Harder-Narasimhan type, as αk\alpha_{k} is not a special root.

Assume on the other hand that GG is of type AA. We have k{i,i+1}k\notin\{i,i+1\} and r{i,i+1}r\in\{i,i+1\} such that Dαr+αk(Z)D_{\alpha_{r}^{\vee}+\alpha_{k}^{\vee}}(Z)\neq\emptyset. So there exists zZz\in Z and a section of ξG,z/Pr,k\xi_{G,z}/P_{r,k} of slope ν𝕏(Z(Lr,k))\nu\in\mathbb{X}_{*}(Z(L_{r,k})^{\circ})_{\mathbb{Q}} satisfying ϖr,ν=ϖk,ν=1\langle\varpi_{r},\nu\rangle=\langle\varpi_{k},\nu\rangle=-1, where Pr,kGP_{r,k}\subseteq G is the standard parabolic of type {αr,αk}\{\alpha_{r},\alpha_{k}\} and Lr,kL_{r,k} its standard Levi factor. But ν\nu is a Harder-Narasimhan vector for Pr,kP_{r,k}, so by Lemma 3.1.3, there exists zZz^{\prime}\in Z such that ξG,z\xi_{G,z^{\prime}} has Harder-Narasimhan reduction to Pr,kP_{r,k} with slope ν\nu. Since Pr,kPP_{r,k}\neq P, we have zZZ0z\in Z\setminus Z_{0}. Again this implies that ξG,z\xi_{G,z^{\prime}} is regular unstable, giving a contradiction.

So Dλ(Z)=D_{\lambda}(Z)=\emptyset for λ{αi,αj,αi+αj}\lambda\notin\{\alpha_{i}^{\vee},\alpha_{j}^{\vee},\alpha_{i}^{\vee}+\alpha_{j}^{\vee}\}, and Dαi(Z),Dαj(Z)D_{\alpha_{i}^{\vee}}(Z),D_{\alpha_{j}^{\vee}}(Z)\neq\emptyset. This implies that Dαi+αj(Z)D_{\alpha_{i}^{\vee}+\alpha_{j}^{\vee}}(Z)\neq\emptyset, for if this were not the case, we would have Dαi(Z)Dαj(Z)=D_{\alpha_{i}^{\vee}}(Z)\cap D_{\alpha_{j}^{\vee}}(Z)=\emptyset and hence χ~Z1(0ΘY1)\tilde{\chi}_{Z}^{-1}(0_{\Theta_{Y}^{-1}}) would be disconnected, contradicting Lemma 3.1.4. ∎

Lemma 3.1.3.

In the setup of Proposition 3.1.1, fix some zZz\in Z with corresponding GG-bundle ξG,z\xi_{G,z}. If there exists a section of ξG,z/Q\xi_{G,z}/Q of degree ν\nu, where QQ is any standard parabolic with Harder-Narasimhan vector ν\nu and (Q,ν)(P,μ)(Q,\nu)\neq(P,\mu), then

  1. (1)

    there exists zZz^{\prime}\in Z such that the corresponding GG-bundle ξG,z\xi_{G,z^{\prime}} has Harder-Narasimhan reduction to QQ with degree ν\nu, and

  2. (2)

    2ρ,νl+2-\langle 2\rho,\nu\rangle\leq l+2.

Proof.

The assumptions imply that the stack Z×BunG,rigBunQ,rigνZ\times_{\mathrm{Bun}_{G,rig}}\mathrm{Bun}_{Q,rig}^{\nu} is nonempty. Since ZBunG,rig/EZ\to\mathrm{Bun}_{G,rig}/E is smooth, the preimage Z×BunG,rigBunQ,rigss,νZ\times_{\mathrm{Bun}_{G,rig}}\mathrm{Bun}_{Q,rig}^{ss,\nu} of BunQ,rigss,ν/E\mathrm{Bun}_{Q,rig}^{ss,\nu}/E under the morphism

Z×BunG,rigBunQ,rigν=Z×BunG,rig/EBunQ,rig/EBunQ,rigν/EZ\times_{\mathrm{Bun}_{G,rig}}\mathrm{Bun}_{Q,rig}^{\nu}=Z\times_{\mathrm{Bun}_{G,rig}/E}\mathrm{Bun}_{Q,rig}/E\longrightarrow\mathrm{Bun}_{Q,rig}^{\nu}/E

is dense, hence nonempty. This proves (1). Since (Q,ν)(P,μ)(Q,\nu)\neq(P,\mu), by uniqueness of Harder-Narasimhan reductions, the Z(L)rigZ(L)_{rig}-invariant locally closed substack Z×BunG,rigBunQ,rigss,νZZ\times_{\mathrm{Bun}_{G,rig}}\mathrm{Bun}_{Q,rig}^{ss,\nu}\subseteq Z is disjoint from the Z(L)rigZ(L)_{rig}-fixed locus Z0ZZ_{0}\subseteq Z. Since Z0SZ_{0}\to S has finite relative stabilisers, Z×BunG,rigBunQ,rigνSZ\times_{\mathrm{Bun}_{G,rig}}\mathrm{Bun}_{Q,rig}^{\nu}\to S is therefore flat of relative dimension at least 11, and hence has codimension at most

dimSZ1=l+2.\dim_{S}Z-1=l+2.

But this codimension is equal to the codimension 2ρ,ν-\langle 2\rho,\nu\rangle of BunQ,rigss,ν/E\mathrm{Bun}_{Q,rig}^{ss,\nu}/E in BunG,rig/E\mathrm{Bun}_{G,rig}/E, so (2) follows. ∎

Lemma 3.1.4.

The morphisms

Bun~GBunG×(Y^//W)/𝔾mΘY1/𝔾m\widetilde{\mathrm{Bun}}_{G}\longrightarrow\mathrm{Bun}_{G}\times_{(\widehat{Y}{/\mkern-6.0mu/}W)/\mathbb{G}_{m}}\Theta_{Y}^{-1}/\mathbb{G}_{m} (3.1.2)

and

χ~Z1(0ΘY1)Y\tilde{\chi}_{Z}^{-1}(0_{\Theta_{Y}^{-1}})\longrightarrow Y (3.1.3)

have connected fibres.

Proof.

Note that the target of (3.1.2) is a local complete intersection, hence Cohen-Macaulay. Moreover, (3.1.2) is an isomorphism over the open substack BunGreg×(Y^//W)/𝔾mΘY1/𝔾m\mathrm{Bun}_{G}^{reg}\times_{(\widehat{Y}{/\mkern-6.0mu/}W)/\mathbb{G}_{m}}\Theta_{Y}^{-1}/\mathbb{G}_{m}, where BunGregBunG\mathrm{Bun}_{G}^{reg}\subseteq\mathrm{Bun}_{G} is the open substack of regular bundles [davis19, §4.4]. This open substack is big (i.e., the complement has codimension at least 22) by [davis19, Proposition 4.4.6], so the target is normal and the pushforward of 𝒪\mathcal{O} is 𝒪\mathcal{O}. Connectedness of the fibres now follows from Zariski’s connectedness theorem [olsson07, Theorem 11.3].

We can write (3.1.3) as a composition

χ~Z1(0ΘY1)χZ1(0)×S0ΘY10ΘY1=Y.\tilde{\chi}_{Z}^{-1}(0_{\Theta_{Y}^{-1}})\longrightarrow\chi_{Z}^{-1}(0)\times_{S}0_{\Theta_{Y}^{-1}}\longrightarrow 0_{\Theta_{Y}^{-1}}=Y. (3.1.4)

The first factor is a pullback of (3.1.2), so has connected fibres. The morphism χZ1(0)S\chi_{Z}^{-1}(0)\to S also has connected fibres, since the H=Z(L)rigH=Z(L)_{rig}-action contracts χZ1(0)\chi_{Z}^{-1}(0) onto Z0Z_{0} and Z0SZ_{0}\to S has connected fibres. So both factors of (3.1.4) have connected fibres, and the first is proper, so their composition has connected fibres also. ∎

3.2. Digression: Bruhat cells for PP-bundles

In this subsection, we consider GG an arbitrary reductive group. (The examples of interest will be our simply connected simple group GG from the rest of the paper, and G=GLnG=GL_{n}.) The material presented here is a brief recap of [davis19a, §3.7].

Given two parabolic subgroups P,PGP,P^{\prime}\subseteq G, which we may as well assume standard, for each ww in the Weyl group WW of GG, there is an associated Bruhat cell

CP,PwBunP×BunGBunP.C^{w}_{P,P^{\prime}}\subseteq\mathrm{Bun}_{P}\times_{\mathrm{Bun}_{G}}\mathrm{Bun}_{P^{\prime}}.

Thinking of the stack on the right as the stack of pairs (ξP,σ)(\xi_{P},\sigma), where ξP\xi_{P} is a PP-bundle and σ\sigma is a section of the partial flag variety bundle ξP×PG/P\xi_{P}\times^{P}G/P^{\prime}, we can define CP,PwC^{w}_{P,P^{\prime}} as the locally closed substack of pairs such that σ\sigma factors through the Bruhat cell ξP×PPwP/P\xi_{P}\times^{P}PwP^{\prime}/P^{\prime}. Since PwP/PP/PwPw1PwP^{\prime}/P^{\prime}\cong P/P\cap wP^{\prime}w^{-1} by the orbit-stabiliser theorem, we have

CP,PwBunPwPw1.C^{w}_{P,P^{\prime}}\cong\mathrm{Bun}_{P\cap wP^{\prime}w^{-1}}.

Under this identification, the map CP,PwBunPC^{w}_{P,P^{\prime}}\to\mathrm{Bun}_{P} (resp. CP,PwBunPC^{w}_{P,P^{\prime}}\to\mathrm{Bun}_{P^{\prime}}) sends a PwPw1P\cap wP^{\prime}w^{-1}-bundle of degree μ𝕏(TPwPw1)\mu\in\mathbb{X}_{*}(T_{P\cap wP^{\prime}w^{-1}}) to a PP-bundle of degree iw(μ)i_{w}(\mu) (resp. a PP^{\prime}-bundle of degree jw(μ)j_{w}(\mu)), where iw:𝕏(TPwPw1)𝕏(TP)i_{w}\colon\mathbb{X}_{*}(T_{P\cap wP^{\prime}w^{-1}})\to\mathbb{X}_{*}(T_{P}) and jw:𝕏(TPwPw1)𝕏(TP)j_{w}\colon\mathbb{X}_{*}(T_{P\cap wP^{\prime}w^{-1}})\to\mathbb{X}_{*}(T_{P^{\prime}}) are induced by the two inclusions

P←-⸧PwPw1⸦-w-1(-)w→P.P\leftarrow\joinrel\relbar\joinrel\rhook P\cap wP^{\prime}w^{-1}{\lhook\joinrel\relbar\joinrel\xrightarrow{w^{-1}(-)w}}P^{\prime}.

The cell CP,PwC^{w}_{P,P^{\prime}} depends only on the double coset WPwWPW_{P}wW_{P^{\prime}}, where WPW_{P} and WPW_{P^{\prime}} are the Weyl groups of (the Levi subgroups of) PP and PP^{\prime}. A particularly nice choice of double coset representatives is given by

WP,P0={wWw1αiΦ+andwαjΦ+forαiΔt(P)andαjΔt(P)}.W^{0}_{P,P^{\prime}}=\{w\in W\mid w^{-1}\alpha_{i}\in\Phi_{+}\;\text{and}\;w\alpha_{j}\in\Phi_{+}\;\text{for}\;\alpha_{i}\in\Delta\setminus t(P)\;\text{and}\;\alpha_{j}\in\Delta\setminus t(P^{\prime})\}. (3.2.1)

These are the coset representatives of minimal length. Note that WPW_{P} (resp., WPW_{P^{\prime}}) are generated by the simple reflections sis_{i} in the roots αiΔt(P)\alpha_{i}\in\Delta\setminus t(P) (resp., αiΔt(P)\alpha_{i}\in\Delta\setminus t(P^{\prime})).

When we want to keep track of the associated PP-bundle and the degree of the associated PP^{\prime}-bundle, we write

CP,Pw,λ=CP,Pw×BunPBunPλandCP,P,ξPw,λ={ξP}×BunPCP,Pw,λC^{w,\lambda}_{P,P^{\prime}}=C^{w}_{P,P^{\prime}}\times_{\mathrm{Bun}_{P^{\prime}}}\mathrm{Bun}_{P^{\prime}}^{\lambda}\quad\text{and}\quad C^{w,\lambda}_{P,P^{\prime},\xi_{P}}=\{\xi_{P}\}\times_{\mathrm{Bun}_{P}}C^{w,\lambda}_{P,P^{\prime}}

for λ𝕏(TP)\lambda\in\mathbb{X}_{*}(T_{P^{\prime}}) and ξPBunP\xi_{P}\in\mathrm{Bun}_{P}.

The following proposition gives an extremely useful criterion for determining when the Bruhat cells cover a given fibre of BunP×BunGBunPλBunP\mathrm{Bun}_{P}\times_{\mathrm{Bun}_{G}}\mathrm{Bun}_{P^{\prime}}^{\lambda}\to\mathrm{Bun}_{P}.

Proposition 3.2.1 ([davis19a, Proposition 3.7.6]).

Let ξPEs\xi_{P}\to E_{s} be a PP-bundle on a geometric fibre of ESE\to S and suppose that there exists a point in BunP×BunGBunPλ\mathrm{Bun}_{P}\times_{\mathrm{Bun}_{G}}\mathrm{Bun}_{P^{\prime}}^{\lambda} over ξP\xi_{P} that does not lie in any Bruhat cell. Then there exists wWP,P0{1}w\in W^{0}_{P,P^{\prime}}\setminus\{1\} and λ<λ\lambda^{\prime}<\lambda such that CP,P,ξL×LPw,λC^{w,\lambda^{\prime}}_{P,P^{\prime},\xi_{L}\times^{L}P}\neq\emptyset, where LL is the Levi factor of PP and ξL=ξP×PL\xi_{L}=\xi_{P}\times^{P}L is the associated LL-bundle.

3.3. Digression: some Bruhat cells for unstable vector bundles

The aim of this subsection is to describe certain spaces of stable maps to partial flag variety bundles associated to particular minimally unstable GLnGL_{n}-bundles on EE. The spaces considered here will crop up again and again in the subregular part of the elliptic Grothendieck-Springer resolution Bun~G\widetilde{\mathrm{Bun}}_{G}.

Recall the notation for the root datum of GLnGL_{n} and the standard parabolic subgroups QknGLnQ^{n}_{k}\subseteq GL_{n} given in §1.2 We also consider the standard parabolic subgroup

Rn\displaystyle R_{n} ={(ap,q)1p,qnGLnap,q=0forq>max(p,n1)}={(00)}\displaystyle=\{(a_{p,q})_{1\leq p,q\leq n}\in GL_{n}\mid a_{p,q}=0\;\text{for}\;q>\mathrm{max}(p,n-1)\}=\left\{\left(\begin{matrix}*&*&\cdots&*&0\\ \vdots&\vdots&&\vdots&\vdots\\ *&*&\cdots&*&0\\ *&*&\cdots&*&*\end{matrix}\right)\right\}

of type {βn1}\{\beta_{n-1}\}. For 1kn1\leq k\leq n, let

Xkn=YQnnen×YQknenKMQkn,GLnen×BunGLn1BunRnss,e1,X_{k}^{n}=Y_{Q^{n}_{n}}^{-e_{n}^{*}}\times_{Y_{Q^{n}_{k}}^{-e_{n}^{*}}}\mathrm{KM}^{-e_{n}^{*}}_{Q^{n}_{k},GL_{n}}\times_{\mathrm{Bun}_{GL_{n}}^{-1}}\mathrm{Bun}_{R_{n}}^{ss,-e_{1}^{*}},

where, for any standard parabolic subgroup PQnnP\supseteq Q^{n}_{n}, we use the same notation for a cocharacter λ𝕏(Qnn)\lambda\in\mathbb{X}_{*}(Q^{n}_{n}) and for its image in 𝕏(P/[P,P])\mathbb{X}_{*}(P/[P,P]). In words, a point of the stack XknX_{k}^{n} over sSs\in S consists of a tuple (y,σ:CξRn×RnGLn/Qkn,ξRn)(y,\sigma\colon C\to\xi_{R_{n}}\times^{R_{n}}GL_{n}/Q^{n}_{k},\xi_{R_{n}}), where ξRnEs\xi_{R_{n}}\to E_{s} is a semistable RnR_{n}-bundle of degree e1-e_{1}^{*} (which is the Harder-Narasimhan reduction of the unstable GLnGL_{n}-bundle of the subsection title), σ\sigma is a stable section of degree en-e_{n}^{*}, and yy is a lift of (the isomorphism class of) the associated TQknT_{Q^{n}_{k}}-bundle to a TQnnT_{Q^{n}_{n}}-bundle of degree en-e_{n}^{*}.

For 1pn11\leq p\leq n-1, let wpWGLn=Snw_{p}\in W_{GL_{n}}=S_{n} be the cyclic permutation

wp=(n,n1,,p+1,p)=sn1sn2spw_{p}=(n,n-1,\ldots,p+1,p)=s_{n-1}s_{n-2}\cdots s_{p}

and let wn=1w_{n}=1 be the identity, where WGLnW_{GL_{n}} is the Weyl group of GLnGL_{n}, and si=(i,i+1)s_{i}=(i,i+1) is the reflection in the root βi\beta_{i}. For 1p,kn1\leq p,k\leq n, we write Ck,pGLnXknC_{k,p}^{GL_{n}}\subseteq X^{n}_{k} for the locally closed substack of tuples (y,σ,ξRn)(y,\sigma,\xi_{R_{n}}) such that the restriction of σ\sigma to the genus 11 component factors through the Bruhat cell

ξRn×RnRnwpQkn/QknξRn×RnGLn/Qkn.\xi_{R_{n}}\times^{R_{n}}R_{n}w_{p}Q^{n}_{k}/Q^{n}_{k}\subseteq\xi_{R_{n}}\times^{R_{n}}GL_{n}/Q^{n}_{k}.
Proposition 3.3.1.

For 1kn1\leq k\leq n, there is a decomposition

Xkn=1p<kCk,pGLnCk,nGLnX_{k}^{n}=\bigcup_{1\leq p<k}C_{k,p}^{GL_{n}}\cup C_{k,n}^{GL_{n}}

into disjoint locally closed substacks.

We break the proof of Proposition 3.3.1 into several lemmas.

Lemma 3.3.2.

Assume that ξRnEs\xi_{R_{n}}\to E_{s} is a semistable RnR_{n}-bundle on a geometric fibre of ESE\to S of degree e1-e_{1}^{*} and that σ:EsξRn×RnGLn/Qnn\sigma\colon E_{s}\to\xi_{R_{n}}\times^{R_{n}}GL_{n}/Q^{n}_{n} is a section of degree λen\lambda\leq-e_{n}^{*}. Then λ{en,en1}\lambda\in\{-e_{n}^{*},-e_{n-1}^{*}\}.

Proof.

The section σ\sigma corresponds to a complete flag

0=VnVn1V0=V,0=V_{n}\subsetneq V_{n-1}\subsetneq\cdots\subsetneq V_{0}=V,

where VV is the vector bundle associated to the GLnGL_{n}-bundle ξGLn=ξRn×RnGLn\xi_{GL_{n}}=\xi_{R_{n}}\times^{R_{n}}GL_{n}, such that Vi1/ViV_{i-1}/V_{i} is a line bundle of degree ei,λ\langle e_{i},\lambda\rangle for i=1,,ni=1,\ldots,n. Since ξRn\xi_{R_{n}} is the Harder-Narasimhan reduction of ξGLn\xi_{GL_{n}}, VV has Harder-Narasimhan decomposition V=MUV=M\oplus U, where UU is a semistable vector bundle of rank n1n-1 and degree 1-1 and MM is a line bundle of degree 0. In particular, any quotient bundle of VV has slope 1/(n1)\geq-1/(n-1), so we deduce that

e1++ei,λ=degV/Viin1\langle e_{1}+\cdots+e_{i},\lambda\rangle=\deg V/V_{i}\geq\frac{-i}{n-1} (3.3.1)

for i=1,,n1i=1,\ldots,n-1.

Since λen\lambda\leq-e_{n}^{*} by assumption, we have

λ=eni=1n1diβi\lambda=-e_{n}^{*}-\sum_{i=1}^{n-1}d_{i}\beta_{i}^{\vee}

for some di0d_{i}\in\mathbb{Z}_{\geq 0}, where βi=eiei+1\beta_{i}^{\vee}=e_{i}^{*}-e_{i+1}^{*}. Applying (3.3.1), we have di=0d_{i}=0 for 1in21\leq i\leq n-2 and dn1{0,1}d_{n-1}\in\{0,1\}, which implies the lemma. ∎

In what follows, we will write

Ckw,λ=BunRnss,e1×BunRnCRn,Qknw,λBunRn×BunGLnBunQknC^{w,\lambda}_{k}=\mathrm{Bun}_{R_{n}}^{ss,-e_{1}^{*}}\times_{\mathrm{Bun}_{R_{n}}}C^{w,\lambda}_{R_{n},Q^{n}_{k}}\subseteq\mathrm{Bun}_{R_{n}}\times_{\mathrm{Bun}_{GL_{n}}}\mathrm{Bun}_{Q^{n}_{k}}

for wWRn,Qkn0w\in W^{0}_{R_{n},Q^{n}_{k}} and λ𝕏(TQkn)\lambda\in\mathbb{X}_{*}(T_{Q^{n}_{k}}). Here CRn,Qknw,λC^{w,\lambda}_{R_{n},Q^{n}_{k}} is the Bruhat cell of §3.2.

Lemma 3.3.3.

Assume that wWRn,Qnn0w\in W^{0}_{R_{n},Q^{n}_{n}} and λ𝕏(TQnn)\lambda\in\mathbb{X}_{*}(T_{Q^{n}_{n}}) with Cnw,λC^{w,\lambda}_{n}\neq\emptyset and λen\lambda\leq-e_{n}^{*}. Then

(w,λ){(1,en1)}{(wp,en)1p<n}.(w,\lambda)\in\{(1,-e_{n-1}^{*})\}\cup\{(w_{p},-e_{n}^{*})\mid 1\leq p<n\}.
Proof.

First note that by Lemma 3.3.2, we know that λ{en,en1}\lambda\in\{-e_{n}^{*},-e_{n-1}^{*}\}. Moreover, we have from the definition (3.2.1) that

WRn,Qnn0={wSnw1(i)<w1(i+1)for 1i<n1}={wp1pn}.W^{0}_{R_{n},Q^{n}_{n}}=\{w\in S_{n}\mid w^{-1}(i)<w^{-1}(i+1)\;\text{for}\;1\leq i<n-1\}=\{w_{p}\mid 1\leq p\leq n\}.

Since QnnGLnQ^{n}_{n}\subseteq GL_{n} is the standard Borel subgroup, the homomorphism

jw:𝕏(TQnn)=𝕏(TRnQnn)=𝕏(TRnwQnnw1)𝕏(TQnn)j_{w}\colon\mathbb{X}_{*}(T_{Q^{n}_{n}})=\mathbb{X}_{*}(T_{R_{n}\cap Q^{n}_{n}})=\mathbb{X}_{*}(T_{R_{n}\cap wQ^{n}_{n}w^{-1}})\longrightarrow\mathbb{X}_{*}(T_{Q^{n}_{n}})

defined in §3.2 is just the isomorphism given by w1w^{-1}. So by nonemptiness of Cnw,λC^{w,\lambda}_{n} there exists a semistable LnL_{n}-bundle ξLnEs\xi_{L_{n}}\to E_{s} on a geometric fibre of ESE\to S of degree e1-e_{1}^{*}, where LnGLn1×𝔾mL_{n}\cong GL_{n-1}\times\mathbb{G}_{m} is the standard Levi factor of RnR_{n} and a section σL:EsξLn/(LnQnn)\sigma_{L}\colon E_{s}\to\xi_{L_{n}}/(L_{n}\cap Q^{n}_{n}) of degree wλw\lambda. In particular, since en𝕏(Ln)e_{n}\in\mathbb{X}^{*}(L_{n}), en,wλ=en,e1=0\langle e_{n},w\lambda\rangle=\langle e_{n},-e_{1}^{*}\rangle=0 and wλw\lambda is the degree of a section

EsσLξLn/(LnQnn)⸦-→ξLn×LnGLn/Qnn.E_{s}\xrightarrow{\sigma_{L}}\xi_{L_{n}}/(L_{n}\cap Q^{n}_{n})\lhook\joinrel\relbar\joinrel\rightarrow\xi_{L_{n}}\times^{L_{n}}GL_{n}/Q^{n}_{n}.

If λ=en\lambda=-e_{n}^{*} and w=wpw=w_{p}, then

wλ={en1,ifp<n,en,ifp=n,w\lambda=\begin{cases}-e_{n-1}^{*},&\text{if}\;\;p<n,\\ -e_{n}^{*},&\text{if}\;\;p=n,\end{cases}

so from the above discussion we must have p{1,,n1}p\in\{1,\ldots,n-1\}. If λ=en1\lambda=-e_{n-1}^{*}, on the other hand, then

wλ={en2,ifp<n1,en,ifp=n1,en1,ifp=n,w\lambda=\begin{cases}-e_{n-2}^{*},&\text{if}\;\;p<n-1,\\ -e_{n}^{*},&\text{if}\;\;p=n-1,\\ -e_{n-1}^{*},&\text{if}\;\;p=n,\end{cases}

so the above discussion and Lemma 3.3.2 imply that p=np=n. Combining these two cases gives that (w,λ)(w,\lambda) is in the desired set. ∎

Lemma 3.3.4.

For all λ𝕏(TQnn)\lambda\in\mathbb{X}_{*}(T_{Q^{n}_{n}}) with λen\lambda\leq-e_{n}^{*}, we have

wWRn,Qnn0Cnw,λ=BunQnnλ×BunGLn1BunRnss,e1.\bigcup_{w\in W^{0}_{R_{n},Q^{n}_{n}}}C^{w,\lambda}_{n}=\mathrm{Bun}_{Q^{n}_{n}}^{\lambda}\times_{\mathrm{Bun}_{GL_{n}}^{-1}}\mathrm{Bun}_{R_{n}}^{ss,-e_{1}^{*}}.
Proof.

Assume for a contradiction that this fails for some λen\lambda\leq-e_{n}^{*}. Then by Proposition 3.2.1 there exist wWRn,Qnn0{1}w\in W^{0}_{R_{n},Q^{n}_{n}}\setminus\{1\} and λ<λ\lambda^{\prime}<\lambda such that Cnw,λC^{w,\lambda^{\prime}}_{n}\neq\emptyset. So Lemmas 3.3.2 and 3.3.3 imply that λ=en\lambda^{\prime}=-e_{n}^{*} and λ{en,en1}\lambda\in\{-e_{n}^{*},-e_{n-1}^{*}\}. But this contradicts λ<λ\lambda^{\prime}<\lambda so we are done. ∎

Lemma 3.3.5.

Let 1k<n1\leq k<n. Then

WRn,Qkn0={wp1p<k}{wn}W^{0}_{R_{n},Q^{n}_{k}}=\{w_{p}\mid 1\leq p<k\}\cup\{w_{n}\}

and

BunQknen×BunGLnBunRnss,e1=wWRn,Qkn0Ckw,en.\mathrm{Bun}_{Q^{n}_{k}}^{-e_{n}^{*}}\times_{\mathrm{Bun}_{GL_{n}}}\mathrm{Bun}_{R_{n}}^{ss,-e_{1}^{*}}=\bigcup_{w\in W^{0}_{R_{n},Q^{n}_{k}}}C^{w,-e_{n}^{*}}_{k}. (3.3.2)
Proof.

From the definition,

WRn,Qkn0={wWRn,Qnn0w(i)<w(i+1)forkin1}={wp1p<k}{wn}W^{0}_{R_{n},Q^{n}_{k}}=\{w\in W^{0}_{R_{n},Q^{n}_{n}}\mid w(i)<w(i+1)\;\text{for}\;k\leq i\leq n-1\}=\{w_{p}\mid 1\leq p<k\}\cup\{w_{n}\}

as claimed. Next, note that by [davis19a, Proposition 3.6.4] the natural morphism

KMQnn,GLnenKMQkn,GLnen\mathrm{KM}_{Q^{n}_{n},GL_{n}}^{-e_{n}^{*}}\longrightarrow\mathrm{KM}_{Q^{n}_{k},GL_{n}}^{-e_{n}^{*}}

is surjective. So any geometric point of BunQknen×BunGLnBunRnss,e1\mathrm{Bun}_{Q^{n}_{k}}^{-e_{n}^{*}}\times_{\mathrm{Bun}_{GL_{n}}}\mathrm{Bun}_{R_{n}}^{ss,-e_{1}^{*}} lifts to a point of BunQnnλ×BunGLnBunRnss,e1\mathrm{Bun}_{Q^{n}_{n}}^{\lambda}\times_{\mathrm{Bun}_{GL_{n}}}\mathrm{Bun}_{R_{n}}^{ss,-e_{1}^{*}} for some λen\lambda\leq-e_{n}^{*}, and hence λ{en,en1}\lambda\in\{-e_{n}^{*},-e_{n-1}^{*}\} by Lemma 3.3.2. So by Lemma 3.3.4, the morphism

wWRn,Qnn0λ{en,en1}Cnw,λwWRn,Qkn0Ckw,enBunQknen×BunGLnBunRnss,e1\coprod_{\begin{subarray}{c}w\in W^{0}_{R_{n},Q^{n}_{n}}\\ \lambda\in\{-e_{n}^{*},-e_{n-1}^{*}\}\end{subarray}}C^{w,\lambda}_{n}\longrightarrow\coprod_{w\in W^{0}_{R_{n},Q^{n}_{k}}}C^{w,-e_{n}^{*}}_{k}\longrightarrow\mathrm{Bun}_{Q^{n}_{k}}^{-e_{n}^{*}}\times_{\mathrm{Bun}_{GL_{n}}}\mathrm{Bun}_{R_{n}}^{ss,-e_{1}^{*}}

is surjective, which proves (3.3.2). ∎

Proof of Proposition 3.3.1.

Suppose first that k<nk<n. Since any QknQ^{n}_{k}-bundle of degree en\leq-e_{n}^{*} can be reduced to a QnnQ^{n}_{n}-bundle of degree en\leq-e_{n}^{*} by [davis19a, Proposition 3.6.4], Lemma 3.3.2 implies that

KMQkn,GLnen×BunGLnBunRnss,e1=BunQknen×BunGLnBunRnss,e1,\mathrm{KM}_{Q_{k}^{n},GL_{n}}^{-e_{n}^{*}}\times_{\mathrm{Bun}_{GL_{n}}}\mathrm{Bun}_{R_{n}}^{ss,-e_{1}^{*}}=\mathrm{Bun}_{Q^{n}_{k}}^{-e_{n}^{*}}\times_{\mathrm{Bun}_{GL_{n}}}\mathrm{Bun}_{R_{n}}^{ss,-e_{1}^{*}},

since en-e_{n}^{*} and en1-e_{n-1}^{*} have the same image in 𝕏(TQkn)\mathbb{X}_{*}(T_{Q^{n}_{k}}). So we have the desired decomposition of XknX_{k}^{n} into locally closed substacks by Lemma 3.3.5 (note that Ck,pGLnC^{GL_{n}}_{k,p} is the preimage of Ckwp,λC^{w_{p},\lambda}_{k} in XknX^{n}_{k} in this case).

On the other hand, if k=nk=n, then Lemma 3.3.2 implies that XnnX^{n}_{n} decomposes as a disjoint union

Xnn=(BunQnnen×BunGLnBunRnss,e1)(BunQnnen1×BunGLnBunRnss,e1×SE)X_{n}^{n}=(\mathrm{Bun}_{Q^{n}_{n}}^{-e_{n}^{*}}\times_{\mathrm{Bun}_{GL_{n}}}\mathrm{Bun}_{R_{n}}^{ss,-e_{1}^{*}})\cup(\mathrm{Bun}_{Q^{n}_{n}}^{-e_{n-1}^{*}}\times_{\mathrm{Bun}_{GL_{n}}}\mathrm{Bun}_{R_{n}}^{ss,-e_{1}^{*}}\times_{S}E)

of locally closed substacks, where the first factor is the locus of stable sections with irreducible domain and the second factor is the locus of stable sections with a single rational component of degree βn1=en1en\beta_{n-1}^{\vee}=e_{n-1}^{*}-e_{n}^{*}. By Lemmas 3.3.3 and 3.3.4, this decomposes further as the desired decomposition

Xnn=1p<nCn,pGLnCn,nGLnX_{n}^{n}=\bigcup_{1\leq p<n}C_{n,p}^{GL_{n}}\cup C_{n,n}^{GL_{n}}

so we are done. ∎

From the proof of Proposition 3.3.1, we have that

Cn,nGLnCn1,en1×SE=BunQnnen1×BunGLnBunRnss,e1×SEC^{GL_{n}}_{n,n}\cong C^{1,-e_{n-1}^{*}}_{n}\times_{S}E=\mathrm{Bun}_{Q^{n}_{n}}^{-e_{n-1}^{*}}\times_{\mathrm{Bun}_{GL_{n}}}\mathrm{Bun}_{R_{n}}^{ss,-e_{1}^{*}}\times_{S}E

is the locus of stable maps with a single rational component of degree βn1\beta_{n-1}^{\vee}. The natural projection to EE keeps track of the point of attachment of the rational component, and the projection to the other factors keeps track of the restriction to the elliptic component. Note that the projection to EE agrees with composition of Cn,nGLnC1,nGLnC_{n,n}^{GL_{n}}\to C_{1,n}^{GL_{n}} with the morphism

C1,nGLnYQnnen×PicS1(E)YRne1\displaystyle C_{1,n}^{GL_{n}}\longrightarrow Y_{Q^{n}_{n}}^{-e_{n}^{*}}\times_{\mathrm{Pic}^{-1}_{S}(E)}Y_{R_{n}}^{-e_{1}^{*}} PicS1(E)=E\displaystyle\longrightarrow\mathrm{Pic}^{1}_{S}(E)=E (3.3.3)
(y,y)\displaystyle(y,y^{\prime}) en(y)en(y).\displaystyle\longmapsto e_{n}(y^{\prime})-e_{n}(y).

For 1p<n1\leq p<n, we let

MpGLnC1,nGLnM_{p}^{GL_{n}}\subseteq C_{1,n}^{GL_{n}}

be the closed substack given by the fibre product

MpGLn{M_{p}^{GL_{n}}}C1,nGLn{C_{1,n}^{GL_{n}}}YQnnen{Y_{Q^{n}_{n}}^{-e_{n}^{*}}}YQnnen×SE,{Y_{Q^{n}_{n}}^{-e_{n}^{*}}\times_{S}E,}θpGLn\scriptstyle{\theta_{p}^{GL_{n}}}

where the morphism C1,nGLnEC_{1,n}^{GL_{n}}\to E is (3.3.3), and the morphism YQnnenYQnnen×SPicS1(E)Y_{Q^{n}_{n}}^{-e_{n}^{*}}\to Y_{Q^{n}_{n}}^{-e_{n}^{*}}\times_{S}\mathrm{Pic}^{1}_{S}(E) is given by

θpGLn:YQnnen\displaystyle\theta^{GL_{n}}_{p}\colon Y_{Q^{n}_{n}}^{-e_{n}^{*}} YQnnen×SPicS1(E)=YQnnen×SE\displaystyle\longrightarrow Y_{Q^{n}_{n}}^{-e_{n}^{*}}\times_{S}\mathrm{Pic}^{1}_{S}(E)=Y_{Q^{n}_{n}}^{-e_{n}^{*}}\times_{S}E
y\displaystyle y (y,ep(y)en(y)).\displaystyle\longmapsto(y,e_{p}(y)-e_{n}(y)).
Proposition 3.3.6.

For all 1k<n1\leq k<n, the morphism Xk+1nXknX^{n}_{k+1}\to X^{n}_{k} restricts to isomorphisms

Ck+1,nGLnCk,nGLnandCk+1,pGLnCk,pGLnC^{GL_{n}}_{k+1,n}\overset{\sim}{\longrightarrow}C^{GL_{n}}_{k,n}\quad\text{and}\quad C^{GL_{n}}_{k+1,p}\overset{\sim}{\longrightarrow}C^{GL_{n}}_{k,p}

for 1p<k1\leq p<k, and a morphism

Ck+1,kGLnMkGLnCk,nGLnC1,nGLnC^{GL_{n}}_{k+1,k}\longrightarrow M_{k}^{GL_{n}}\subseteq C^{GL_{n}}_{k,n}\cong C^{GL_{n}}_{1,n}

that exhibits Ck+1,kGLnC_{k+1,k}^{GL_{n}} as an 𝔸1\mathbb{A}^{1}-bundle over MkGLnM_{k}^{GL_{n}}.

Proof.

If k<n1k<n-1, then the morphism Ck+1,nGLnCk,nGLnC_{k+1,n}^{GL_{n}}\to C_{k,n}^{GL_{n}} can be identified with

YQnnen×YQk+1nenBunRnQk+1nen1×BunRne1BunRnss,e1YQnnen×YQknenBunRnQknen1×BunRne1BunRnss,e1.Y_{Q^{n}_{n}}^{-e_{n}^{*}}\times_{Y_{Q^{n}_{k+1}}^{-e_{n}^{*}}}\mathrm{Bun}_{R_{n}\cap Q^{n}_{k+1}}^{-e_{n-1}^{*}}\times_{\mathrm{Bun}_{R_{n}}^{-e_{1}^{*}}}\mathrm{Bun}_{R_{n}}^{ss,-e_{1}^{*}}\longrightarrow Y_{Q^{n}_{n}}^{-e_{n}^{*}}\times_{Y_{Q^{n}_{k}}^{-e_{n}^{*}}}\mathrm{Bun}_{R_{n}\cap Q^{n}_{k}}^{-e_{n-1}^{*}}\times_{\mathrm{Bun}_{R_{n}}^{-e_{1}^{*}}}\mathrm{Bun}_{R_{n}}^{ss,-e_{1}^{*}}.

This is a pullback of

BunQk+1n1en1×BunGLn1BunGLn1ss,1YQk+1n1en1×YQkn1en1BunQkn1en1×BunGLn1BunGLn1ss,1\mathrm{Bun}_{Q^{n-1}_{k+1}}^{-e_{n-1}^{*}}\times_{\mathrm{Bun}_{GL_{n-1}}}\mathrm{Bun}_{GL_{n-1}}^{ss,-1}\longrightarrow Y^{-e_{n-1}^{*}}_{Q^{n-1}_{k+1}}\times_{Y^{-e_{n-1}^{*}}_{Q^{n-1}_{k}}}\mathrm{Bun}_{Q^{n-1}_{k}}^{-e_{n-1}^{*}}\times_{\mathrm{Bun}_{GL_{n-1}}}\mathrm{Bun}_{GL_{n-1}}^{ss,-1}

under the morphism RnGLn1R_{n}\to GL_{n-1} forgetting the last row and column, hence an isomorphism by [davis19, Lemma 4.3.7].

If k=n1k=n-1, then we can identify Ck+1,nGLnCk,nGLnC^{GL_{n}}_{k+1,n}\to C^{GL_{n}}_{k,n} with the morphism

BunQnnRnen1×BunRnBunRnss,e1×SEYQnnen×YQn1nenBunRnQn1nen1×BunRnBunRnss,e1.\mathrm{Bun}_{Q^{n}_{n}\cap R_{n}}^{-e_{n-1}^{*}}\times_{\mathrm{Bun}_{R_{n}}}\mathrm{Bun}_{R_{n}}^{ss,-e_{1}^{*}}\times_{S}E\longrightarrow Y_{Q^{n}_{n}}^{-e_{n}^{*}}\times_{Y_{Q^{n}_{n-1}}^{-e_{n}^{*}}}\mathrm{Bun}_{R_{n}\cap Q^{n}_{n-1}}^{-e_{n-1}^{*}}\times_{\mathrm{Bun}_{R_{n}}}\mathrm{Bun}_{R_{n}}^{ss,-e_{1}^{*}}.

Since RnQnn=RnQn1n=QnnR_{n}\cap Q^{n}_{n}=R_{n}\cap Q^{n}_{n-1}=Q^{n}_{n}, this is naturally a pullback of the isomorphism

YQnnen1×SE\displaystyle Y_{Q^{n}_{n}}^{-e_{n-1}^{*}}\times_{S}E YQnnen×YQn1nenYQnnen1\displaystyle\overset{\sim}{\longrightarrow}Y_{Q^{n}_{n}}^{-e_{n}^{*}}\times_{Y_{Q^{n}_{n-1}}^{-e_{n}^{*}}}Y_{Q_{n}^{n}}^{-e_{n-1}^{*}}
(y,x)\displaystyle(y,x) (y+βn1(x),y),\displaystyle\longmapsto(y+\beta_{n-1}^{\vee}(x),y),

hence an isomorphism itself.

If knk\leq n and 1p<k1\leq p<k, then LnwpQknwp1=LnQk1nL_{n}\cap w_{p}Q^{n}_{k}w_{p}^{-1}=L_{n}\cap Q^{n}_{k-1}, where LnRnL_{n}\subseteq R_{n} is the standard Levi factor. One easily checks that, in the notation of §3.2, the morphism

(iwp,jwp):𝕏(TRnwpQknwp1)𝕏(TRn)𝕏(TQkn)(i_{w_{p}},j_{w_{p}})\colon\mathbb{X}_{*}(T_{R_{n}\cap w_{p}Q^{n}_{k}w_{p}^{-1}})\longrightarrow\mathbb{X}_{*}(T_{R_{n}})\oplus\mathbb{X}_{*}(T_{Q^{n}_{k}})

is injective and sends en1-e_{n-1}^{*} to (e1,en)(-e_{1}^{*},-e_{n}^{*}). So

Ckwp,en=BunRnwpQknwp1en1×BunRnBunRnss,e1.C_{k}^{w_{p},-e_{n}^{*}}=\mathrm{Bun}_{R_{n}\cap w_{p}Q^{n}_{k}w_{p}^{-1}}^{-e_{n-1}^{*}}\times_{\mathrm{Bun}_{R_{n}}}\mathrm{Bun}_{R_{n}}^{ss,-e_{1}^{*}}.

By general nonsense, the right hand side is the relative space of sections of

ηk,p×(LnwpQknwp1)Ru(Rn)Ru(Rn)Ru(Rn)wpQknwp1BunLnwpQknwp1en1×BunLnBunRnss,e1×SE\eta_{k,p}\times^{(L_{n}\cap w_{p}Q^{n}_{k}w_{p}^{-1})R_{u}(R_{n})}\frac{R_{u}(R_{n})}{R_{u}(R_{n})\cap w_{p}Q^{n}_{k}w_{p}^{-1}}\longrightarrow\mathrm{Bun}_{L_{n}\cap w_{p}Q^{n}_{k}w_{p}^{-1}}^{-e_{n-1}^{*}}\times_{\mathrm{Bun}_{L_{n}}}\mathrm{Bun}_{R_{n}}^{ss,-e_{1}^{*}}\times_{S}E

over

BunLnwpQknwp1en1×BunLnBunRnss,e1Bun(LnwpQknwp1)Ru(Rn),\mathrm{Bun}_{L_{n}\cap w_{p}Q^{n}_{k}w_{p}^{-1}}^{-e_{n-1}^{*}}\times_{\mathrm{Bun}_{L_{n}}}\mathrm{Bun}_{R_{n}}^{ss,-e_{1}^{*}}\subseteq\mathrm{Bun}_{(L_{n}\cap w_{p}Q^{n}_{k}w_{p}^{-1})R_{u}(R_{n})},

where ηk,p\eta_{k,p} is the universal (LnwpQknwp1)Ru(Rn)(L_{n}\cap w_{p}Q^{n}_{k}w_{p}^{-1})R_{u}(R_{n})-bundle. By Lemma 3.3.7 below, we can therefore identify Ck,pGLn=YQnnen×YQknenCkwp,enC_{k,p}^{GL_{n}}=Y_{Q^{n}_{n}}^{-e_{n}^{*}}\times_{Y_{Q^{n}_{k}}^{-e_{n}^{*}}}C_{k}^{w_{p},-e_{n}^{*}} with the relative space of sections of

η¯k,p×(LnwpQknwp1)Ru(Rn)Ru(Rn)Ru(Rn)wpQknwp1MpGLn×SE\bar{\eta}_{k,p}\times^{(L_{n}\cap w_{p}Q^{n}_{k}w_{p}^{-1})R_{u}(R_{n})}\frac{R_{u}(R_{n})}{R_{u}(R_{n})\cap w_{p}Q^{n}_{k}w_{p}^{-1}}\longrightarrow M_{p}^{GL_{n}}\times_{S}E (3.3.4)

over MpC1,nGLnM_{p}\subseteq C_{1,n}^{GL_{n}}, where η¯k,p\bar{\eta}_{k,p} is a pullback of ηk,p\eta_{k,p}. Note that by Lemma 3.3.8 below, there is an isomorphism

Ru(Rn)Ru(Rn)wpQknwp1Uk,pen,\frac{R_{u}(R_{n})}{R_{u}(R_{n})\cap w_{p}Q^{n}_{k}w_{p}^{-1}}\cong U_{k,p}^{\vee}\otimes\mathbb{Z}_{e_{n}},

of LnwpQknwp1L_{n}\cap w_{p}Q^{n}_{k}w_{p}^{-1}-varieties, where Uk,pU_{k,p} is the representation described immediately before Lemma 3.3.8. So after pulling back along the smooth surjection BunLnss,e1BunRnss,e1\mathrm{Bun}_{L_{n}}^{ss,-e_{1}^{*}}\to\mathrm{Bun}_{R_{n}}^{ss,-e_{1}^{*}}, (3.3.4) becomes a family of stable vector bundles on EE of degree 11.

If kn1k\leq n-1, then by the above discussion, the morphism Ck+1,pGLnCk,pGLnC^{GL_{n}}_{k+1,p}\to C^{GL_{n}}_{k,p} becomes the pushforward of a surjective morphism between families of stable vector bundles of degree 11 over MpGLnM_{p}^{GL_{n}} after pulling back along BunLnss,e1BunRnss,e1\mathrm{Bun}_{L_{n}}^{ss,-e_{1}^{*}}\to\mathrm{Bun}_{R_{n}}^{ss,-e_{1}^{*}}, and is therefore an isomorphism as claimed. On the other hand, the morphism Ck+1,kGLnCk,nGLnC_{k+1,k}^{GL_{n}}\to C_{k,n}^{GL_{n}} becomes the relative space of sections over MkGLnC1,nGLnCk,nGLnM_{k}^{GL_{n}}\subseteq C_{1,n}^{GL_{n}}\cong C_{k,n}^{GL_{n}} of a family of stable vector bundles of degree 11, and is therefore an 𝔸1\mathbb{A}^{1}-bundle over MkGLnM_{k}^{GL_{n}}. ∎

Lemma 3.3.7.

If p<knp<k\leq n, then the morphism

YQnnen×YQknen(BunLnwpQknwp1en1×BunLne1BunRnss,e1)YQnnen×PicS1(E)BunRnss,e1=C1,nGLn=X1nY_{Q^{n}_{n}}^{-e_{n}^{*}}\times_{Y_{Q^{n}_{k}}^{-e_{n}^{*}}}(\mathrm{Bun}_{L_{n}\cap w_{p}Q^{n}_{k}w_{p}^{-1}}^{-e_{n-1}^{*}}\times_{\mathrm{Bun}_{L_{n}}^{-e_{1}^{*}}}\mathrm{Bun}_{R_{n}}^{ss,-e_{1}^{*}})\longrightarrow Y_{Q^{n}_{n}}^{-e_{n}^{*}}\times_{\mathrm{Pic}^{-1}_{S}(E)}\mathrm{Bun}_{R_{n}}^{ss,-e_{1}^{*}}=C^{GL_{n}}_{1,n}=X^{n}_{1} (3.3.5)

induced by the inclusion LnwpQknwp1LnL_{n}\cap w_{p}Q^{n}_{k}w_{p}^{-1}\subseteq L_{n} factors through an isomorphism onto MpGLnM_{p}^{GL_{n}}. Here the morphisms to PicS1(E)\mathrm{Pic}^{-1}_{S}(E) in the fibre product in the right hand side of (3.3.5) are both given by the determinant.

Proof.

First note that the morphism

BunLnwpQknwp1en1×BunLne1BunRnss,e1YLnwpQknwp1en1×YRne1BunRnss,e1\mathrm{Bun}_{L_{n}\cap w_{p}Q^{n}_{k}w_{p}^{-1}}^{-e_{n-1}^{*}}\times_{\mathrm{Bun}_{L_{n}}^{-e_{1}^{*}}}\mathrm{Bun}_{R_{n}}^{ss,-e_{1}^{*}}\longrightarrow Y_{L_{n}\cap w_{p}Q^{n}_{k}w_{p}^{-1}}^{-e_{n-1}^{*}}\times_{Y_{R_{n}}^{-e_{1}^{*}}}\mathrm{Bun}_{R_{n}}^{ss,-e_{1}^{*}}

is a pullback of

BunQk1n1en1×BunGLn1BunGLn1ss,1YQk1n1en1×PicS1(E)BunGLn1ss,1\mathrm{Bun}_{Q^{n-1}_{k-1}}^{-e_{n-1}^{*}}\times_{\mathrm{Bun}_{GL_{n-1}}}\mathrm{Bun}_{GL_{n-1}}^{ss,-1}\longrightarrow Y_{Q^{n-1}_{k-1}}^{-e_{n-1}^{*}}\times_{\mathrm{Pic}^{-1}_{S}(E)}\mathrm{Bun}_{GL_{n-1}}^{ss,-1}

and hence an isomorphism by [davis19, Lemma 4.3.7]. Composing with the isomorphism

jwp:YLnwpQknwp1en1YQknenj_{w_{p}}\colon Y_{L_{n}\cap w_{p}Q^{n}_{k}w_{p}^{-1}}^{-e_{n-1}^{*}}\overset{\sim}{\longrightarrow}Y_{Q^{n}_{k}}^{-e_{n}^{*}}

allows us to identify (3.3.5) with the closed immerison

YQnnen×YRne1BunRnss,e1YQnnen×PicS1(E)BunRnss,e1,Y_{Q^{n}_{n}}^{-e_{n}^{*}}\times_{Y_{R_{n}}^{-e_{1}^{*}}}\mathrm{Bun}_{R_{n}}^{ss,-e_{1}^{*}}\longrightarrow Y_{Q^{n}_{n}}^{-e_{n}^{*}}\times_{\mathrm{Pic}^{-1}_{S}(E)}\mathrm{Bun}_{R_{n}}^{ss,-e_{1}^{*}},

where the morphism YQnnenYRne1Y_{Q^{n}_{n}}^{-e_{n}^{*}}\to Y_{R_{n}}^{-e_{1}^{*}} is the composition

YQnnenYQknenjwp1YLnwpQknwp1en1iwpYRne1.Y_{Q^{n}_{n}}^{-e_{n}^{*}}\longrightarrow Y_{Q^{n}_{k}}^{-e_{n}^{*}}\xrightarrow{j_{w_{p}}^{-1}}Y_{L_{n}\cap w_{p}Q^{n}_{k}w_{p}^{-1}}^{-e_{n-1}^{*}}\xrightarrow{i_{w_{p}}}Y_{R_{n}}^{-e_{1}^{*}}.

Chasing through the various definitions now shows that the source of this morphism is precisely MpGLnM_{p}^{GL_{n}}, so we are done. ∎

In the following lemma, we write Uk,pU_{k,p} for the LnwpQknwp1L_{n}\cap w_{p}Q^{n}_{k}w_{p}^{-1}-representation induced by the homomorphism

LnwpQknwp1=Qk1n1×𝔾mQk1n1GLnpL_{n}\cap w_{p}Q^{n}_{k}w_{p}^{-1}=Q^{n-1}_{k-1}\times\mathbb{G}_{m}\longrightarrow Q^{n-1}_{k-1}\longrightarrow GL_{n-p}

given by deleting the last row and column and the first p1p-1 rows and columns.

Lemma 3.3.8.

If p<kp<k, then there is an LnwpQknwp1L_{n}\cap w_{p}Q^{n}_{k}w_{p}^{-1}-equivariant isomorphism

Ru(Rn)/(Ru(Rn)wpQknwp1)Uk,pen.R_{u}(R_{n})/(R_{u}(R_{n})\cap w_{p}Q^{n}_{k}w_{p}^{-1})\overset{\sim}{\longrightarrow}U_{k,p}^{\vee}\otimes\mathbb{Z}_{e_{n}}. (3.3.6)
Proof.

If β\beta is a root of Ru(Rn)R_{u}(R_{n}), then the root subgroup Uβ𝔾aRu(Rn)U_{\beta}\cong\mathbb{G}_{a}\subseteq R_{u}(R_{n}) maps injectively into Ru(Rn)/(Ru(Rn)wpQknwp1)R_{u}(R_{n})/(R_{u}(R_{n})\cap w_{p}Q^{n}_{k}w_{p}^{-1}) if and only if wp1βw_{p}^{-1}\beta is not a root of QknQ^{n}_{k}. In particular, this implies that β\beta is a negative root and wp1βw_{p}^{-1}\beta is a positive root, and hence that

βΣ={βn1,βn1βn2,,βn1βn2βp},\beta\in\Sigma=\{-\beta_{n-1},-\beta_{n-1}-\beta_{n-2},\ldots,-\beta_{n-1}-\beta_{n-2}-\cdots-\beta_{p}\},

and

wp1β{βn1+βn2++βp,βn2++βp,,βp}.w_{p}^{-1}\beta\in\{\beta_{n-1}+\beta_{n-2}+\cdots+\beta_{p},\beta_{n-2}+\cdots+\beta_{p},\ldots,\beta_{p}\}.

Note that if βΣ\beta\in\Sigma, then UβRu(P)U_{\beta}\subseteq R_{u}(P), and wp1βw_{p}^{-1}\beta is not a root of QknQ^{n}_{k}, so Σ\Sigma is precisely the set of roots appearing in Ru(Rn)/(Ru(Rn)wpQknwp1)R_{u}(R_{n})/(R_{u}(R_{n})\cap w_{p}Q^{n}_{k}w_{p}^{-1}).

It is clear from the above that Ru(Rn)/(Ru(Rn)wpQknwp1)R_{u}(R_{n})/(R_{u}(R_{n})\cap w_{p}Q^{n}_{k}w_{p}^{-1}) is isomorphic to an LnwpQknwp1L_{n}\cap w_{p}Q^{n}_{k}w_{p}^{-1}-representation. The isomorphism (3.3.6) follows by inspection of the weights of this representation. ∎

3.4. The divisor Dαj(Z)D_{\alpha_{j}^{\vee}}(Z)

The purpose of this subsection is to prove Proposition 3.4.1 below, which refines Theorem 1.0.3, (2). For the statement, recall Notation 2.3.2. For 1kn0+11\leq k\leq n_{0}+1, we write θk\theta_{k} for the section

θk:Y\displaystyle\theta_{k}\colon Y Y×SPicS0(E)\displaystyle\longrightarrow Y\times_{S}\mathrm{Pic}^{0}_{S}(E)
y\displaystyle y {(y,ϖj(y)ϖi(y)ϖc0,1(y)),ifk=1,(y,ϖj(y)ϖi(y)ϖc0,k(y)+ϖc0,k1(y)),if  1<kn0,(y,0),ifk=n0+1.\displaystyle\longmapsto\begin{cases}(y,\varpi_{j}(y)-\varpi_{i}(y)-\varpi_{c_{0},1}(y)),&\text{if}\;\;k=1,\\ (y,\varpi_{j}(y)-\varpi_{i}(y)-\varpi_{c_{0},k}(y)+\varpi_{c_{0},k-1}(y)),&\text{if}\;\;1<k\leq n_{0},\\ (y,0),&\text{if}\;\;k=n_{0}+1.\end{cases}
Proposition 3.4.1.

Assume we are in the setup of Proposition 3.1.1. Then there is a sequence of n0+1n_{0}+1 morphisms

Dαj(Z)=Dn0+2Dn0+1D1D_{\alpha_{j}^{\vee}}(Z)=D_{n_{0}+2}\longrightarrow D_{n_{0}+1}\longrightarrow\cdots\longrightarrow D_{1}

over Y×SZY\times_{S}Z such that D1D_{1} is a line bundle over Y×SPicS0(E)Y\times_{S}\mathrm{Pic}^{0}_{S}(E) and Dk+1DkD_{k+1}\to D_{k} is the blowup along the section θk:YY×SPicS0(E)Dk\theta_{k}\colon Y\to Y\times_{S}\mathrm{Pic}^{0}_{S}(E)\subseteq D_{k} of the proper transform of the zero section of D1D_{1}.

Proof.

The spaces DkD_{k} are defined as follows. For 1kn01\leq k\leq n_{0}, let PkGP_{k}\subseteq G be the standard parabolic with type t(Pk)=Δ{αc0,k,,αc0,n0}=Δ{αc0,k,,αc0,n01,αi}t(P_{k})=\Delta\setminus\{\alpha_{c_{0},k},\ldots,\alpha_{c_{0},n_{0}}\}=\Delta\setminus\{\alpha_{c_{0},k},\ldots,\alpha_{c_{0},n_{0}-1},\alpha_{i}\}, and let Pn0+1=BP_{n_{0}+1}=B. Then for 1kn0+11\leq k\leq n_{0}+1, we define

Dk\displaystyle D_{k} =YBαj×YPkαjKMPk,G,rigαj×BunG,rigZ×SE\displaystyle=Y_{B}^{-\alpha_{j}^{\vee}}\times_{Y_{P_{k}}^{-\alpha_{j}^{\vee}}}\mathrm{KM}_{P_{k},G,rig}^{-\alpha_{j}^{\vee}}\times_{\mathrm{Bun}_{G,rig}}Z\times_{S}E
Y×YPk(KMPk,G,rigαj×BunG,rigZ×SE),\displaystyle\cong Y\times_{Y_{P_{k}}}(\mathrm{KM}_{P_{k},G,rig}^{-\alpha_{j}^{\vee}}\times_{\mathrm{Bun}_{G,rig}}Z\times_{S}E),

where the morphism to YPkY_{P_{k}} in the last fibre product is given by the composition

KMPk,G/S,rigαj×BunG,rigZ×SEBlPkYPkαj×SE\displaystyle\mathrm{KM}_{P_{k},G/S,rig}^{-\alpha_{j}^{\vee}}\times_{\mathrm{Bun}_{G,rig}}Z\times_{S}E\xrightarrow{\mathrm{Bl}_{P_{k}}}Y_{P_{k}}^{-\alpha_{j}^{\vee}}\times_{S}E YPk\displaystyle\longrightarrow Y_{P_{k}}
(y,x)\displaystyle(y,x) y+αj(x).\displaystyle\longmapsto y+\alpha_{j}^{\vee}(x).

For kn0k\leq n_{0}, the morphism Dk+1DkD_{k+1}\to D_{k} is the obvious one induced by the inclusion Pk+1PkP_{k+1}\subseteq P_{k}. To describe the morphism Dαj(Z)Dn0+1D_{\alpha_{j}^{\vee}}(Z)\to D_{n_{0}+1}, note that every stable map parametrised by a point in Dαj(Z)D_{\alpha_{j}^{\vee}}(Z) has a unique rational irreducible component of degree αj\alpha_{j}^{\vee}. Deleting this rational component and recording the point of EE over which it was attached defines the morphism

Dαj(Z)KMB,G,rigαj×BunG,rigZ×SE=Dn0+1.D_{\alpha_{j}^{\vee}}(Z)\longrightarrow\mathrm{KM}_{B,G,rig}^{-\alpha_{j}^{\vee}}\times_{\mathrm{Bun}_{G,rig}}Z\times_{S}E=D_{n_{0}+1}.

For kn0+1k\leq n_{0}+1, the spaces DkD_{k} can be decomposed into locally closed subsets as follows. First, by Proposition 3.4.2, for zZ0Zz\in Z_{0}\subseteq Z, every stable section of ξG,z/P1=ξP,z×PG/P1\xi_{G,z}/P_{1}=\xi_{P,z}\times^{P}G/P_{1} of degree αj-\alpha_{j}^{\vee} is in fact a genuine section of the subvariety ξP,z×PPP1/P1ξP,z/(PP1)\xi_{P,z}\times^{P}PP_{1}/P_{1}\cong\xi_{P,z}/(P\cap P_{1}). So there is an isomorphism

KMPk,G,rigαj×BunG,rigZ0KMPk,P1,rigαj×BunP1,rigBunPP1,rigαiαj×BunP,rigZ0.\mathrm{KM}_{P_{k},G,rig}^{-\alpha_{j}^{\vee}}\times_{\mathrm{Bun}_{G,rig}}Z_{0}\cong\mathrm{KM}_{P_{k},P_{1},rig}^{-\alpha_{j}^{\vee}}\times_{\mathrm{Bun}_{P_{1},rig}}\mathrm{Bun}_{P\cap P_{1},rig}^{-\alpha_{i}^{\vee}-\alpha_{j}^{\vee}}\times_{\mathrm{Bun}_{P,rig}}Z_{0}. (3.4.1)

Explicitly, the right hand side is tautologically identified with the space of pairs (z,σ)(z,\sigma) where zZ0z\in Z_{0} and σ\sigma is a stable section of ξP,z×PP1P1/Pk\xi_{P,z}\times^{P\cap P_{1}}P_{1}/P_{k} of appropriate degree such that the image in ξP,z/(PP1)\xi_{P,z}/(P\cap P_{1}) is a genuine section, while the left hand side is the space of pairs (z,σ)(z,\sigma) where zZ0z\in Z_{0} and σ\sigma is a stable section of ξP,z×PG/Pk\xi_{P,z}\times^{P}G/P_{k}. The isomorphism (3.4.1) is obtained in this description by applying the isomorphism P×PP1P1/PkPP1/PkP\times^{P\cap P_{1}}P_{1}/P_{k}\cong PP_{1}/P_{k} and the inclusion PP1/PkG/PkPP_{1}/P_{k}\hookrightarrow G/P_{k}. The homomorphism πP1:P1GLn0+1\pi_{P_{1}}\colon P_{1}\to GL_{n_{0}+1} of Proposition 3.4.5 therefore induces a morphism

Dk×ZZ0Xk,rign0+1,D_{k}\times_{Z}Z_{0}\longrightarrow X_{k,rig}^{n_{0}+1}, (3.4.2)

where Xk,rign0+1X_{k,rig}^{n_{0}+1} is the rigidification of the space Xkn0+1X_{k}^{n_{0}+1} of §3.3 with respect to the image of Z(G)Z(G) in Z(GLn0+1)Z(GL_{n_{0}+1}). We therefore get a decomposition

Dk=(Dk×Z(ZZ0))1p<kCk,pCk,n0+1,D_{k}=(D_{k}\times_{Z}(Z\setminus Z_{0}))\cup\bigcup_{1\leq p<k}C_{k,p}\cup C_{k,n_{0}+1}, (3.4.3)

where Ck,pC_{k,p} is the preimage of Ck,pGLn0+1Xkn0+1C_{k,p}^{GL_{n_{0}+1}}\subseteq X^{n_{0}+1}_{k} under (3.4.2). The behaviour of these locally closed subsets under the morphisms Dk+1DkD_{k+1}\to D_{k} is described by Proposition 3.4.6.

There is a morphism C1,n0+1Y×SPicS0(E)C_{1,n_{0}+1}\to Y\times_{S}\mathrm{Pic}^{0}_{S}(E) (3.4.6), which is defined so that the image of a stable section in Dαj(Z)D_{\alpha_{j}^{\vee}}(Z) over yYy\in Y with two rational components is sent to (y,xjxi)(y,x_{j}-x_{i}), where xjEx_{j}\in E (resp. xiEx_{i}\in E) is the point of attachment of the rational component of degree αj\alpha_{j}^{\vee} (resp. αi\alpha_{i}^{\vee}). This morphism is an isomorphism by Lemma 3.4.8.

Since KMPk,G,rigαj×BunG,rigZ\mathrm{KM}_{P_{k},G,rig}^{-\alpha_{j}^{\vee}}\times_{\mathrm{Bun}_{G,rig}}Z is smooth over YPkαjY_{P_{k}}^{-\alpha_{j}^{\vee}}, each space DkD_{k} is smooth over YY. Proposition 3.4.6 implies that they are all isomorphic to Dn0+1D_{n_{0}+1}, and hence to Dαj(Z)D_{\alpha_{j}^{\vee}}(Z), outside Z0Z_{0}. So the spaces DkD_{k} are all smooth surfaces over YY.

In particular, C1,n0+1=D1×ZZ0C_{1,n_{0}+1}=D_{1}\times_{Z}Z_{0} is a Cartier divisor on D1D_{1}. Moreover, choosing any cocharacter of the torus Z(L)rigZ(L)_{rig} whose negative is a Harder-Narasimhan vector for the parabolic P+P^{+} opposite to PP, we get compatible actions of 𝔾m\mathbb{G}_{m} on ZZ and D1D_{1} acting trivially on Z0Z_{0} and D1×ZZ0D_{1}\times_{Z}Z_{0}, such that 𝔾m\mathbb{G}_{m} acts on the fibres of the affine space bundle ZZ0Z\to Z_{0} with positive weights. Since the normal cone of D1×ZZ0D_{1}\times_{Z}Z_{0} in D1D_{1} is a line bundle and 𝔾m\mathbb{G}_{m} acts nontrivially on it, 𝔾m\mathbb{G}_{m} acts on it with a single nonzero weight. So [davis19, Lemma 4.3.11] shows that D1D_{1} is isomorphic to a line bundle over C1,n0+1=Y×SPicS0(E)C_{1,n_{0}+1}=Y\times_{S}\mathrm{Pic}^{0}_{S}(E) as claimed.

It remains to show that Dk+1DkD_{k+1}\to D_{k} is the blowup along the proper transform of θk\theta_{k} for 1kn0+11\leq k\leq n_{0}+1. If kn0k\leq n_{0}, this follows from Proposition 3.4.6 and Lemma 3.4.10. For k=n0+1k=n_{0}+1, note that Dn0+2=Dαj(Z)Dn0+1D_{n_{0}+2}=D_{\alpha_{j}^{\vee}}(Z)\to D_{n_{0}+1} is an isomorphism outside the proper transform of θn0+1\theta_{n_{0}+1} (the locus of curves with a degree αi\alpha_{i}^{\vee} rational component over the point of attachment of the degree αj\alpha_{j}^{\vee} rational curve), and the fibre over any point in this proper transform is an irreducible curve. The claim now follows by Lemma 3.4.10 again. ∎

The rest of the subsection is devoted to the various lemmas and propositions quoted in the proof of Proposition 3.4.1.

Proposition 3.4.2.

Let zZ0Zz\in Z_{0}\subseteq Z and let ξP,z\xi_{P,z} and ξG,z=ξP,z×PG\xi_{G,z}=\xi_{P,z}\times^{P}G be the corresponding PP and GG-bundles. Then any stable section of ξG,z/P1=ξP,z×GG/P1\xi_{G,z}/P_{1}=\xi_{P,z}\times^{G}G/P_{1} of degree αj-\alpha_{j}^{\vee} is a genuine section, and factors through ξP,z×PPP1/P1\xi_{P,z}\times^{P}PP_{1}/P_{1}.

Proof.

The proposition is equivalent to the claim that

CP11,αj(Z0)⸦-→BunP1,rigαj×BunG,rigZ0⸦-→KMP1,G,rigαj×BunG,rigZ0C^{1,-\alpha_{j}^{\vee}}_{P_{1}}(Z_{0})\lhook\joinrel\relbar\joinrel\rightarrow\mathrm{Bun}_{P_{1},rig}^{-\alpha_{j}^{\vee}}\times_{\mathrm{Bun}_{G,rig}}Z_{0}\lhook\joinrel\relbar\joinrel\rightarrow\mathrm{KM}_{P_{1},G,rig}^{-\alpha_{j}^{\vee}}\times_{\mathrm{Bun}_{G,rig}}Z_{0} (3.4.4)

is surjective, where, for 1kn0+11\leq k\leq n_{0}+1, wWP,Pk0w\in W^{0}_{P,P_{k}} and λ𝕏(TPk)\lambda\in\mathbb{X}_{*}(T_{P_{k}}), we write

CPkw,λ(Z0)=CP,Pk,rigw,λ×BunP,rigZ0,C^{w,\lambda}_{P_{k}}(Z_{0})=C^{w,\lambda}_{P,P_{k},rig}\times_{\mathrm{Bun}_{P,rig}}Z_{0},

where CP,Pk,rigw,λC^{w,\lambda}_{P,P_{k},rig} is the rigidification of the Bruhat cell CP,Pkw,λC^{w,\lambda}_{P,P_{k}} of §3.2. Lemma 3.4.3 below and Proposition 3.2.1 imply that the morphism

wWP,B0WL1λ=w1(αi+αj)Cw,λ(Z0)BunB,rigλ×BunG,rigZ0\coprod_{\begin{subarray}{c}w\in W^{0}_{P,B}\cap W_{L_{1}}\\ \lambda=-w^{-1}(\alpha_{i}^{\vee}+\alpha_{j}^{\vee})\end{subarray}}C^{w,\lambda}(Z_{0})\longrightarrow\mathrm{Bun}_{B,rig}^{\lambda}\times_{\mathrm{Bun}_{G,rig}}Z_{0}

is surjective for all λαj\lambda\leq-\alpha_{j}^{\vee}, where WL1W_{L_{1}} is the Weyl group of the Levi factor L1P1L_{1}\subseteq P_{1} and Cw,λ(Z0)=CPn0+1w,λ(Z0)C^{w,\lambda}(Z_{0})=C^{w,\lambda}_{P_{n_{0}+1}}(Z_{0}). Since the morphism KMB,GαjKMP1,Gαj\mathrm{KM}_{B,G}^{-\alpha_{j}^{\vee}}\to\mathrm{KM}_{P_{1},G}^{-\alpha_{j}^{\vee}} is also surjective by [davis19a, Proposition 3.6.4], and maps sections coming from Cw,λ(Z0)C^{w,\lambda}(Z_{0}) to CP11,αj(Z0)C^{1,-\alpha_{j}^{\vee}}_{P_{1}}(Z_{0}), surjectivity of (3.4.4) now follows. ∎

Lemma 3.4.3.

Assume that wWP,B0w\in W^{0}_{P,B}, λαj\lambda\leq-\alpha_{j}^{\vee} and Cw,λ(Z0)C^{w,\lambda}(Z_{0})\neq\emptyset. Then wWL1w\in W_{L_{1}} and λ=w1(αi+αj){αj,αiαj}\lambda=-w^{-1}(\alpha_{i}^{\vee}+\alpha_{j}^{\vee})\in\{-\alpha_{j}^{\vee},-\alpha_{i}^{\vee}-\alpha_{j}^{\vee}\}, where L1P1L_{1}\subseteq P_{1} is the standard Levi subgroup.

Proof.

It is immediate from Lemma 3.1.2 that λ{αj,αiαj}\lambda\in\{-\alpha_{j}^{\vee},-\alpha_{i}^{\vee}-\alpha_{j}^{\vee}\}. If Cw,λ(Z0)C^{w,\lambda}(Z_{0})\neq\emptyset, then there exists a geometric point z:SpeckZ0z\colon\operatorname{\mathrm{Spec}}k\to Z_{0} over s:SpeckSs\colon\operatorname{\mathrm{Spec}}k\to S and a section σL:EsξL,z/(LB)\sigma_{L}\colon E_{s}\to\xi_{L,z}/(L\cap B) of degree wλ𝕏(T)w\lambda\in\mathbb{X}_{*}(T). Since ξL,z\xi_{L,z} has slope μ\mu, we must have

ϖi,wλ=ϖi,μ=1.\langle\varpi_{i},w\lambda\rangle=\langle\varpi_{i},\mu\rangle=-1.

Since λ\lambda and hence wλw\lambda is a coroot, we therefore have wλΦ𝕏(T)w\lambda\in\Phi^{\vee}_{-}\subseteq\mathbb{X}_{*}(T)_{-}. Since composing σL\sigma_{L} with the inclusion ξL,z/(LB)ξG,z/B\xi_{L,z}/(L\cap B)\to\xi_{G,z}/B defines a section of degree wλw\lambda, we deduce that Dwλ(Z)D_{-w\lambda}(Z)\neq\emptyset, and hence that wλ{αi,αiαj}w\lambda\in\{-\alpha_{i}^{\vee},-\alpha_{i}^{\vee}-\alpha_{j}^{\vee}\}.

If wλ=αiw\lambda=-\alpha_{i}^{\vee}, then w1αiΦ+w^{-1}\alpha_{i}^{\vee}\in\Phi^{\vee}_{+}, so w=1w=1 since wWP,B0w\in W^{0}_{P,B}. So λ=αi\lambda=-\alpha_{i}^{\vee}, contradicting λαj\lambda\leq-\alpha_{j}^{\vee}. So we must have wλ=αiαjw\lambda=-\alpha_{i}^{\vee}-\alpha_{j}^{\vee}, and in particular w1(αi+αj)Φ+w^{-1}(\alpha_{i}^{\vee}+\alpha_{j}^{\vee})\in\Phi_{+}^{\vee}.

If (G,P,μ)(G,P,\mu) is not of type AA, then w1(αk)Φ+w^{-1}(\alpha_{k}^{\vee})\in\Phi_{+}^{\vee} for αkαi\alpha_{k}\neq\alpha_{i} (since wWP,B0w\in W^{0}_{P,B} and t(P)={αi}t(P)=\{\alpha_{i}\}) so Lemma 3.4.4 implies that wWL1w\in W_{L_{1}}. If (G,P,μ)(G,P,\mu) is of type AA, then w1(αk)Φ+w^{-1}(\alpha_{k}^{\vee})\in\Phi_{+}^{\vee} for αkαi,αj\alpha_{k}\neq\alpha_{i},\alpha_{j}. If w1(αj)Φ+w^{-1}(\alpha_{j}^{\vee})\in\Phi_{+}^{\vee} then wWL1w\in W_{L_{1}} by Lemma 3.4.4 again. Otherwise, we must have w1(αi)Φ+w^{-1}(\alpha_{i}^{\vee})\in\Phi_{+}^{\vee} and hence

w{si+1si+2ski<kl}w\in\{s_{i+1}s_{i+2}\cdots s_{k}\mid i<k\leq l\}

by Lemma 3.4.4. But this implies that λ=w1(αiαj)=w1(αiαi+1)=αi\lambda=w^{-1}(-\alpha_{i}^{\vee}-\alpha_{j}^{\vee})=w^{-1}(-\alpha_{i}^{\vee}-\alpha_{i+1}^{\vee})=-\alpha_{i}^{\vee}, contradicting λαj\lambda\leq-\alpha_{j}^{\vee}, so we are done. ∎

Lemma 3.4.4.

Let (M,Ψ,M,Ψ)(M,\Psi,M^{\vee},\Psi^{\vee}) be a root datum with Weyl group W(Ψ)W(\Psi), and let ΓΨ\Gamma\subseteq\Psi be a complete set of positive simple roots. Let βjΓ\beta_{j}\in\Gamma be a simple root, and let cπ0(Γ{βj})c\in\pi_{0}(\Gamma\setminus\{\beta_{j}\}) be a connected component of the Dynkin diagram of Γ{βj}\Gamma\setminus\{\beta_{j}\} of type AnA_{n} such that βj\beta_{j} is adjacent to one end of cc. Let βc,1,,βc,nΓ\beta_{c,1},\ldots,\beta_{c,n}\in\Gamma denote the nodes of cc, labelled so that βc,k\beta_{c,k} is adjacent to βc,k+1\beta_{c,k+1} for all kk and βc,n\beta_{c,n} is adjacent to βj\beta_{j}, and let

Σ={wW(Ψ)w1βkΨ+for allβkΓ{βc,n}andw1(βc,n+βj)Ψ+},\Sigma=\{w\in W(\Psi)\mid w^{-1}\beta_{k}^{\vee}\in\Psi^{\vee}_{+}\;\text{for all}\;\beta_{k}\in\Gamma\setminus\{\beta_{c,n}\}\;\text{and}\;w^{-1}(\beta_{c,n}^{\vee}+\beta_{j}^{\vee})\in\Psi^{\vee}_{+}\},

Then

Σ={1}{sc,nsc,n1sc,k1kn}\Sigma=\{1\}\cup\{s_{c,n}s_{c,n-1}\cdots s_{c,k}\mid 1\leq k\leq n\}

where sc,kW(Ψ)s_{c,k}\in W(\Psi) is the reflection in the root βc,k\beta_{c,k}

Proof.

First note that an easy inspection shows that

{1}{sc,nsc,n1sc,k1kn}Σ,\{1\}\cup\{s_{c,n}s_{c,n-1}\cdots s_{c,k}\mid 1\leq k\leq n\}\subseteq\Sigma,

so it suffices to prove the reverse inclusion.

We prove the claim by induction on n1n\geq 1. Suppose that wΣw\in\Sigma. Then either w=1w=1 or w1βc,nΨw^{-1}\beta_{c,n}\in\Psi_{-}. In the second case, we see that (sc,nw)1βkΨ+(s_{c,n}w)^{-1}\beta_{k}^{\vee}\in\Psi^{\vee}_{+} for βkΓ{βc,n1}\beta_{k}\in\Gamma\setminus\{\beta_{c,n-1}\} and (sc,nw)1(βc,n1+βc,n)Ψ+(s_{c,n}w)^{-1}(\beta_{c,n-1}^{\vee}+\beta_{c,n}^{\vee})\in\Psi^{\vee}_{+} if n>1n>1. So either n=1n=1 and w{1,sc,n}w\in\{1,s_{c,n}\}, or n>1n>1 and by induction we have

sc,nw{sc,n1sc,k1kn1},s_{c,n}w\in\{s_{c,n-1}\cdots s_{c,k}\mid 1\leq k\leq n-1\},

and hence

w{1}{sc,nsc,n1sc,k1kn}.w\in\{1\}\cup\{s_{c,n}s_{c,n-1}\cdots s_{c,k}\mid 1\leq k\leq n\}.

This proves the lemma. ∎

Proposition 3.4.5.

There exists a surjective homomorphism

πP1:P1GLn0+1\pi_{P_{1}}\colon P_{1}\longrightarrow GL_{n_{0}+1}

such that πP11(Rn0+1)=PP1\pi_{P_{1}}^{-1}(R_{n_{0}+1})=P\cap P_{1} and πP11(Qkn0+1)=Pk\pi_{P_{1}}^{-1}(Q^{n_{0}+1}_{k})=P_{k} for 1kn0+11\leq k\leq n_{0}+1, and such that the induced map T=TPn0+1Qn0+1n0+1T=T_{P_{n_{0}+1}}\to Q^{n_{0}+1}_{n_{0}+1} is given on cocharacters by

𝕏(T)\displaystyle\mathbb{X}_{*}(T) 𝕏(TQn0+1n0+1)\displaystyle\longrightarrow\mathbb{X}_{*}(T_{Q^{n_{0}+1}_{n_{0}+1}})
αc0,k\displaystyle\alpha_{c_{0},k}^{\vee} ekek+1\displaystyle\longmapsto e_{k}^{*}-e_{k+1}^{*}
αj\displaystyle\alpha_{j}^{\vee} en0+1\displaystyle\longmapsto e_{n_{0}+1}^{*}
αp\displaystyle\alpha_{p}^{\vee} 0,ifαp{αc0,1,,αc0,n0,αj}.\displaystyle\longmapsto 0,\quad\text{if}\;\;\alpha_{p}\notin\{\alpha_{c_{0},1},\ldots,\alpha_{c_{0},n_{0}},\alpha_{j}\}.
Proof.

Since the Dynkin diagram Δt(P1)\Delta\setminus t(P_{1}) has exactly one connected component of type An0A_{n_{0}}, Proposition 2.3.1 gives an embedding

L1⸦-→GLn0+1×𝔾mn1.L_{1}\lhook\joinrel\relbar\joinrel\rightarrow GL_{n_{0}+1}\times\mathbb{G}_{m}^{n_{1}}. (3.4.5)

Let πL1\pi_{L_{1}} be the composition of (3.4.5) with the projection to the first factor, and let πP1\pi_{P_{1}} be the composition of πL1\pi_{L_{1}} with the quotient P1L1P_{1}\to L_{1}. The remaining claims can now be checked routinely using the explicit isomorphism of Proposition 2.3.1. ∎

By construction, the morphism C1,n0+1C1,n0+1GLn0+1C_{1,n_{0}+1}\to C_{1,n_{0}+1}^{GL_{n_{0}+1}} factors through a morphism

C1,n0+1Yαj×YQn0+1n0+1en0+1(C1,n0+1GLn0+1×SE)=Y×YQn0+1n0+1(C1,n0+1GLn0+1×SE),C_{1,n_{0}+1}\longrightarrow Y^{-\alpha_{j}^{\vee}}\times_{Y_{Q^{n_{0}+1}_{n_{0}+1}}^{-e_{n_{0}+1}^{*}}}(C_{1,n_{0}+1}^{GL_{n_{0}+1}}\times_{S}E)=Y\times_{Y_{Q^{n_{0}+1}_{n_{0}+1}}}(C_{1,n_{0}+1}^{GL_{n_{0}+1}}\times_{S}E),

where the morphism C1,n0+1GLn0+1×SEYQn0+1n0+1C_{1,n_{0}+1}^{GL_{n_{0}+1}}\times_{S}E\to Y_{Q^{n_{0}+1}_{n_{0}+1}} is given by the natural morphism to YQn0+1n0+1en0+1×SEY_{Q^{n_{0}+1}_{n_{0}+1}}^{-e_{n_{0}+1}^{*}}\times_{S}E composed with (y,x)y+en0+1(x)(y,x)\mapsto y+e_{n_{0}+1}^{*}(x). Composing with the morphism (3.3.3) gives a morphism C1,n0+1Y×SE×SEC_{1,n_{0}+1}\to Y\times_{S}E\times_{S}E and hence a morphism

C1,n0+1Y×SE×SE\displaystyle C_{1,n_{0}+1}\longrightarrow Y\times_{S}E\times_{S}E Y×SPicS0(E)\displaystyle\longrightarrow Y\times_{S}\mathrm{Pic}^{0}_{S}(E) (3.4.6)
(y,xi,xj)\displaystyle(y,x_{i},x_{j}) (y,xjxi)\displaystyle\longmapsto(y,x_{j}-x_{i})

over YY. We remark that for the image of a stable map with two rational components of degree αi\alpha_{i}^{\vee} and αj\alpha_{j}^{\vee}, xix_{i} and xjx_{j} above are just the points of attachment of the two rational curves.

For 1pn0+11\leq p\leq n_{0}+1, we let

MpC1,n0+1M_{p}\subseteq C_{1,n_{0}+1}

be the closed substack given by the fibre product

Mp{M_{p}}C1,n0+1{C_{1,n_{0}+1}}Y{Y}Y×SPicS0(E).{Y\times_{S}\mathrm{Pic}^{0}_{S}(E).}(3.4.6)\scriptstyle{\eqref{eq:bruhatcellcurvecomparison1}}θp\scriptstyle{\theta_{p}}
Proposition 3.4.6.

For all 1kn01\leq k\leq n_{0}, the morphism Dk+1DkD_{k+1}\to D_{k} restricts to isomorphisms

Dk+1×Z(ZZ0)Dk×Z(ZZ0),Ck+1,n0+1Ck,n0+1andCk+1,pCk,pD_{k+1}\times_{Z}(Z\setminus Z_{0})\overset{\sim}{\longrightarrow}D_{k}\times_{Z}(Z\setminus Z_{0}),\quad C_{k+1,n_{0}+1}\overset{\sim}{\longrightarrow}C_{k,n_{0}+1}\quad\text{and}\quad C_{k+1,p}\overset{\sim}{\longrightarrow}C_{k,p}

for 1p<k1\leq p<k, and a morphism

Ck+1,kMkCk,n0+1C1,n0+1C_{k+1,k}\longrightarrow M_{k}\subseteq C_{k,n_{0}+1}\cong C_{1,n_{0}+1}

that realises Ck+1,kC_{k+1,k} as an 𝔸1\mathbb{A}^{1}-bundle over MkM_{k}.

Proof.

Chasing through the definitions, we have

Mk=C1,n0+1×C1,n0+1GLn0+1MkGLn0+1.M_{k}=C_{1,n_{0}+1}\times_{C_{1,n_{0}+1}^{GL_{n_{0}+1}}}M_{k}^{GL_{n_{0}+1}}.

Since the diagram

Dk+1×ZZ0{D_{k+1}\times_{Z}Z_{0}}Xk+1,rign0+1{X^{n_{0}+1}_{k+1,rig}}Dk×ZZ0{D_{k}\times_{Z}Z_{0}}Xk,rign0+1{X^{n_{0}+1}_{k,rig}}

is Cartesian, Proposition 3.3.6 implies everything except the claim that

Dk+1×Z(ZZ0)Dk×Z(ZZ0)D_{k+1}\times_{Z}(Z\setminus Z_{0})\longrightarrow D_{k}\times_{Z}(Z\setminus Z_{0}) (3.4.7)

is an isomorphism. To prove this, first note that every GG-bundle in the image of Dk×Z(ZZ0)D_{k}\times_{Z}(Z\setminus Z_{0}) is regular unstable, as follows from comparing the codimensions of its Z(L)rigZ(L)_{rig}-orbit in ZZ and in BunG,rig/E\mathrm{Bun}_{G,rig}/E. Since KMB,GαjKMPk,Gαj\mathrm{KM}_{B,G}^{-\alpha_{j}^{\vee}}\to\mathrm{KM}_{P_{k},G}^{-\alpha_{j}^{\vee}} is surjective for all kk by [davis19a, Proposition 3.6.4], all such bundles necessarily have Harder-Narasimhan reduction to the parabolic QQ of type t(Q)={αj}t(Q)=\{\alpha_{j}\} by [davis19, Lemma 4.3.4]. So the morphism to BunG,rig\mathrm{Bun}_{G,rig} factors as

Dk×Z(ZZ0)BunQ,rigss,αj⸦-→BunG,rig.D_{k}\times_{Z}(Z\setminus Z_{0})\longrightarrow\mathrm{Bun}_{Q,rig}^{ss,-\alpha_{j}^{\vee}}\lhook\joinrel\relbar\joinrel\rightarrow\mathrm{Bun}_{G,rig}.

The argument of the proof of [davis19, Proposition 4.3.8], shows that we have isomorphisms

Dk×Z(ZZ0)Y×YPk(BunMPk,rigαj×BunM,rigαjBunQ,rigss,αj×BunG,rig(ZZ0)×SE)D_{k}\times_{Z}(Z\setminus Z_{0})\cong Y\times_{Y_{P_{k}}}(\mathrm{Bun}_{M\cap P_{k},rig}^{-\alpha_{j}^{\vee}}\times_{\mathrm{Bun}_{M,rig}^{-\alpha_{j}^{\vee}}}\mathrm{Bun}_{Q,rig}^{ss,-\alpha_{j}^{\vee}}\times_{\mathrm{Bun}_{G,rig}}(Z\setminus Z_{0})\times_{S}E)

for all kk, where MM is the Levi factor of QQ (note that the first two factors in the parentheses above are just the Bruhat cell CPk,Q1,αjC^{1,-\alpha_{j}^{\vee}}_{P_{k},Q}). So Proposition 2.3.1 and [davis19, Lemma 4.3.7] show that (3.4.7) is an isomorphism as claimed. ∎

Lemma 3.4.7.

The morphism (3.4.6) is smooth with connected fibres.

Proof.

From the construction, we have

C1,n0+1=D1×ZZ0=Yαj×YP1αjBunLP1,rigαiαj×BunL,rigZ0×SE.C_{1,n_{0}+1}=D_{1}\times_{Z}Z_{0}=Y^{-\alpha_{j}^{\vee}}\times_{Y_{P_{1}}^{-\alpha_{j}^{\vee}}}\mathrm{Bun}_{L\cap P_{1},rig}^{-\alpha_{i}^{\vee}-\alpha_{j}^{\vee}}\times_{\mathrm{Bun}_{L,rig}}Z_{0}\times_{S}E.

There is an isomorphism

Y×YP1YLP1\displaystyle Y\times_{Y_{P_{1}}}Y_{L\cap P_{1}} Y×SPicS0(E)\displaystyle\overset{\sim}{\longrightarrow}Y\times_{S}\mathrm{Pic}^{0}_{S}(E) (3.4.8)
(y1,y2)\displaystyle(y_{1},y_{2}) (y1,ϖi(y2)ϖi(y1)).\displaystyle\longmapsto(y_{1},\varpi_{i}(y_{2})-\varpi_{i}(y_{1})).

Chasing through the definitions of the various morphisms involved, we deduce that there is a pullback

C1,n0+1{C_{1,n_{0}+1}}Y×SPicS0(E){Y\times_{S}\mathrm{Pic}^{0}_{S}(E)}BunLP1,rigαiαj×BunL,rigZ0×SE{\mathrm{Bun}_{L\cap P_{1},rig}^{-\alpha_{i}^{\vee}-\alpha_{j}^{\vee}}\times_{\mathrm{Bun}_{L,rig}}Z_{0}\times_{S}E}YLP1αiαj×SE{Y_{L\cap P_{1}}^{-\alpha_{i}^{\vee}-\alpha_{j}^{\vee}}\times_{S}E}YLP1,{Y_{L\cap P_{1}},}(3.4.6)\scriptstyle{\eqref{eq:bruhatcellcurvecomparison1}}

where the morphisms Y×SPicS0(E)YLP1Y\times_{S}\mathrm{Pic}^{0}_{S}(E)\to Y_{L\cap P_{1}} is the composition of the inverse to (3.4.8) with the natural projection.

It therefore suffices to show that the composition ff of the first two morphisms in the bottom row is smooth with connected fibres. Note that the morphism

YLP1αiαj×SEYLP1Y_{L\cap P_{1}}^{-\alpha_{i}^{\vee}-\alpha_{j}^{\vee}}\times_{S}E\longrightarrow Y_{L\cap P_{1}}

naturally identifies YLP1Y_{L\cap P_{1}} with the quotient (YLP1αiαj×SE)/E(Y_{L\cap P_{1}}^{-\alpha_{i}^{\vee}-\alpha_{j}^{\vee}}\times_{S}E)/E by the diagonal action of EE by translations. So we can identify ff with the composition of the middle vertical arrows in the diagram

BunLP1,rigαiαj×BunL,rigμZ0×SE{\mathrm{Bun}_{L\cap P_{1},rig}^{-\alpha_{i}^{\vee}-\alpha_{j}^{\vee}}\times_{\mathrm{Bun}_{L,rig}^{\mu}}Z_{0}\times_{S}E}Z0{Z_{0}}BunLP1,rigαiαj×SE{\mathrm{Bun}_{L\cap P_{1},rig}^{-\alpha_{i}^{\vee}-\alpha_{j}^{\vee}}\times_{S}E}(BunLP1,rigαiαj×SE)/E{(\mathrm{Bun}_{L\cap P_{1},rig}^{-\alpha_{i}^{\vee}-\alpha_{j}^{\vee}}\times_{S}E)/E}BunL,rigμ/E{\mathrm{Bun}_{L,rig}^{\mu}/E}YLP1αiαj×SE{Y_{L\cap P_{1}}^{-\alpha_{i}^{\vee}-\alpha_{j}^{\vee}}\times_{S}E}(YLP1αiαj×SE)/E.{(Y_{L\cap P_{1}}^{-\alpha_{i}^{\vee}-\alpha_{j}^{\vee}}\times_{S}E)/E.} (3.4.9)

The vertical arrow on the left in (3.4.9) is smooth, and has connected fibres since the semisimple part of LP1L\cap P_{1} is simply connected. The vertical arrow on the right in (3.4.9) is smooth with connected fibres by assumption. Since both squares are Cartesian, and the horizontal arrows in the square on the left are faithfully flat, it follows that both vertical arrows in the middle are smooth with connected fibres, and hence so is their composition ff. ∎

Lemma 3.4.8.

The morphism (3.4.6) is an isomorphism.

Proof.

Observe that the cell

Cn0+1,n0+1Dn0+1=KMB,G,rigαj×BunG,rigZ×SEC_{n_{0}+1,n_{0}+1}\subseteq D_{n_{0}+1}=\mathrm{KM}_{B,G,rig}^{-\alpha_{j}^{\vee}}\times_{\mathrm{Bun}_{G,rig}}Z\times_{S}E

is equal to the locus of singular domain curves, and is therefore a divisor in Dn0+1D_{n_{0}+1} flat over YY. Since Dn0+1YD_{n_{0}+1}\to Y has relative dimension 22, Cn0+1,n0+1YC_{n_{0}+1,n_{0}+1}\to Y therefore has relative dimension 11. So by Lemma 3.4.7, (3.4.6) is a smooth proper morphism with connected fibres and finite relative stabilisers between smooth stacks of the same dimension over SS. Since Cn0+1,n0+1SC_{n_{0}+1,n_{0}+1}\to S is representable over the dense open substack where Z0SZ_{0}\to S is representable (note that Z~Z\tilde{Z}\to Z is representable, and even projective since all stable maps involved in the definition have trivial automorphism group), so is Cn0+1,n0+1Y×SPicS0(E)C_{n_{0}+1,n_{0}+1}\to Y\times_{S}\mathrm{Pic}^{0}_{S}(E). Since Y×SPicS0(E)SY\times_{S}\mathrm{Pic}^{0}_{S}(E)\to S has irreducible fibres, (3.4.6) is therefore surjective, so by Lemma 3.4.9 below, it is an isomorphism as claimed. ∎

Lemma 3.4.9.

Let XX and XX^{\prime} be stacks that are smooth and of the same dimension over SS, and let f:XXf\colon X\to X^{\prime} be a smooth surjective proper morphism with connected fibres and finite relative stabilisers. Assume that there exists some open set UXU\subseteq X that is dense in every fibre of XSX\to S such that f|Uf|_{U} is representable. Then ff is an isomorphism.

Proof.

First note that f|U:UXf|_{U}\colon U\to X^{\prime} is étale and representable with connected fibres, and hence an open immersion. Moreover, the morphism X×XXXX\times_{X^{\prime}}X\to X is smooth with connected fibres, so the preimage of UU under either projection is dense. So the diagonal XX×XXX\to X\times_{X^{\prime}}X, which is finite by assumption, is an isomorphism over the dense open subset UU, and hence surjective. Since X×XXX\times_{X^{\prime}}X is smooth over SS, and hence normal, it follows that XX×XXX\to X\times_{X^{\prime}}X is an isomorphism. Since ff is smooth and surjective, by flat descent it follows that f:XXf\colon X\to X^{\prime} is also an isomorphism as claimed. ∎

Lemma 3.4.10.

Let UU be a regular stack, let XUX\to U and XUX^{\prime}\to U be smooth representable morphisms of relative dimension 22, and let f:XXf\colon X\to X^{\prime} be a projective morphism over UU. Suppose that there exists a section g:UXg\colon U\to X^{\prime} such that f1(Xg(U))Xg(U)f^{-1}(X^{\prime}\setminus g(U))\to X^{\prime}\setminus g(U) is an isomorphism, and such that every fibre of ff over a point in g(U)g(U) is an irreducible curve. Then ff is the blowup of XX^{\prime} along g(U)g(U).

Proof.

Since the claim is local in the smooth topology on UU and in the étale topology on XX^{\prime}, we can reduce to the case where XUX^{\prime}\to U is a smooth morphism of schemes with UU connected and regular.

First note that the underlying reduced scheme DD of the exceptional locus f1(g(U))f^{-1}(g(U)) is an integral closed subscheme of codimension 11 in a regular scheme, and hence a Cartier divisor. Since XX and XX^{\prime} are smooth over UU and ff is an isomorphism outside DD, we therefore have KX/U=fKX/U(nD)K_{X/U}=f^{*}K_{X^{\prime}/U}(nD) for some n>0n>0. If kk is any field and u:SpeckUu\colon\operatorname{\mathrm{Spec}}k\to U is a kk-point, we have D|Xu=muCuD|_{X_{u}}=m_{u}C_{u} for some mu>0m_{u}>0, where CuXuC_{u}\subseteq X_{u} is the irreducible curve contracted under ff, and hence, by adjunction

2degKCu=(mun+1)Cu2.-2\leq\deg K_{C_{u}}=(m_{u}n+1)C_{u}^{2}.

Since Cu2<0C_{u}^{2}<0, we deduce that mu=n=1m_{u}=n=1, Cu2=1C_{u}^{2}=-1, degKCu=2\deg K_{C_{u}}=-2, and hence that CuC_{u} is a smooth rational curve. In particular, by Castelnuovo’s theorem, fu:XuXuf_{u}\colon X_{u}\to X^{\prime}_{u} is the blowup at g(u)g(u).

Now let η:SpecKU\eta\colon\operatorname{\mathrm{Spec}}K\to U be the generic point of UU. We have shown that on the generic fibre fη:XηXηf_{\eta}\colon X_{\eta}\to X^{\prime}_{\eta} is the blowup along g(η)g(\eta), so the same must be true on some dense open set UU with complement VV. So we get an isomorphism

h:Xf1(g(V))X~π1(g(V))h\colon X\setminus f^{-1}(g(V))\overset{\sim}{\longrightarrow}\tilde{X}^{\prime}\setminus\pi^{-1}(g(V))

over XX^{\prime}, where π:X~X\pi\colon\tilde{X}^{\prime}\to X^{\prime} is the blowup of XX^{\prime} along g(U)g(U). Since ff is projective and is an isomorphism outside DD, it follows that either DD or D-D is ff-ample. Since DCu=(Cu2)Xu=1D\cdot C_{u}=(C_{u}^{2})_{X_{u}}=-1 for all points u:SpeckXu\colon\operatorname{\mathrm{Spec}}k\to X^{\prime}, it follows that D-D is ff-ample. But hh is an isomorphism in codimension 11 between regular schemes projective over XX^{\prime}, h(Df1(g(V)))=π1(g(U))π1(g(V))h(D\setminus f^{-1}(g(V)))=\pi^{-1}(g(U))\setminus\pi^{-1}(g(V)), and π1(g(U))-\pi^{-1}(g(U)) is ff-ample, so

XProjXd0f𝒪(dD)ProjXd0π𝒪(dπ1(g(U)))X~,X\overset{\sim}{\longrightarrow}\mathrm{Proj}_{X^{\prime}}\bigoplus_{d\geq 0}f_{*}\mathcal{O}(-dD)\cong\mathrm{Proj}_{X^{\prime}}\bigoplus_{d\geq 0}\pi_{*}\mathcal{O}(-d\pi^{-1}(g(U)))\overset{\sim}{\longleftarrow}\tilde{X}^{\prime},

which proves that XX is the blowup as claimed. ∎

3.5. The divisor Dαi+αj(Z)D_{\alpha_{i}^{\vee}+\alpha_{j}^{\vee}}(Z)

In this subsection, we prove the following proposition, which is essentially Theorem 1.0.3 (3).

Proposition 3.5.1.

Assume we are in the setup of Proposition 3.1.1. Then every fibre of the morphism

Dαi+αj(Z)YD_{\alpha_{i}^{\vee}+\alpha_{j}^{\vee}}(Z)\longrightarrow Y

is isomorphic to the Hirzebruch surface 𝔽d1\mathbb{F}_{d-1}.

Proof.

By Proposition 3.1.1, we have Dλ(Z)=D_{\lambda}(Z)=\emptyset for all λ>αi+αj\lambda>\alpha_{i}^{\vee}+\alpha_{j}^{\vee}, so any stable map parametrised by a point in Dαi+αj(Z)D_{\alpha_{i}^{\vee}+\alpha_{j}^{\vee}}(Z) must be the union of a section of the relevant G/BG/B-bundle of degree αiαj-\alpha_{i}^{\vee}-\alpha_{j}^{\vee} and a single connected stable map of genus 0 and degree αi+αj\alpha_{i}^{\vee}+\alpha_{j}^{\vee} to a fibre of the G/BG/B-bundle. We deduce that

Dαi+αj(Z)η×B/Z(G)M¯0,1+(G/B,αi+αj),D_{\alpha_{i}^{\vee}+\alpha_{j}^{\vee}}(Z)\cong\eta\times^{B/Z(G)}\bar{M}_{0,1}^{+}(G/B,\alpha_{i}^{\vee}+\alpha_{j}^{\vee}),

where

ηBunB,rigαiαj×BunG,rigZ×SE\eta\longrightarrow\mathrm{Bun}_{B,rig}^{-\alpha_{i}^{\vee}-\alpha_{j}^{\vee}}\times_{\mathrm{Bun}_{G,rig}}Z\times_{S}E

is the pullback of the universal B/Z(G)B/Z(G)-bundle on BunB,rig×SE\mathrm{Bun}_{B,rig}\times_{S}E, and

M¯0,1+(G/B,αi+αj)\bar{M}_{0,1}^{+}(G/B,\alpha_{i}^{\vee}+\alpha_{j}^{\vee})

is the moduli space of 11-pointed stable maps of genus 0 and degree αi+αj\alpha_{i}^{\vee}+\alpha_{j}^{\vee} sending the marked point to the base point B/BG/BB/B\in G/B.The morphism Dαi+αj(Z)YD_{\alpha_{i}^{\vee}+\alpha_{j}^{\vee}}(Z)\to Y factors through

BunB,rigαiαj×BunG,rigZ×SEΔEBunB,rigαiαj×BunG,rigZ×SE×SE=Cn0+1,n0+1Y,\mathrm{Bun}_{B,rig}^{-\alpha_{i}^{\vee}-\alpha_{j}^{\vee}}\times_{\mathrm{Bun}_{G,rig}}Z\times_{S}E\xrightarrow{\Delta_{E}}\mathrm{Bun}_{B,rig}^{-\alpha_{i}^{\vee}-\alpha_{j}^{\vee}}\times_{\mathrm{Bun}_{G,rig}}Z\times_{S}E\times_{S}E=C_{n_{0}+1,n_{0}+1}\longrightarrow Y, (3.5.1)

where the last morphism is the map Cn0+1,n0+1C1,n0+1C_{n_{0}+1,n_{0}+1}\to C_{1,n_{0}+1} composed with (3.4.6) and the projection to YY. Proposition 3.4.6 and Lemma 3.4.8 identify the last morphism with Y×SPicS0(E)YY\times_{S}\mathrm{Pic}^{0}_{S}(E)\to Y and the first with the zero section. So (3.5.1) is an isomorphism, so we can identify Dαi+αj(Z)D_{\alpha_{i}^{\vee}+\alpha_{j}^{\vee}}(Z) with the morphism

η×B/Z(G)M¯0,1+(G/B,αi+αj)Y\eta\times^{B/Z(G)}\bar{M}_{0,1}^{+}(G/B,\alpha_{i}^{\vee}+\alpha_{j}^{\vee})\to Y

for some B/Z(G)B/Z(G)-bundle ηY\eta\to Y. The proposition now follows from Proposition 3.5.2 below. ∎

Proposition 3.5.2.

With notation as in the proof of Proposition 3.5.1, there is an isomorphism

M¯0,1+(G/B,αi+αj)𝔽d1.\bar{M}^{+}_{0,1}(G/B,\alpha_{i}^{\vee}+\alpha_{j}^{\vee})\cong\mathbb{F}_{d-1}.

such that the closure of the locus of stable maps with dual graph

αi\scriptstyle{\alpha_{i}^{\vee}}αj\scriptstyle{\alpha_{j}^{\vee}}(resp.αj\scriptstyle{\alpha_{j}^{\vee}}αi\scriptstyle{\alpha_{i}^{\vee}})

is a fibre of 𝔽d11\mathbb{F}_{d-1}\to\mathbb{P}^{1} (resp. a section 1𝔽d1\mathbb{P}^{1}\to\mathbb{F}_{d-1} with self-intersection 1d1-d).

An important role in the proof of Proposition 3.5.2 is played by the Schubert varieties in G/BG/B. Given wWw\in W, recall that the Schubert variety associated to ww is the closed subvariety

Xw=BwB/B¯G/B.X_{w}=\overline{BwB/B}\subseteq G/B.

In what follows, we write Qi,QjGQ_{i},Q_{j}\subseteq G for the standard minimal parabolics of types t(Qi)=Δ{αi}t(Q_{i})=\Delta\setminus\{\alpha_{i}\} and t(Qj)=Δ{αj}t(Q_{j})=\Delta\setminus\{\alpha_{j}\}.

Lemma 3.5.3.

There are isomorphisms

Xsisj𝔽d,(resp.Xsjsi𝔽1)X_{s_{i}s_{j}}\cong\mathbb{F}_{d},\quad\text{(resp.}\quad X_{s_{j}s_{i}}\cong\mathbb{F}_{1}\;\;\text{)}

such that XsjX_{s_{j}} is identified with a fibre of 𝔽d1\mathbb{F}_{d}\to\mathbb{P}^{1} (resp., the unique section 1𝔽1\mathbb{P}^{1}\to\mathbb{F}_{1} of self-intersection 1-1) and XsiX_{s_{i}} is identified with the unique section 1𝔽d\mathbb{P}^{1}\to\mathbb{F}_{d} of self-intersection d-d (resp., a fibre of 𝔽11\mathbb{F}_{1}\to\mathbb{P}^{1}).

Proof.

We prove the claim for XsisjX_{s_{i}s_{j}}; the proof for XsjsiX_{s_{j}s_{i}} is identical after noting that αi,αj=1\langle\alpha_{i},\alpha_{j}^{\vee}\rangle=-1.

There is an isomorphism

SL2×BSL2,ραiQj/B=Qi×BQj/BXsisj,SL_{2}\times^{B_{SL_{2}},\rho_{\alpha_{i}}}Q_{j}/B=Q_{i}\times^{B}Q_{j}/B\overset{\sim}{\longrightarrow}X_{s_{i}s_{j}},

given by multiplication, where BSL2SL2B_{SL_{2}}\subseteq SL_{2} is the Borel subgroup of lower triangular matrices, and ραi:SL2G\rho_{\alpha_{i}}\colon SL_{2}\to G is the root homomorphism corresponding to αi\alpha_{i}. We also have an isomorphism of QjQ_{j}-varieties Qj/B(V)Q_{j}/B\cong\mathbb{P}(V^{\vee}), where VV is the QjQ_{j}-representation V=IndBQj(ϖj)V=\mathrm{Ind}_{B}^{Q_{j}}(\mathbb{Z}_{\varpi_{j}}), and an exact sequence

0ϖjαjVϖj00\longrightarrow\mathbb{Z}_{\varpi_{j}-\alpha_{j}}\longrightarrow V\longrightarrow\mathbb{Z}_{\varpi_{j}}\longrightarrow 0

of BB-representations, which splits uniquely as an exact sequence of BSL2B_{SL_{2}}-representations. So we have

Xsisj=SL2×BSL2(V)=1(𝒪(ϖj,αi)𝒪(ϖjαj,αi))=1(𝒪𝒪(d)))=𝔽d.X_{s_{i}s_{j}}=SL_{2}\times^{B_{SL_{2}}}\mathbb{P}(V^{\vee})=\mathbb{P}_{\mathbb{P}^{1}}(\mathcal{O}(-\langle\varpi_{j},\alpha_{i}^{\vee}\rangle)\oplus\mathcal{O}(-\langle\varpi_{j}-\alpha_{j},\alpha_{i}^{\vee}\rangle))=\mathbb{P}_{\mathbb{P}^{1}}(\mathcal{O}\oplus\mathcal{O}(-d)))=\mathbb{F}_{d}.

(Recall that ()\mathbb{P}(-) denotes the projective space of 11-dimensional subspaces or rank 11 subbundles of a vector space or vector bundle.) The identifications of Xsi=Qi/BX_{s_{i}}=Q_{i}/B and Xsj=Qj/BX_{s_{j}}=Q_{j}/B under this isomorphism follow immediately. ∎

Lemma 3.5.4.

The partial Schubert variety Xsisj/Qi=BsisjQi/Qi¯G/QiX_{s_{i}s_{j}}/Q_{i}=\overline{Bs_{i}s_{j}Q_{i}/Q_{i}}\subseteq G/Q_{i} is isomorphic to the projective cone ^d1\widehat{\mathbb{P}}^{1}_{d} on 1\mathbb{P}^{1} of degree dd, and the morphism

XsisjXsisj/QiX_{s_{i}s_{j}}\longrightarrow X_{s_{i}s_{j}}/Q_{i} (3.5.2)

is the blowup of Xsisj/QiX_{s_{i}s_{j}}/Q_{i} at the origin Qi/QiQ_{i}/Q_{i}.

Proof.

First note that the morphisms BsisjB/BBsisjQi/QiBs_{i}s_{j}B/B\to Bs_{i}s_{j}Q_{i}/Q_{i} and BsjB/BBsjQi/QiBs_{j}B/B\to Bs_{j}Q_{i}/Q_{i} are isomorphisms. So (3.5.2) is birational and finite outside Qi/QiQ_{i}/Q_{i}, and hence an isomorphism outside Qi/QiQ_{i}/Q_{i} since partial Schubert varieties are always normal. Since the preimage of Qi/QiQ_{i}/Q_{i} under (3.5.2) is Qi/B=XsiQ_{i}/B=X_{s_{i}}, using normality of Xsisj/QiX_{s_{i}s_{j}}/Q_{i} and of ^d1\widehat{\mathbb{P}}^{1}_{d}, we can conclude from Lemma 3.5.3 that (3.5.2) can be identified with the morphism

𝔽d^d1\mathbb{F}_{d}\longrightarrow\widehat{\mathbb{P}}^{1}_{d}

contracting the curve of self-intersection d-d. But this is indeed the blowup at the cone point, so we are done. ∎

Lemma 3.5.5.

There is a QiQ_{i}-equivariant isomorphism

M¯0,1+(Xsisj/Qi,αj)Qi/B1,\bar{M}^{+}_{0,1}(X_{s_{i}s_{j}}/Q_{i},\alpha_{j}^{\vee})\cong Q_{i}/B\cong\mathbb{P}^{1},

identifying the universal stable map with

Qi×BQj/BXsisjXsisj/Qi.Q_{i}\times^{B}Q_{j}/B\longrightarrow X_{s_{i}s_{j}}\longrightarrow X_{s_{i}s_{j}}/Q_{i}. (3.5.3)
Proof.

Assume that UU is a scheme and (f:CXsisj/Qi,x:UC)(f\colon C\to X_{s_{i}s_{j}}/Q_{i},x\colon U\to C) is a 11-pointed stable map over UU of degree αj\alpha_{j}^{\vee} sending xx to the base point. We need to show that there is a unique morphism UQi/BU\to Q_{i}/B such that ff and xx are the pullbacks of (3.5.3) and the canonical section Qi/B=Qi×BB/BQi×BQj/BQ_{i}/B=Q_{i}\times^{B}B/B\to Q_{i}\times^{B}Q_{j}/B.

We first claim that CUC\to U is smooth and that every geometric fibre of f1(Qi/Qi)Uf^{-1}(Q_{i}/Q_{i})\to U is a reduced point. Since f1(Qi/Qi)Uf^{-1}(Q_{i}/Q_{i})\to U has a section xx, it then follows that it is an isomorphism.

To prove the claim, fix a geometric point u:SpeckUu\colon\operatorname{\mathrm{Spec}}k\to U, and consider the stable map fu:Cu(Xsisj/Qi)kf_{u}\colon C_{u}\to(X_{s_{i}s_{j}}/Q_{i})_{k}. Since αj\alpha_{j}^{\vee} is not the sum of two nonzero effective curve classes, it follows that CuC_{u} is irreducible, hence smooth over Speck\operatorname{\mathrm{Spec}}k. Since this holds for all geometric points, CUC\to U is smooth as claimed, and fu1(Qi/Qi)f_{u}^{-1}(Q_{i}/Q_{i}) is a Cartier divisor on CuC_{u}. So by Lemmas 3.5.3 and 3.5.4, fuf_{u} lifts to a morphism f¯u:Cu(Xsisj)k(𝔽d)k\bar{f}_{u}\colon C_{u}\to(X_{s_{i}s_{j}})_{k}\cong(\mathbb{F}_{d})_{k} such that CuXsi>0C_{u}\cdot X_{s_{i}}>0 and Cu(dXsj+Xsi)=1C_{u}\cdot(dX_{s_{j}}+X_{s_{i}})=1. (Note that dXsj+XsidX_{s_{j}}+X_{s_{i}} is linearly equivalent to the pullback of ϖj\mathcal{L}_{\varpi_{j}}.) Since d>0d>0, it follows that CuXsi=1C_{u}\cdot X_{s_{i}}=1 and CuXsj=0C_{u}\cdot X_{s_{j}}=0. In particular, fu1(Qi/Qi)=CuXsif_{u}^{-1}(Q_{i}/Q_{i})=C_{u}\cap X_{s_{i}} is a reduced closed point on CuC_{u}, so fu1(Qi/Qi)Speckf_{u}^{-1}(Q_{i}/Q_{i})\cong\operatorname{\mathrm{Spec}}k as claimed.

Since f1(Qi/Qi)Cf^{-1}(Q_{i}/Q_{i})\subseteq C is a section of the smooth curve CUC\to U, it is a Cartier divisor, so by Lemma 3.5.4, ff lifts uniquely to a morphism f¯:CXsisj\bar{f}\colon C\to X_{s_{i}s_{j}}. Since the above argument shows that the composition f¯:CXsisj=Qi×BQj/BQi/B\bar{f}\colon C\to X_{s_{i}s_{j}}=Q_{i}\times^{B}Q_{j}/B\to Q_{i}/B has degree 0 on every fibre, this descends to a unique morphism UQi/BU\to Q_{i}/B. The induced morphism

CU×Qi/B(Qi×BQj/B)C\longrightarrow U\times_{Q_{i}/B}(Q_{i}\times^{B}Q_{j}/B) (3.5.4)

has degree 11 on every fibre and is therefore an isomorphism. Since (3.5.4) sends the section xx to the section Qi/BQi×BQj/BQ_{i}/B\to Q_{i}\times^{B}Q_{j}/B (as both are the preimage of Qi/QiXsisj/QiQ_{i}/Q_{i}\subseteq X_{s_{i}s_{j}}/Q_{i}), this proves the lemma. ∎

Proof of Proposition 3.5.2.

For the sake of brevity, write

M=M¯0,1+(G/B,αi+αj).M=\bar{M}_{0,1}^{+}(G/B,\alpha_{i}^{\vee}+\alpha_{j}^{\vee}).

We first claim that MM is connected. To see this, observe that BB acts on MM, that any BB-fixed point corresponds to a stable map factoring through XsiXsjG/BX_{s_{i}}\cup X_{s_{j}}\subseteq G/B, and that there is a unique such pointed stable map of class αi+αj\alpha_{i}^{\vee}+\alpha_{j}^{\vee} defined over kk for any algebraically closed field kk. Since every connected component of MM must have at least one BB-fixed point over every algebraically closed field, connectedness of MM follows immediately.

We now compute the closed subscheme

M=M¯0,1+(Xsisjsi,αi+αj)MM^{\prime}=\bar{M}_{0,1}^{+}(X_{s_{i}s_{j}s_{i}},\alpha_{i}^{\vee}+\alpha_{j}^{\vee})\subseteq M

consisting of stable maps factoring through the Schubert variety XsisjsiX_{s_{i}s_{j}s_{i}}. We will show that M𝔽dM^{\prime}\cong\mathbb{F}_{d} is smooth and projective of relative dimension 22 over Spec\operatorname{\mathrm{Spec}}\mathbb{Z}. Since the same is true for MM and MM is connected, it follows that M=MM^{\prime}=M.

Since Xsisjsi/Qi=Xsisj/QiX_{s_{i}s_{j}s_{i}}/Q_{i}=X_{s_{i}s_{j}}/Q_{i}, by Lemma 3.5.5 we have a morphism

MM¯0,1+(Xsisj/Qi,αj)Qi/B=1M^{\prime}\longrightarrow\bar{M}_{0,1}^{+}(X_{s_{i}s_{j}}/Q_{i},\alpha_{j}^{\vee})\cong Q_{i}/B=\mathbb{P}^{1}

sending a stable map to the stabilisation of its composition with G/BG/QiG/B\to G/Q_{i}. The pullback of the universal domain curve of M¯0,1+(Xsisj/Qi,αj)\bar{M}_{0,1}^{+}(X_{s_{i}s_{j}}/Q_{i},\alpha_{j}^{\vee}) along XsisjsiXsisj/QiX_{s_{i}s_{j}s_{i}}\to X_{s_{i}s_{j}}/Q_{i} is

Xsisjsi×Xsisj/Qi(Qi×BQj/B)=G/B×G/Qi(Qi×BQj/B),X_{s_{i}s_{j}s_{i}}\times_{X_{s_{i}s_{j}}/Q_{i}}(Q_{i}\times^{B}Q_{j}/B)=G/B\times_{G/Q_{i}}(Q_{i}\times^{B}Q_{j}/B),

which is identified with the Bott-Samelson variety X~sisjsi\tilde{X}_{s_{i}s_{j}s_{i}} via

X~sisjsi=Qi×BQj×BQi/B\displaystyle\tilde{X}_{s_{i}s_{j}s_{i}}=Q_{i}\times^{B}Q_{j}\times^{B}Q_{i}/B G/B×G/Qi(Qi×BQj/B)\displaystyle\overset{\sim}{\longrightarrow}G/B\times_{G/Q_{i}}(Q_{i}\times^{B}Q_{j}/B)
(g1,g2,g3B)\displaystyle(g_{1},g_{2},g_{3}B) (g1g2g3B,(g1,g2B)).\displaystyle\longmapsto(g_{1}g_{2}g_{3}B,(g_{1},g_{2}B)).

So we can identify MM^{\prime} with the relative space of stable maps

MM¯0,1,Qi/B+(X~sisjsi,αi+αj),M^{\prime}\cong\bar{M}_{0,1,Q_{i}/B}^{+}(\tilde{X}_{s_{i}s_{j}s_{i}},\alpha_{i}^{\vee}+\alpha_{j}^{\vee}),

where M¯0,1,Qi/B+(X~sisjsi,αi+αj)\bar{M}_{0,1,Q_{i}/B}^{+}(\tilde{X}_{s_{i}s_{j}s_{i}},\alpha_{i}^{\vee}+\alpha_{j}^{\vee}) is the fibre product

M¯0,1,Qi/B+(X~sisjsi,αi+αj){\bar{M}_{0,1,Q_{i}/B}^{+}(\tilde{X}_{s_{i}s_{j}s_{i}},\alpha_{i}^{\vee}+\alpha_{j}^{\vee})}Qi/B{Q_{i}/B}M¯0,1,Qi/B(X~sisjsi,αi+αj){\bar{M}_{0,1,Q_{i}/B}(\tilde{X}_{s_{i}s_{j}s_{i}},\alpha_{i}^{\vee}+\alpha_{j}^{\vee})}X~sisjsi.{\tilde{X}_{s_{i}s_{j}s_{i}}.}σ\scriptstyle{\sigma}

Here σ\sigma is the section defined by Qi/Bm1(B/B)X~sisjsiQ_{i}/B\cong m^{-1}(B/B)\to\tilde{X}_{s_{i}s_{j}s_{i}}, for m:X~sisjsiG/Bm\colon\tilde{X}_{s_{i}s_{j}s_{i}}\to G/B the natural morphism given by multiplication. Note that M¯0,1,Qi/B(X~sisjsi,αi+αj)\bar{M}_{0,1,Q_{i}/B}(\tilde{X}_{s_{i}s_{j}s_{i}},\alpha_{i}^{\vee}+\alpha_{j}^{\vee}) is naturally identified with the universal domain curve over the space M¯0,Qi/B(X~sisjsi,αi+αj)\bar{M}_{0,Q_{i}/B}(\tilde{X}_{s_{i}s_{j}s_{i}},\alpha_{i}^{\vee}+\alpha_{j}^{\vee}) of unpointed stable maps.

By Lemma 3.5.3, every fibre of X~sisjsjQi/B\tilde{X}_{s_{i}s_{j}s_{j}}\to Q_{i}/B is isomorphic to 𝔽1=Xsjsi=Qj×BQi/B\mathbb{F}_{1}=X_{s_{j}s_{i}}=Q_{j}\times^{B}Q_{i}/B, and αi+αj\alpha_{i}^{\vee}+\alpha_{j}^{\vee} is the class Xsi+XsjX_{s_{i}}+X_{s_{j}} of the (1)(-1)-curve plus a fibre of 𝔽11\mathbb{F}_{1}\to\mathbb{P}^{1}. Unpointed stable maps of class αi+αj\alpha_{i}^{\vee}+\alpha_{j}^{\vee} are the same things as closed subschemes with ideal sheaf 𝒪(XsiXsj)=mϖi\mathcal{O}(-X_{s_{i}}-X_{s_{j}})=m^{*}\mathcal{L}_{-\varpi_{i}}. So we can identify M0,Qi/B(X~sisjsi,αi+αj)M_{0,Q_{i}/B}(\tilde{X}_{s_{i}s_{j}s_{i}},\alpha_{i}^{\vee}+\alpha_{j}^{\vee}) with the Hilbert scheme Qi/B(πmϖi)\mathbb{P}_{Q_{i}/B}(\pi_{*}m^{*}\mathcal{L}_{\varpi_{i}}) and MM^{\prime} with the closed subscheme

M=Qi/B(kerπmϖiσmϖi)M^{\prime}=\mathbb{P}_{Q_{i}/B}(\ker\pi_{*}m^{*}\mathcal{L}_{\varpi_{i}}\to\sigma^{*}m^{*}\mathcal{L}_{\varpi_{i}})

of curves meeting σ(Qi/B)\sigma(Q_{i}/B), where π:X~sisjsiQi/B\pi\colon\tilde{X}_{s_{i}s_{j}s_{i}}\to Q_{i}/B is the natural projection.

It therefore remains to identify the vector bundle πmϖi\pi_{*}m^{*}\mathcal{L}_{\varpi_{i}} on Qi/B1Q_{i}/B\cong\mathbb{P}^{1} and the morphism πmϖiσmϖi=𝒪\pi_{*}m^{*}\mathcal{L}_{\varpi_{i}}\to\sigma^{*}m^{*}\mathcal{L}_{\varpi_{i}}=\mathcal{O}. It is clear from the identification X~sisjsi=Qi×BQj×BQi/B\tilde{X}_{s_{i}s_{j}s_{i}}=Q_{i}\times^{B}Q_{j}\times^{B}Q_{i}/B that πmϖi\pi_{*}m^{*}\mathcal{L}_{\varpi_{i}} is the QiQ_{i}-linearised vector bundle associated to the BB-representation

V=IndBQjIndBQiϖi.V=\mathrm{Ind}_{B}^{Q_{j}}\mathrm{Ind}_{B}^{Q_{i}}\mathbb{Z}_{\varpi_{i}}.

The representation VV has rank 33, with weights ϖi\varpi_{i}, ϖiαi\varpi_{i}-\alpha_{i} and ϖiαiαj\varpi_{i}-\alpha_{i}-\alpha_{j}, and restricting VV to a BSL2B_{SL_{2}}-representation via the root homomorphism ραi:SL2QiG\rho_{\alpha_{i}}\colon SL_{2}\to Q_{i}\subseteq G, we have

V=Uϖiαiαj,αi=Ud1,V=U\oplus\mathbb{Z}_{\langle\varpi_{i}-\alpha_{i}-\alpha_{j},\alpha_{i}^{\vee}\rangle}=U\oplus\mathbb{Z}_{d-1},

where UU is the standard representation of SL2SL_{2} and d1\mathbb{Z}_{d-1} is the rank 11 BSL2B_{SL_{2}}-module of weight d1d-1. So we get

πmϖi=U𝒪1𝒪(d1).\pi_{*}m^{*}\mathcal{L}_{\varpi_{i}}=U\otimes\mathcal{O}_{\mathbb{P}^{1}}\oplus\mathcal{O}(d-1).

Since d>0d>0, the kernel of

πmϖi=𝒪𝒪𝒪(d1)𝒪=σmϖi\pi_{*}m^{*}\mathcal{L}_{\varpi_{i}}=\mathcal{O}\oplus\mathcal{O}\oplus\mathcal{O}(d-1)\longrightarrow\mathcal{O}=\sigma^{*}m^{*}\mathcal{L}_{\varpi_{i}}

must be isomorphic to 𝒪𝒪(d1)\mathcal{O}\oplus\mathcal{O}(d-1), which gives the desired isomorphism M=M𝔽d1M=M^{\prime}\cong\mathbb{F}_{d-1}.

Finally, to identify the loci of stable maps with given dual graphs in the statement of the proposition, notice that each closure is isomorphic to 1\mathbb{P}^{1} (since there are unique curves of classes αi\alpha_{i}^{\vee} and αj\alpha_{j}^{\vee} through every point in G/BG/B), and that the closure of curves with dual graph

αi\scriptstyle{\alpha_{i}^{\vee}}αj\scriptstyle{\alpha_{j}^{\vee}}

is contracted under the map to M0,1+(G/Qi,αj)M_{0,1}^{+}(G/Q_{i},\alpha_{j}^{\vee}), and is hence a fibre of 𝔽d11\mathbb{F}_{d-1}\to\mathbb{P}^{1} as claimed. For the other statement, note that the map

πmϖiπ(σ)mϖi\pi_{*}m^{*}\mathcal{L}_{\varpi_{i}}\longrightarrow\pi^{\prime}_{*}(\sigma^{\prime})^{*}m^{*}\mathcal{L}_{\varpi_{i}}

is just the quotient map U𝒪1𝒪(d1)U𝒪1U\otimes\mathcal{O}_{\mathbb{P}^{1}}\oplus\mathcal{O}(d-1)\to U\otimes\mathcal{O}_{\mathbb{P}^{1}}, where σ\sigma^{\prime} is the morphism Qi×BQi/B=Qi×BB×BQi/BX~sisjsiQ_{i}\times^{B}Q_{i}/B=Q_{i}\times^{B}B\times^{B}Q_{i}/B\to\tilde{X}_{s_{i}s_{j}s_{i}} and π:Qi×BQi/BQi/B\pi^{\prime}\colon Q_{i}\times^{B}Q_{i}/B\to Q_{i}/B is the natural projection onto the first factor. So the subscheme

Qi/B(kerπmϖiπ(σ)mϖi)Qi/B(kerπmϖiσmϖi)=M\mathbb{P}_{Q_{i}/B}(\ker\pi_{*}m^{*}\mathcal{L}_{\varpi_{i}}\to\pi^{\prime}_{*}(\sigma^{\prime})^{*}m^{*}\mathcal{L}_{\varpi_{i}})\subseteq\mathbb{P}_{Q_{i}/B}(\ker\pi_{*}m^{*}\mathcal{L}_{\varpi_{i}}\to\sigma^{*}m^{*}\mathcal{L}_{\varpi_{i}})=M^{\prime}

is the canonical section of 𝔽d1\mathbb{F}_{d-1} of degree 1d1-d. But this parametrises curves of class αi+αj\alpha_{i}^{\vee}+\alpha_{j}^{\vee} containing some curve of class αi\alpha_{i}^{\vee}, so this must be the closure of the locus of curves with dual graph

αj\scriptstyle{\alpha_{j}^{\vee}}αi\scriptstyle{\alpha_{i}^{\vee}}

as claimed. ∎

3.6. The divisor Dαi(Z)D_{\alpha_{i}^{\vee}}(Z)

In this subsection, we complete the proof of Theorem 1.0.3 by proving Proposition 3.6.1 below.

For the statement, we let

N={n1+1,in typeA,n11,in typeF,n1,otherwise.N=\begin{cases}n_{1}+1,&\text{in type}\;\;A,\\ n_{1}-1,&\text{in type}\;\;F,\\ n_{1},&\text{otherwise}.\end{cases}

We let θN:YY×SPicS0(E)\theta_{N}^{\prime}\colon Y\to Y\times_{S}\mathrm{Pic}^{0}_{S}(E) be the section θN(y)=(y,0)\theta_{N}^{\prime}(y)=(y,0), and for 1k<N1\leq k<N, we let θk:YY×SPicS0(E)\theta_{k}^{\prime}\colon Y\to Y\times_{S}\mathrm{Pic}^{0}_{S}(E) be the section given in type AA by

θk(y)={(y,ϖi(y)+ϖi+1(y)+ϖl(y)),ifk=1,(y,ϖi(y)+ϖi+1(y)+ϖlk+1(y)ϖlk+2(y)),if  1<kli+1=N1,\theta_{k}^{\prime}(y)=\begin{cases}(y,-\varpi_{i}(y)+\varpi_{i+1}(y)+\varpi_{l}(y)),&\text{if}\;\;k=1,\\ (y,-\varpi_{i}(y)+\varpi_{i+1}(y)+\varpi_{l-k+1}(y)-\varpi_{l-k+2}(y)),&\text{if}\;\;1<k\leq l-i+1=N-1,\end{cases}

and in types BB, DD and EE by

θk(y)={(y,αl1(y)),in typeB,(y,αl2(y)++αlk(y)),in typeD,(y,αk(y)+αk+1(y)++α3(y)),in typeE.\theta_{k}^{\prime}(y)=\begin{cases}(y,\alpha_{l-1}(y)),&\text{in type}\;B,\\ (y,\alpha_{l-2}(y)+\cdots+\alpha_{l-k}(y)),&\text{in type}\;D,\\ (y,\alpha_{k}(y)+\alpha_{k+1}(y)+\cdots+\alpha_{3}(y)),&\text{in type}\;E.\end{cases}

Note that N=1N=1 in types CC, FF and GG.

Proposition 3.6.1.

Assume we are in the setup of Proposition 3.1.1, and moreover assume for simplicity of notation that i=l3i=l-3 if (G,P,μ)(G,P,\mu) is of type DD, and that i=5i=5 if (G,P,μ)(G,P,\mu) is of type EE. Then there is a sequence of NN morphisms

Dαi(Z)=DN+1DND1D_{\alpha_{i}^{\vee}}(Z)=D_{N+1}^{\prime}\longrightarrow D_{N}^{\prime}\longrightarrow\cdots\longrightarrow D_{1}^{\prime}

over Y×SZY\times_{S}Z such that D1D_{1}^{\prime} is a family of smooth surfaces over YY containing Y×SPicS0(E)Y\times_{S}\mathrm{Pic}^{0}_{S}(E) as a closed substack, and Dk+1DkD_{k+1}^{\prime}\to D_{k}^{\prime} is the blowup along the section θk:YY×SPicS0(E)Dk\theta_{k}^{\prime}\colon Y\to Y\times_{S}\mathrm{Pic}^{0}_{S}(E)\subseteq D_{k}^{\prime} of the proper transform of Y×SPicS0(E)D1Y\times_{S}\mathrm{Pic}^{0}_{S}(E)\subseteq D_{1}^{\prime}. Moreover, we have the following descriptions of D1D_{1}^{\prime} in each type.

  1. (1)

    In type AA, D1Y×SZ0=Y×SPic0(E)D_{1}^{\prime}\to Y\times_{S}Z_{0}=Y\times_{S}\mathrm{Pic}^{0}(E) is a line bundle.

  2. (2)

    In type BB, the morphism D1Y×SZ0D_{1}^{\prime}\to Y\times_{S}Z_{0} is a 1\mathbb{P}^{1}-bundle such that the fibre of D1YD_{1}^{\prime}\to Y over a point yYy\in Y is isomorphic to the stacky Hirzebruch surface

    (D1)y{(1,2)(𝒪𝒪(1)),ifϖl(y)0,(1,2)(𝒪𝒪(3)),ifϖl(y)=0.(D_{1}^{\prime})_{y}\cong\begin{cases}\mathbb{P}_{\mathbb{P}(1,2)}(\mathcal{O}\oplus\mathcal{O}(1)),&\text{if}\;\;\varpi_{l}(y)\neq 0,\\ \mathbb{P}_{\mathbb{P}(1,2)}(\mathcal{O}\oplus\mathcal{O}(3)),&\text{if}\;\;\varpi_{l}(y)=0.\end{cases}
  3. (3)

    In types CC and DD, the morphism D1Y×SZ0D_{1}^{\prime}\to Y\times_{S}Z_{0} is a 1\mathbb{P}^{1}-bundle such that the fibre of D1YD_{1}^{\prime}\to Y over a point yYy\in Y is isomorphic to the Hirzebruch surface

    (D1)y{𝔽0,ifϖl(y)0,𝔽2,ifϖl(y)=0.(D_{1}^{\prime})_{y}\cong\begin{cases}\mathbb{F}_{0},&\text{if}\;\;\varpi_{l}(y)\neq 0,\\ \mathbb{F}_{2},&\text{if}\;\;\varpi_{l}(y)=0.\end{cases}
  4. (4)

    In types EE and GG, the morphism D1Y×SZ0=YD_{1}^{\prime}\to Y\times_{S}Z_{0}=Y is a 2\mathbb{P}^{2}-bundle.

  5. (5)

    In type FF, the morphism D1Y×SZ0=YD_{1}^{\prime}\to Y\times_{S}Z_{0}=Y factors as a sequence of two 1\mathbb{P}^{1}-bundles D1D1′′YD_{1}^{\prime}\to D_{1}^{\prime\prime}\to Y, and the fibre over a point yYy\in Y is isomorphic to the Hirzebruch surface

    (D1)y{𝔽0,ifα1(y)0,𝔽2,ifα1(y)=0.(D_{1}^{\prime})_{y}\cong\begin{cases}\mathbb{F}_{0},&\text{if}\;\;\alpha_{1}(y)\neq 0,\\ \mathbb{F}_{2},&\text{if}\;\;\alpha_{1}(y)=0.\end{cases}
Proof.

First note that in type AA, the roots αi\alpha_{i} and αj=αi+1\alpha_{j}=\alpha_{i+1} play completely symmetric roles. So applying Proposition 3.4.1 with the vertices of the Dynkin diagram AlA_{l} labelled in reverse order gives contractions

Dαi(Z)=Dli+2Dli+1D1D_{\alpha_{i}^{\vee}}(Z)=D_{l-i+2}^{\prime}\longrightarrow D_{l-i+1}^{\prime}\longrightarrow\cdots\longrightarrow D_{1}^{\prime}

with the desired properties, where to get the correct blowup loci we have composed the identification of D1D_{1}^{\prime} with a line bundle over Y×SPicS0(E)Y\times_{S}\mathrm{Pic}^{0}_{S}(E) given by Proposition 3.4.1 with the isomorphism

Y×SPicS0(E)\displaystyle Y\times_{S}\mathrm{Pic}^{0}_{S}(E) Y×SPicS0(E)\displaystyle\overset{\sim}{\longrightarrow}Y\times_{S}\mathrm{Pic}^{0}_{S}(E)
(y,x)\displaystyle(y,x) (y,x).\displaystyle\longmapsto(y,-x).

If GG is not of type AA, then we define

Dαi(Z)DN:=KMB,G,rigαi×BunG,rigZ×SED_{\alpha_{i}^{\vee}}(Z)\longrightarrow D_{N}^{\prime}:=\mathrm{KM}_{B,G,rig}^{-\alpha_{i}^{\vee}}\times_{\mathrm{Bun}_{G,rig}}Z\times_{S}E (3.6.1)

to be the map given by deleting the unique degree αi\alpha_{i}^{\vee} rational component of a stable section and recording its image in EE.

Let

(DN)0=BunB,rigαi×BunG,rigZ×SEDN(D_{N}^{\prime})_{0}=\mathrm{Bun}_{B,rig}^{-\alpha_{i}^{\vee}}\times_{\mathrm{Bun}_{G,rig}}Z\times_{S}E\subseteq D_{N}^{\prime}

and let (DN)1=DN(DN)0(D_{N}^{\prime})_{1}=D_{N}^{\prime}\setminus(D_{N}^{\prime})_{0}. Then (DN)1(D_{N}^{\prime})_{1} is a smooth divisor in DND_{N}^{\prime} isomorphic to

(DN)1BunB,rigαiαj×BunG,rigZ×SE×SE,(D_{N}^{\prime})_{1}\cong\mathrm{Bun}_{B,rig}^{-\alpha_{i}^{\vee}-\alpha_{j}^{\vee}}\times_{\mathrm{Bun}_{G,rig}}Z\times_{S}E\times_{S}E,

where the first (resp., second) factor of EE above keeps track of the point of attachment of an αj\alpha_{j}^{\vee} curve. There is a morphism

(DN)1Y×SPicS0(E)(D_{N}^{\prime})_{1}\longrightarrow Y\times_{S}\mathrm{Pic}^{0}_{S}(E) (3.6.2)

given on the first factor by the morphism (DN)1DNY(D_{N}^{\prime})_{1}\to D_{N}^{\prime}\to Y and on the second by the morphism

(DN)1E×SE\displaystyle(D_{N}^{\prime})_{1}\longrightarrow E\times_{S}E PicS0(E)\displaystyle\longrightarrow\mathrm{Pic}^{0}_{S}(E)
(xj,xi)\displaystyle(x_{j},x_{i}) xjxi.\displaystyle\longmapsto x_{j}-x_{i}.

Using the fact that Dαi(Z)D_{\alpha_{i}^{\vee}}(Z) is naturally identified with the pullback of the universal domain curve over KMB,G,rigαi×BunG,rigZ\mathrm{KM}_{B,G,rig}^{-\alpha_{i}^{\vee}}\times_{\mathrm{Bun}_{G,rig}}Z, one can deduce from [davis19, Proposition 2.1.7] that (3.6.1) is the blow up at the preimage of the section θN:YY×SPicS0(E)\theta_{N}\colon Y\to Y\times_{S}\mathrm{Pic}^{0}_{S}(E) under (3.6.2). It follows that the strict transform Dαi(Z)Dαj(Z)D_{\alpha_{i}^{\vee}}(Z)\cap D_{\alpha_{j}^{\vee}}(Z) of (DN)1(D_{N}^{\prime})_{1} maps isomorphically to it. By construction, the composition

Dαi(Z)Dαj(Z)(DN)1(3.6.2)Y×SPicS0(E)D_{\alpha_{i}^{\vee}}(Z)\cap D_{\alpha_{j}^{\vee}}(Z)\overset{\sim}{\longrightarrow}(D_{N}^{\prime})_{1}\xrightarrow{\eqref{eq:subregularresolutions4:2}}Y\times_{S}\mathrm{Pic}^{0}_{S}(E)

agrees with the composition

Dαi(Z)Dαj(Z)C1,n0+1(3.4.6)Y×SPicS0(E),D_{\alpha_{i}^{\vee}}(Z)\cap D_{\alpha_{j}^{\vee}}(Z)\overset{\sim}{\longrightarrow}C_{1,n_{0}+1}\xrightarrow{\eqref{eq:bruhatcellcurvecomparison1}}Y\times_{S}\mathrm{Pic}^{0}_{S}(E),

and is therefore an isomorphism by Lemma 3.4.8.

The next step is to construct the spaces DkD_{k}^{\prime} for 1k<N1\leq k<N. This is vacuous for types CC, FF and GG (since N=1N=1 in these cases). In the remaining types, we define standard parabolics PkGP_{k}^{\prime}\subseteq G for 1k<N1\leq k<N and set

Dk=Y×YPk(KMPk,G,rig×BunG,rigZ×SE).D_{k}^{\prime}=Y\times_{Y_{P_{k}^{\prime}}}(\mathrm{KM}_{P_{k}^{\prime},G,rig}\times_{\mathrm{Bun}_{G,rig}}Z\times_{S}E).

In type BB, N=2N=2, and we let P1P_{1}^{\prime} be the standard parabolic with type t(P1)={αi,αl}={αl2,αl}t(P_{1}^{\prime})=\{\alpha_{i},\alpha_{l}\}=\{\alpha_{l-2},\alpha_{l}\}. In type DD, N=3N=3, and we let t(P1)={αi,αl}={αl3,αl}t(P_{1}^{\prime})=\{\alpha_{i},\alpha_{l}\}=\{\alpha_{l-3},\alpha_{l}\} and t(P2)={αl3,αl1,αl}t(P_{2}^{\prime})=\{\alpha_{l-3},\alpha_{l-1},\alpha_{l}\}. Finally, in type EE, N=4N=4, and we let t(P1)={αi,α4}={α4,α5}t(P_{1}^{\prime})=\{\alpha_{i},\alpha_{4}\}=\{\alpha_{4},\alpha_{5}\}, t(P2)={α1,α4,α5}t(P_{2}^{\prime})=\{\alpha_{1},\alpha_{4},\alpha_{5}\} and t(P3)={α1,α2,α4,α5}t(P_{3}^{\prime})=\{\alpha_{1},\alpha_{2},\alpha_{4},\alpha_{5}\}. Note that in each case, we have a sequence of morphisms

DNDN1D1D_{N}^{\prime}\longrightarrow D_{N-1}^{\prime}\longrightarrow\cdots\longrightarrow D_{1}^{\prime}

coming from the inclusions of the parabolics.

We prove below in Proposition 3.6.6 that the spaces D1D_{1}^{\prime} are as described in the statement of the proposition. This completes the proof of the proposition in types CC, FF and GG. In types BB, DD and EE, we still need to show that Dk+1DkD_{k+1}^{\prime}\to D_{k}^{\prime} is the blowup at the desired section for 1k<N1\leq k<N. As in the proof of Proposition 3.4.1, the proof relies on a decomposition into locally closed substacks coming from the Bruhat cells of §3.3.

We define representations πP1:P1GLn1\pi_{P_{1}^{\prime}}\colon P_{1}^{\prime}\to GL_{n_{1}} of P1P_{1}^{\prime} as follows. In type BB, we let πP1\pi_{P_{1}^{\prime}} be given by

P1P1/Ru(P)=LP1ρLGSp4R4,P_{1}^{\prime}\longrightarrow P_{1}^{\prime}/R_{u}(P)=L\cap P_{1}^{\prime}\overset{\rho_{L}}{\longrightarrow}GSp_{4}\cap R_{4},

composed with the homomorphism

GSp4R4\displaystyle GSp_{4}\cap R_{4} GL2\displaystyle\longrightarrow GL_{2}
(λ1detA000000A0000λ)\displaystyle\left(\begin{array}[]{c|cc|c}\lambda^{-1}\det A&0&0&0\\ \hline\cr 0&&&0\\ 0&\lx@intercol\hfil\smash{\raisebox{6.0pt}{$A$}}\hfil\lx@intercol\vrule\lx@intercol&0\\ \hline\cr\\[-12.0pt] 0&0&0&\lambda\end{array}\right) A,\displaystyle\longmapsto A,

where ρL\rho_{L} is the representation defined in §2.4. In types DD and EE, we let πP1:P1GLn1\pi_{P_{1}^{\prime}}\colon P_{1}^{\prime}\to GL_{n_{1}} be the composition

πP1:P1LP1ρLRn1+1GLn1,\pi_{P_{1}^{\prime}}\colon P_{1}^{\prime}\longrightarrow L\cap P_{1}^{\prime}\overset{\rho_{L}}{\longrightarrow}R_{n_{1}+1}\longrightarrow GL_{n_{1}},

where the last homomorphism is given by deleting the last row and column, and ρL\rho_{L} is the composition of the isomorphism of Lemma 2.3.4 with the projection to the second factor.

In each of types BB, DD and EE, we have Pk=(πP1)1(Qkn1)P_{k}^{\prime}=(\pi_{P_{1}^{\prime}})^{-1}(Q_{k}^{n_{1}}) for 1kN1\leq k\leq N, where we set PN=PP1P_{N}^{\prime}=P\cap P_{1} in the notation of §3.4. Note that the morphism

DNY×YPN(KMPN,G,rigαi×BunG,rigZ×SE)D_{N}^{\prime}\longrightarrow Y\times_{Y_{P_{N}^{\prime}}}(\mathrm{KM}_{P_{N}^{\prime},G,rig}^{-\alpha_{i}^{\vee}}\times_{\mathrm{Bun}_{G,rig}}Z\times_{S}E) (3.6.3)

is an isomorphism by [davis19, Lemma 4.3.7], since there is an isomorphism PN/BGLn0/Qn0n0P_{N}^{\prime}/B\cong GL_{n_{0}}/Q^{n_{0}}_{n_{0}} identifying sections of degree αi-\alpha_{i}^{\vee} with sections of degree en0-e_{n_{0}}^{*}. So we have a sequence of pullback squares

Dk+1{D_{k+1}^{\prime}}YQn1n1en1×YQk+1n1en1KMQk+1n1,GLn1,rigen1{Y_{Q^{n_{1}}_{n_{1}}}^{-e_{n_{1}}^{*}}\times_{Y_{Q_{k+1}^{n_{1}}}^{-e_{n_{1}}^{*}}}\mathrm{KM}_{Q^{n_{1}}_{k+1},GL_{n_{1}},rig}^{-e_{n_{1}}^{*}}}Dk{D_{k}^{\prime}}YQn1n1en1×YQkn1en1KMQkn1,GLn1,rigen1,{Y_{Q^{n_{1}}_{n_{1}}}^{-e_{n_{1}}^{*}}\times_{Y_{Q_{k}^{n_{1}}}^{-e_{n_{1}}^{*}}}\mathrm{KM}_{Q^{n_{1}}_{k},GL_{n_{1}},rig}^{-e_{n_{1}}^{*}},} (3.6.4)

where the subscript ()rig(-)_{rig} denotes the rigidification with respect to the image of Z(G)Z(G) in Z(GLn1)Z(GL_{n_{1}}).

By Lemma 3.6.2, there is a stable section of ξG,z×GG/P\xi_{G,z}\times^{G}G/P of degree αi-\alpha_{i}^{\vee} if and only if zZ0Zz\in Z_{0}\subseteq Z, and for such zz, the unique such section is the canonical (Harder-Narasimhan) one of ξG,z×GG/P=ξL,z×LG/P\xi_{G,z}\times^{G}G/P=\xi_{L,z}\times^{L}G/P. Since PkPP_{k}^{\prime}\subseteq P, one can use this fact, the definition of the slice Z0Z_{0} and elementary slope arguments (see e.g., [davis19a, Lemma 6.6.11]) to show in each case that any unstable GLn1GL_{n_{1}}-bundle in the image of D1BunGLn1,rig1D_{1}^{\prime}\to\mathrm{Bun}_{GL_{n_{1}},rig}^{-1} has Harder-Narasimhan reduction to Rn1R_{n_{1}} of degree e1-e_{1}^{*}. By Proposition 3.3.1, we therefore have a decomposition

Dk=(Dk×BunGLn1,rig1BunGLn1,rigss,1)1p<kCk,pCk,n1D_{k}^{\prime}=(D_{k}^{\prime}\times_{\mathrm{Bun}_{GL_{n_{1}},rig}^{-1}}\mathrm{Bun}_{GL_{n_{1}},rig}^{ss,-1})\cup\bigcup_{1\leq p<k}C_{k,p}^{\prime}\cup C_{k,n_{1}}^{\prime}

into disjoint locally closed substacks for 1kn1=N1\leq k\leq n_{1}=N, where Ck,pDkC_{k,p}^{\prime}\subseteq D_{k}^{\prime} is the preimage of Ck,p,rigGLn1Xk,rign1C_{k,p,rig}^{GL_{n_{1}}}\subseteq X^{n_{1}}_{k,rig} in DkD_{k}^{\prime}. We remark that CN,n1=(DN)1Y×SPicS0(E)C_{N,n_{1}}^{\prime}=(D^{\prime}_{N})_{1}\cong Y\times_{S}\mathrm{Pic}^{0}_{S}(E).

Using Proposition 3.3.6, Lemma 3.4.10, [davis19, Lemma 4.3.7] and the pullback squares (3.6.4), one can now check that Dk+1DkD_{k+1}^{\prime}\to D_{k}^{\prime} is the blowup along the desired section θk\theta_{k}^{\prime} of Ck,n1CN,n1=Y×SPicS0(E)C_{k,n_{1}}^{\prime}\cong C_{N,n_{1}}^{\prime}=Y\times_{S}\mathrm{Pic}^{0}_{S}(E) exactly as in the proof of Proposition 3.4.1. ∎

In the rest of this subsection, we will establish the propositions and lemmas quoted in the proof of Proposition 3.6.1. We will assume from now on that (G,P,μ)(G,P,\mu) is not of type AA.

Lemma 3.6.2.

Assume zZz\in Z is such that there exists a section of ξG,z×GG/P\xi_{G,z}\times^{G}G/P of degree αi\leq-\alpha_{i}^{\vee}. Then zZ0Zz\in Z_{0}\subseteq Z, and the only such section is the canonical (Harder-Narasimhan) one of ξG,z×GG/P=ξP,z×PG/P\xi_{G,z}\times^{G}G/P=\xi_{P,z}\times^{P}G/P.

Proof.

First note that KMB,GαiKMP,Gαi\mathrm{KM}_{B,G}^{-\alpha_{i}^{\vee}}\to\mathrm{KM}_{P,G}^{-\alpha_{i}^{\vee}} is surjective by [davis19a, Proposition 3.6.4]. Since zZ0z\notin Z_{0} implies that ξG,z\xi_{G,z} is either semistable or regular unstable, we must therefore have zZ0z\in Z_{0} by [davis19, Lemma 4.3.4] as αi\alpha_{i} is not a special root. Given a section σ\sigma of ξG,z×GG/P\xi_{G,z}\times^{G}G/P of degree αi\leq-\alpha_{i}^{\vee}, any lift to a section of ξG,z×GG/B\xi_{G,z}\times^{G}G/B of degree αi\leq-\alpha_{i}^{\vee} must factor through ξP,z×PP/B\xi_{P,z}\times^{P}P/B by Lemma 3.6.3 and Proposition 3.2.1, so σ\sigma must be the canonical section as claimed. ∎

Lemma 3.6.3.

Assume wWP,B0w\in W^{0}_{P,B}, λαi\lambda\leq-\alpha_{i}^{\vee} and Cw,λ(Z0)C^{w,\lambda}(Z_{0})\neq\emptyset. Then w=1w=1 and λ{αi,αiαj}\lambda\in\{-\alpha_{i}^{\vee},-\alpha_{i}^{\vee}-\alpha_{j}^{\vee}\}.

Proof.

From the proof of Lemma 3.4.3, we have either wλ=αiw\lambda=-\alpha_{i}^{\vee} and w=1w=1, or wλ=αiαjw\lambda=-\alpha_{i}^{\vee}-\alpha_{j}^{\vee} and

w{1}{sc0,n0sc0,n01sc0,k1kn0}.w\in\{1\}\cup\{s_{c_{0},n_{0}}s_{c_{0},n_{0}-1}\cdots s_{c_{0},k}\mid 1\leq k\leq n_{0}\}.

If w1w\neq 1, then this implies that λ=w1(αi+αj)=αj\lambda=-w^{-1}(\alpha_{i}^{\vee}+\alpha_{j}^{\vee})=-\alpha_{j}^{\vee}, contradicting λαi\lambda\leq-\alpha_{i}^{\vee}. So this proves the lemma. ∎

It now remains only to describe the maps D1Y×SZ0D_{1}^{\prime}\to Y\times_{S}Z_{0}. We do this in Proposition 3.6.6 after a few preparations.

Since the statement is local on SS, we will assume from now on that the initial section SBunL,rigss,μS\to\mathrm{Bun}_{L,rig}^{ss,\mu} (resp. 𝔹S𝔾mBunL,rigss,μ\mathbb{B}_{S}\mathbb{G}_{m}\to\mathrm{Bun}_{L^{\prime},rig}^{ss,\mu^{\prime}}) used in the construction of the slice Z0Z_{0} in types EE, FF and GG (resp. BB, CC and DD) lifts to a section SBunLss,μS\to\mathrm{Bun}_{L}^{ss,\mu} (resp. SBunLss,μS\to\mathrm{Bun}_{L^{\prime}}^{ss,\mu^{\prime}}). We will also write Z1=Z0=SZ_{1}=Z_{0}=S in types EE, FF and GG and Z1=IndLL(S)SZ_{1}=\mathrm{Ind}_{L^{\prime}}^{L}(S)\setminus S in types BB, CC and DD; our assumption implies that Z0BunL,rigss,μZ_{0}\to\mathrm{Bun}_{L,rig}^{ss,\mu} lifts to Z1BunLss,μZ_{1}\to\mathrm{Bun}_{L}^{ss,\mu}.

We first relate D1Y×SZ0D_{1}^{\prime}\to Y\times_{S}Z_{0} to the projectivisation of a vector bundle. Let ρL\rho_{L} be the representation of LL given by the isomorphism of Lemmas 2.3.4 and 2.3.5 composed with the projection to the second factor in types CC, DD, EE, FF and GG, and given by the isomorphism of Lemma 2.3.6 composed with the projection to the second factor and the inclusion GSp4GL4GSp_{4}\subseteq GL_{4} in type BB. We will write WW for the vector bundle on Z1×SEZ_{1}\times_{S}E induced by Z1BunLss,μZ_{1}\to\mathrm{Bun}_{L}^{ss,\mu} and ρL\rho_{L}. We will also write λ𝕏(T)\lambda\in\mathbb{X}^{*}(T) for the character

λ={ϖl,in typesB,C,D,ϖ4,in typeE,ϖ2,in typeG.\lambda=\begin{cases}\varpi_{l},&\text{in types}\;B,C,D,\\ \varpi_{4},&\text{in type}\;E,\\ \varpi_{2},&\text{in type}\;G.\end{cases}
Lemma 3.6.4.

In types BB, CC, DD, EE and GG, there is an isomorphism

D1×Z0Z1Y×SZ1π(Mλ𝒪(dOE)W),D_{1}^{\prime}\times_{Z_{0}}Z_{1}\cong\mathbb{P}_{Y\times_{S}Z_{1}}\pi_{*}(M_{\lambda}\otimes\mathcal{O}(dO_{E})\otimes W),

where π:Y×SZ1×SEY×SZ1\pi\colon Y\times_{S}Z_{1}\times_{S}E\to Y\times_{S}Z_{1} is the natural projection and MλM_{\lambda} is the line bundle on Y×SZ1×SEY\times_{S}Z_{1}\times_{S}E classified by the morphism

Y×SZ1Y𝜆PicS0(E).Y\times_{S}Z_{1}\longrightarrow Y\overset{\lambda}{\longrightarrow}\mathrm{Pic}^{0}_{S}(E).
Proof.

We first prove the lemma in types BB, DD and EE. Let

X=Y×YP1(BunLP1αi×BunLZ1×SE)D1×Z0Z1,X=Y\times_{Y_{P_{1}^{\prime}}}(\mathrm{Bun}_{L\cap P_{1}^{\prime}}^{-\alpha_{i}^{\vee}}\times_{\mathrm{Bun}_{L}}Z_{1}\times_{S}E)\subseteq D_{1}^{\prime}\times_{Z_{0}}Z_{1},

where we note that Lemma 3.6.2 implies that

D1=Y×YP1(KMLP1,L,rigαi×BunL,rigZ0×SE).D_{1}^{\prime}=Y\times_{Y_{P_{1}^{\prime}}}(\mathrm{KM}_{L\cap P_{1}^{\prime},L,rig}^{-\alpha_{i}^{\vee}}\times_{\mathrm{Bun}_{L,rig}}Z_{0}\times_{S}E).

Lemmas 2.3.4 and 2.4.3 show that XX is the stack of tuples (y,z,Mλ,y1𝒪(OE)Wz)(y,z,M_{\lambda,y}^{-1}\otimes\mathcal{O}(-O_{E})\subseteq W_{z}), where yYy\in Y, zZ1z\in Z_{1}, Mλ,yM_{\lambda,y} is the line bundle on EE corresponding to λ(y)PicS0(E)\lambda(y)\in\mathrm{Pic}^{0}_{S}(E), and WzW_{z} is the restriction of WW to the fibre over zZ1z\in Z_{1}. Since the vector bundle WzW_{z} is semistable of slope <0<0, any nonzero morphism Mλ,y1𝒪(OE)WzM_{\lambda,y}^{-1}\otimes\mathcal{O}(-O_{E})\to W_{z} must be a subbundle, so we have an isomorphism

XY×SZ1π(Mλ𝒪(OE)W).X\cong\mathbb{P}_{Y\times_{S}Z_{1}}\pi_{*}(M_{\lambda}\otimes\mathcal{O}(O_{E})\otimes W).

Since this implies in particular that XX is already proper over Y×SZ1=Y×YP1(Z1×SE)Y\times_{S}Z_{1}=Y\times_{Y_{P_{1}^{\prime}}}(Z_{1}\times_{S}E), we conclude that X=D1×Z0Z1X=D_{1}^{\prime}\times_{Z_{0}}Z_{1} and the claim is proved.

In types CC and GG, we argue instead as follows. Observe that there is a pullback

D1×Z0Z1{D_{1}^{\prime}\times_{Z_{0}}Z_{1}}KMQ22,GL2de2×BunGL2BunGL2ss,d{\mathrm{KM}_{Q^{2}_{2},GL_{2}}^{-de_{2}^{*}}\times_{\mathrm{Bun}_{GL_{2}}}\mathrm{Bun}_{GL_{2}}^{ss,-d}}Y×SZ1{Y\times_{S}Z_{1}}PicSd(E)×SBunGL2ss,d,{\mathrm{Pic}^{-d}_{S}(E)\times_{S}\mathrm{Bun}_{GL_{2}}^{ss,-d},} (3.6.5)

where the bottom morphism is given by

(y,z)(Mλ,y1𝒪(dOE),Wz)(y,z)\longmapsto(M_{\lambda,y}^{-1}\otimes\mathcal{O}(-dO_{E}),W_{z})

and the right morphism is given on the first factor by the blow down to TQ22T_{Q^{2}_{2}}-bundles composed with the character e2e_{2}. If (y,z)Y×SZ1(y,z)\in Y\times_{S}Z_{1} lies over a geometric point s:SpeckSs\colon\operatorname{\mathrm{Spec}}k\to S, then any stable map to the GL2GL_{2} flag variety bundle (Wz)\mathbb{P}(W_{z}^{\vee}) corresponding to a point in D1×Z0Z1D_{1}^{\prime}\times_{Z_{0}}Z_{1} over (y,z)(y,z) is a closed immersion with ideal sheaf p(Mλ,y1𝒪(d)OE))𝒪(1)p^{*}(M_{\lambda,y}^{-1}\otimes\mathcal{O}(-d)O_{E}))\otimes\mathcal{O}(-1), where p:(Wz)Esp\colon\mathbb{P}(W_{z}^{\vee})\to E_{s} is the structure morphism. So we deduce that

D1×Z0Z1=Y×SZ1πp(p(Mλ𝒪(dOE))𝒪(1))=Y×SZ1π(Mλ𝒪(dOE)Wz)D_{1}^{\prime}\times_{Z_{0}}Z_{1}=\mathbb{P}_{Y\times_{S}Z_{1}}\pi_{*}p_{*}(p^{*}(M_{\lambda}\otimes\mathcal{O}(dO_{E}))\otimes\mathcal{O}(1))=\mathbb{P}_{Y\times_{S}Z_{1}}\pi_{*}(M_{\lambda}\otimes\mathcal{O}(dO_{E})\otimes W_{z})

as claimed. ∎

The situation in type FF is similar. In this case, we let P1′′LP_{1}^{\prime\prime}\subseteq L be the standard parabolic subgroup of type t(P1′′)={α1}t(P_{1}^{\prime\prime})=\{\alpha_{1}\}, and define

D1′′=Y×YP1′′(KMP1′′,L,rigαi×BunL,rigμZ0×SE).D_{1}^{\prime\prime}=Y\times_{Y_{P_{1}^{\prime\prime}}}(\mathrm{KM}_{P_{1}^{\prime\prime},L,rig}^{-\alpha_{i}^{\vee}}\times_{\mathrm{Bun}_{L,rig}^{\mu}}Z_{0}\times_{S}E).
Lemma 3.6.5.

In type FF, there are isomorphisms

D1′′Y×SZ1π(Mϖ1W)D_{1}^{\prime\prime}\cong\mathbb{P}_{Y\times_{S}Z_{1}}\pi_{*}(M_{\varpi_{1}}\otimes W^{\vee})

and

D1D1′′π(pMϖ2𝒪(2OE)ker(pWpMϖ1𝒪D1′′(1))),D_{1}^{\prime}\cong\mathbb{P}_{D_{1}^{\prime\prime}}\pi^{\prime}_{*}(p^{*}M_{\varpi_{2}}\otimes\mathcal{O}(2O_{E})\otimes\ker(p^{*}W\to p^{*}M_{\varpi_{1}}\otimes\mathcal{O}_{D_{1}^{\prime\prime}}(1))),

where π:Y×SZ1×SEY×SZ1\pi\colon Y\times_{S}Z_{1}\times_{S}E\to Y\times_{S}Z_{1} and π:D1×SZ1×SEY×SZ1\pi^{\prime}\colon D_{1}^{\prime}\times_{S}Z_{1}\times_{S}E\to Y\times_{S}Z_{1} are the natural projections, and p:D1′′Y×SZ1p\colon D_{1}^{\prime\prime}\to Y\times_{S}Z_{1} is the structure morphism.

Proof.

Recall that αi=α3\alpha_{i}=\alpha_{3} and Z1=SZ_{1}=S in this case and let

X=Y×YP1′′(BunP1′′α3×BunLμZ1×SE)D1′′.X=Y\times_{Y_{P_{1}^{\prime\prime}}}(\mathrm{Bun}_{P_{1}^{\prime\prime}}^{-\alpha_{3}^{\vee}}\times_{\mathrm{Bun}_{L}^{\mu}}Z_{1}\times_{S}E)\subseteq D_{1}^{\prime\prime}.

Then Lemma 2.3.4 shows that XX is the stack of tuples (y,z,WzMϖ1,y)(y,z,W_{z}\twoheadrightarrow M_{\varpi_{1},y}), where yYy\in Y and zZ1z\in Z_{1}. Since the vector bundle WzW_{z} is semistable of slope >1>-1, any nonzero morphism WzMϖ1,yW_{z}\to M_{\varpi_{1},y} is surjective, so we have an isomorphism

XY×SZ1(π(Mϖ1W)).X\cong\mathbb{P}_{Y\times_{S}Z_{1}}(\pi_{*}(M_{\varpi_{1}}\otimes W^{\vee})).

Since this shows that XX is already proper over Y×SZ1=Y×YP1′′(Z1×SE)Y\times_{S}Z_{1}=Y\times_{Y_{P_{1}^{\prime\prime}}}(Z_{1}\times_{S}E), it follows that X=D1′′X=D_{1}^{\prime\prime}, so this gives the first of the desired isomorphisms.

For the second isomorphism, there is a pullback

D1{D_{1}^{\prime}}KMQ22,GL22e2×BunGL22BunGL2ss,2{\mathrm{KM}_{Q^{2}_{2},GL_{2}}^{-2e_{2}^{*}}\times_{\mathrm{Bun}_{GL_{2}}^{-2}}\mathrm{Bun}_{GL_{2}}^{ss,-2}}D1′′{D_{1}^{\prime\prime}}PicS2(E)×SBunGL2ss,2{\mathrm{Pic}^{-2}_{S}(E)\times_{S}\mathrm{Bun}_{GL_{2}}^{ss,-2}}

where the bottom horizontal morphism is classified by the pair (pMϖ21𝒪(2OE),ker(pWpMϖ1𝒪D1′′(1)))(p^{*}M_{\varpi_{2}}^{-1}\otimes\mathcal{O}(-2O_{E}),\ker(p^{*}W\to p^{*}M_{\varpi_{1}}\otimes\mathcal{O}_{D_{1}^{\prime\prime}}(1))) of line bundle and vector bundle on D1′′×SED_{1}^{\prime\prime}\times_{S}E. Since any stable map to the associated flag variety bundle appearing in D1D_{1}^{\prime} is again a closed immersion, the argument used in the proof of Lemma 3.6.4 for types CC and GG gives the desired isomorphism

D1D1′′π(pMϖ2𝒪(2OE)ker(pWpMϖ1𝒪D1′′(1))).D_{1}^{\prime}\cong\mathbb{P}_{D_{1}^{\prime\prime}}\pi^{\prime}_{*}(p^{*}M_{\varpi_{2}}\otimes\mathcal{O}(2O_{E})\otimes\ker(p^{*}W\to p^{*}M_{\varpi_{1}}\otimes\mathcal{O}_{D_{1}^{\prime\prime}}(1))).

Proposition 3.6.6.

The descriptions given in Proposition 3.6.1 for the maps D1Y×SZ0D_{1}^{\prime}\to Y\times_{S}Z_{0} are correct.

Proof.

First observe that in types EE and GG, Mλ𝒪(dOE)WM_{\lambda}\otimes\mathcal{O}(dO_{E})\otimes W is a family of semistable vector bundles of degree 33, so Lemma 3.6.4 shows that D1Y×SZ1=YD_{1}^{\prime}\to Y\times_{S}Z_{1}=Y is a 2\mathbb{P}^{2}-bundle, which proves (4).

In types BB, CC and DD, Mλ𝒪((d+1)OE)WM_{\lambda}\otimes\mathcal{O}((d+1)O_{E})\otimes W is a family of semistable vector bundles of degree 22, so Lemma 3.6.4 shows that D1×Z0Z1Y×SZ1D_{1}^{\prime}\times_{Z_{0}}Z_{1}\to Y\times_{S}Z_{1} is a 1\mathbb{P}^{1}-bundle, and hence that D1Y×SZ0D_{1}^{\prime}\to Y\times_{S}Z_{0} is also.

To complete the proof of (2), note that in type BB, we have a canonical Z(L)Z(L^{\prime})-invariant subbundle 𝒪(OE)W\mathcal{O}(-O_{E})\subseteq W and a Z(L)Z(L^{\prime})-equivariant exact sequence

0UW/𝒪(OE)𝒪0,0\longrightarrow U\longrightarrow W/\mathcal{O}(-O_{E})\longrightarrow\mathcal{O}\longrightarrow 0,

where UU is a family of stable vector bundles on EE of rank 22 and determinant 𝒪(OE)\mathcal{O}(-O_{E}). So if we fix a geometric point y:SpeckYy\colon\operatorname{\mathrm{Spec}}k\to Y over s:SpeckSs\colon\operatorname{\mathrm{Spec}}k\to S, we have Z(L)Z(L^{\prime})-equivariant exact sequences

0π(Mϖl,y)π\displaystyle 0\longrightarrow\pi_{*}(M_{\varpi_{l},y})\longrightarrow\pi_{*} (Mϖl,y𝒪(OE)Ws)\displaystyle(M_{\varpi_{l},y}\otimes\mathcal{O}(O_{E})\otimes W_{s}) (3.6.6)
π(Mϖl,y𝒪(OE)(Ws/𝒪(OE)))1π(Mϖl,y)0,\displaystyle\longrightarrow\pi_{*}(M_{\varpi_{l},y}\otimes\mathcal{O}(O_{E})\otimes(W_{s}/\mathcal{O}(-O_{E})))\longrightarrow\mathbb{R}^{1}\pi_{*}(M_{\varpi_{l},y})\longrightarrow 0,

and

0π(Mϖl,y𝒪(OE)Us)π(Mϖl,y𝒪(OE)\displaystyle 0\longrightarrow\pi_{*}(M_{\varpi_{l},y}\otimes\mathcal{O}(O_{E})\otimes U_{s})\longrightarrow\pi_{*}(M_{\varpi_{l},y}\otimes\mathcal{O}(O_{E}) (Ws/𝒪(OE)))\displaystyle\otimes(W_{s}/\mathcal{O}(-O_{E}))) (3.6.7)
π(Mϖl,y𝒪(OE))0\displaystyle\longrightarrow\pi_{*}(M_{\varpi_{l},y}\otimes\mathcal{O}(O_{E}))\longrightarrow 0

of Z(L)Z(L^{\prime})-linearised vector bundles on (Z1)s(Z_{1})_{s}. Note that π(Mϖl,y)\pi_{*}(M_{\varpi_{l},y}), 1π(Mϖl,y)\mathbb{R}^{1}\pi_{*}(M_{\varpi_{l},y}), π(Mϖl,y𝒪(OE)Us)\pi_{*}(M_{\varpi_{l},y}\otimes\mathcal{O}(O_{E})\otimes U_{s}) and π(Mϖl,y𝒪(OE))\pi_{*}(M_{\varpi_{l},y}\otimes\mathcal{O}(O_{E})) are each either a trivial line bundle or zero, with Z(L)Z(L^{\prime})-weights f4f_{4}, f4f_{4}, f2=f3f_{2}=f_{3} and f1f_{1} respectively, where we use the notation of the proof of Lemma 2.3.6. So after tensoring with the character f1-f_{1} of Z(L)Z(L^{\prime}), Z(G)Z(G) acts trivially on (3.6.6) and (3.6.7), so they descend to exact sequences of vector bundles on (Z0)s=(Z1)s/𝔾m(1,2)(Z_{0})_{s}=(Z_{1})_{s}/\mathbb{G}_{m}\cong\mathbb{P}(1,2). Examining the 𝔾m\mathbb{G}_{m}-weights, the sequence (3.6.7) descends to a sequence of the form

0𝒪(1)W𝒪0.0\longrightarrow\mathcal{O}(1)\longrightarrow W^{\prime}\longrightarrow\mathcal{O}\longrightarrow 0.

Since any such sequence splits, we must have W𝒪𝒪(1)W^{\prime}\cong\mathcal{O}\oplus\mathcal{O}(1) as vector bundles on (1,2)\mathbb{P}(1,2).

If ϖl(y)0\varpi_{l}(y)\neq 0, then π(Mϖl,y)=1π(Mϖl,y)=0\pi_{*}(M_{\varpi_{l},y})=\mathbb{R}^{1}\pi_{*}(M_{\varpi_{l},y})=0, so we have

π(Mϖl,y𝒪(OE)Ws)π(Mϖl,y𝒪(OE)Ws/𝒪(OE)),\pi_{*}(M_{\varpi_{l},y}\otimes\mathcal{O}(O_{E})\otimes W_{s})\cong\pi_{*}(M_{\varpi_{l},y}\otimes\mathcal{O}(O_{E})\otimes W_{s}/\mathcal{O}(-O_{E})),

and hence (D1)y=(1,2)(W)=(1,2)(𝒪𝒪(1))(D_{1}^{\prime})_{y}=\mathbb{P}_{\mathbb{P}(1,2)}(W^{\prime})=\mathbb{P}_{\mathbb{P}(1,2)}(\mathcal{O}\oplus\mathcal{O}(1)). Otherwise, (3.6.6) tensored with f1-f_{1} descends to an exact sequence

0𝒪(2)W′′W=𝒪𝒪(1)𝒪(2)00\longrightarrow\mathcal{O}(2)\longrightarrow W^{\prime\prime}\longrightarrow W^{\prime}=\mathcal{O}\oplus\mathcal{O}(1)\longrightarrow\mathcal{O}(2)\longrightarrow 0

such that (D1)y=(1,2)(W′′)(D_{1}^{\prime})_{y}=\mathbb{P}_{\mathbb{P}(1,2)}(W^{\prime\prime}). But since the kernel of any surjection 𝒪𝒪(1)𝒪(2)\mathcal{O}\oplus\mathcal{O}(1)\to\mathcal{O}(2) on (1,2)\mathbb{P}(1,2) must be isomorphic to 𝒪(1)\mathcal{O}(-1), this means that we must have W′′=𝒪(1)𝒪(2)W^{\prime\prime}=\mathcal{O}(-1)\oplus\mathcal{O}(2), so

(D1)y=(1,2)(𝒪(1)𝒪(2))=(1,2)(𝒪𝒪(3)).(D_{1}^{\prime})_{y}=\mathbb{P}_{\mathbb{P}(1,2)}(\mathcal{O}(-1)\oplus\mathcal{O}(2))=\mathbb{P}_{\mathbb{P}(1,2)}(\mathcal{O}\oplus\mathcal{O}(3)).

This proves (2).

Similarly, to prove (3), note that in types CC and DD we have a canonical Z(L)Z(L^{\prime})-equivariant exact sequence

0𝒪(dOE)WU0,0\longrightarrow\mathcal{O}(-dO_{E})\longrightarrow W\longrightarrow U\longrightarrow 0,

where UU is semistable and Z(L)Z(L^{\prime}) acts on 𝒪(dOE)\mathcal{O}(-dO_{E}) and 𝒪\mathcal{O} respectively with weights

en1+1=ϖl+(d+1)ϖi={ϖl+2ϖl1,in typeC,ϖl+ϖl3,in typeD,ande1={ϖl,in typeC,ϖl1,in typeD.e_{n_{1}+1}=-\varpi_{l}+(d+1)\varpi_{i}=\begin{cases}-\varpi_{l}+2\varpi_{l-1},&\text{in type}\;C,\\ -\varpi_{l}+\varpi_{l-3},&\text{in type}\;D,\end{cases}\quad\text{and}\quad e_{1}=\begin{cases}\varpi_{l},&\text{in type}\;C,\\ \varpi_{l-1},&\text{in type}\;D.\end{cases}

So over any geometric point y:SpeckYy\colon\operatorname{\mathrm{Spec}}k\to Y over s:SpeckSs\colon\operatorname{\mathrm{Spec}}k\to S, we have an exact sequence

0π(Mϖl,y)π\displaystyle 0\longrightarrow\pi_{*}(M_{\varpi_{l},y})\longrightarrow\pi_{*} (Mϖl,y𝒪(dOE)Ws)\displaystyle(M_{\varpi_{l},y}\otimes\mathcal{O}(dO_{E})\otimes W_{s}) (3.6.8)
π(Mϖl,y𝒪(dOE)Us)1π(Mϖl,y)0,\displaystyle\longrightarrow\pi_{*}(M_{\varpi_{l},y}\otimes\mathcal{O}(dO_{E})\otimes U_{s})\longrightarrow\mathbb{R}^{1}\pi_{*}(M_{\varpi_{l},y})\longrightarrow 0,

of Z(L)Z(L^{\prime})-linearised vector bundles on (Z1)s(Z_{1})_{s}, which descends to an exact sequence of vector bundles on 1=(Z0)s=(Z1)s/𝔾m\mathbb{P}^{1}=(Z_{0})_{s}=(Z_{1})_{s}/\mathbb{G}_{m} after tensoring with e1-e_{1}. Note that in both cases Mϖl,y𝒪((d+1)OE)UsM_{\varpi_{l},y}\otimes\mathcal{O}((d+1)O_{E})\otimes U_{s} is a semistable vector bundle of degree 22 on which Z(L)Z(L^{\prime}) acts with the single weight e1e_{1}, so π(Mϖl,y𝒪((d+1)OE)Us)e1\pi_{*}(M_{\varpi_{l},y}\otimes\mathcal{O}((d+1)O_{E})\otimes U_{s})\otimes\mathbb{Z}_{-e_{1}} descends to a trivial rank 22 vector bundle 𝒪𝒪\mathcal{O}\oplus\mathcal{O} on 1\mathbb{P}^{1}.

If ϖl(y)0\varpi_{l}(y)\neq 0, then π(Mϖl,y)=1π(Mϖl,y)=0\pi_{*}(M_{\varpi_{l},y})=\mathbb{R}^{1}\pi_{*}(M_{\varpi_{l},y})=0, so

π(Mϖl,y𝒪((d+1)OE)Ws)e1=π(Mϖl,y𝒪((d+1)OE)Us)e1\pi_{*}(M_{\varpi_{l},y}\otimes\mathcal{O}((d+1)O_{E})\otimes W_{s})\otimes\mathbb{Z}_{-e_{1}}=\pi_{*}(M_{\varpi_{l},y}\otimes\mathcal{O}((d+1)O_{E})\otimes U_{s})\otimes\mathbb{Z}_{-e_{1}}

descends to 𝒪𝒪\mathcal{O}\oplus\mathcal{O} on 1\mathbb{P}^{1}, which together with Lemma 3.6.4 shows that (D1)y=1(𝒪𝒪)=𝔽0(D_{1}^{\prime})_{y}=\mathbb{P}_{\mathbb{P}^{1}}(\mathcal{O}\oplus\mathcal{O})=\mathbb{F}_{0}. Otherwise, (3.6.8) descends to an exact sequence

0𝒪(1)W𝒪𝒪𝒪(1)00\longrightarrow\mathcal{O}(1)\longrightarrow W^{\prime}\longrightarrow\mathcal{O}\oplus\mathcal{O}\longrightarrow\mathcal{O}(1)\longrightarrow 0

such that (D1)y1(W)(D_{1}^{\prime})_{y}\cong\mathbb{P}_{\mathbb{P}^{1}}(W^{\prime}). Since the kernel of any surjection 𝒪𝒪𝒪(1)\mathcal{O}\oplus\mathcal{O}\to\mathcal{O}(1) must be isomorphic to 𝒪(1)\mathcal{O}(-1), this implies that W𝒪(1)𝒪(1)W^{\prime}\cong\mathcal{O}(-1)\oplus\mathcal{O}(1) and hence that

(D1)y1(𝒪(1)𝒪(1))𝔽2.(D_{1}^{\prime})_{y}\cong\mathbb{P}_{\mathbb{P}^{1}}(\mathcal{O}(-1)\oplus\mathcal{O}(1))\cong\mathbb{F}_{2}.

This proves (3).

Finally, in type FF, we have already constructed the morphisms D1D1′′Y=Y×SZ0D_{1}^{\prime}\to D_{1}^{\prime\prime}\to Y=Y\times_{S}Z_{0}. Since Mϖ1WM_{\varpi_{1}}\otimes W^{\vee} is a family of semistable vector bundles of degree 22, Lemma 3.6.5 shows that D1′′YD_{1}^{\prime\prime}\to Y is a 1\mathbb{P}^{1}-bundle as claimed. Moreover, any rank 22 degree 2-2 subbundle of WW is necessarily also semistable, so Lemma 3.6.5 also shows that D1D1′′D_{1}^{\prime}\to D_{1}^{\prime\prime} is a 1\mathbb{P}^{1}-bundle.

If y:SpeckYy\colon\operatorname{\mathrm{Spec}}k\to Y is a geometric point over s:SpeckSs\colon\operatorname{\mathrm{Spec}}k\to S, then by Lemma 3.6.5 we have an exact sequence

0Uq(Mϖ2,y𝒪(2OE)Ws)q(Mϖ1+ϖ2,y𝒪(2OE))(π)𝒪(1)00\longrightarrow U\longrightarrow q^{*}(M_{\varpi_{2},y}\otimes\mathcal{O}(2O_{E})\otimes W_{s})\longrightarrow q^{*}(M_{\varpi_{1}+\varpi_{2},y}\otimes\mathcal{O}(2O_{E}))\otimes(\pi^{\prime})^{*}\mathcal{O}(1)\longrightarrow 0 (3.6.9)

of vector bundles on 1×Es\mathbb{P}^{1}\times E_{s} such that (D1)y=πU(D_{1}^{\prime})_{y}=\mathbb{P}\pi^{\prime}_{*}U, where π\pi^{\prime} and qq are the projections to the first and second factors respectively. Since UU is a vector bundle of rank 22 and determinant q(Mϖ1+2ϖ2,y𝒪(2OE))(π)𝒪(1)q^{*}(M_{-\varpi_{1}+2\varpi_{2},y}\otimes\mathcal{O}(2O_{E}))\otimes(\pi^{\prime})^{*}\mathcal{O}(-1), it follows that we have an isomorphism

UUdetU=Uq(Mϖ1+2ϖ2,y𝒪(2OE))(π)𝒪(1).U\overset{\sim}{\longrightarrow}U^{\vee}\otimes\det U=U^{\vee}\otimes q^{*}(M_{-\varpi_{1}+2\varpi_{2},y}\otimes\mathcal{O}(2O_{E}))\otimes(\pi^{\prime})^{*}\mathcal{O}(-1).

So the dual of (3.6.9) gives an exact sequence

0qM2ϖ1+ϖ2,y(π)𝒪(2)q(Mϖ1+ϖ2,yWs)(π)𝒪(1)U0,0\longrightarrow q^{*}M_{-2\varpi_{1}+\varpi_{2},y}\otimes(\pi^{\prime})^{*}\mathcal{O}(-2)\longrightarrow q^{*}(M_{-\varpi_{1}+\varpi_{2},y}\otimes W_{s}^{\vee})\otimes(\pi^{\prime})^{*}\mathcal{O}(-1)\longrightarrow U\longrightarrow 0,

and hence an exact sequence

0H0(Es,M2ϖ1+ϖ2,y)𝒪(2)\displaystyle 0\longrightarrow H^{0}(E_{s},M_{-2\varpi_{1}+\varpi_{2},y})\otimes\mathcal{O}(-2) H0(Es,Mϖ1+ϖ2,yWs)𝒪(1)\displaystyle\longrightarrow H^{0}(E_{s},M_{-\varpi_{1}+\varpi_{2},y}\otimes W_{s}^{\vee})\otimes\mathcal{O}(-1)
(π)UH1(Es,M2ϖ1+ϖ2,y)𝒪(2)0.\displaystyle\longrightarrow(\pi^{\prime})_{*}U\longrightarrow H^{1}(E_{s},M_{-2\varpi_{1}+\varpi_{2},y})\otimes\mathcal{O}(-2)\longrightarrow 0. (3.6.10)

If α1(y)=2ϖ1(y)ϖ2(y)0\alpha_{1}(y)=2\varpi_{1}(y)-\varpi_{2}(y)\neq 0, then H0(Es,M2ϖ1+ϖ2,y)=H1(Es,M2ϖ1+ϖ2,y)=0H^{0}(E_{s},M_{-2\varpi_{1}+\varpi_{2},y})=H^{1}(E_{s},M_{-2\varpi_{1}+\varpi_{2},y})=0, so (3.6.10) gives an isomorphism

(π)UH0(Es,Mϖ1+ϖ2,yWs)𝒪(1)=𝒪(1)𝒪(1),(\pi^{\prime})_{*}U\cong H^{0}(E_{s},M_{-\varpi_{1}+\varpi_{2},y}\otimes W_{s}^{\vee})\otimes\mathcal{O}(-1)=\mathcal{O}(-1)\oplus\mathcal{O}(-1),

so (D1)y1(𝒪(1)𝒪(1))=𝔽0(D_{1}^{\prime})_{y}\cong\mathbb{P}_{\mathbb{P}^{1}}(\mathcal{O}(-1)\oplus\mathcal{O}(-1))=\mathbb{F}_{0}. Otherwise, (3.6.10) gives an exact sequence

0𝒪(2)𝒪(1)𝒪(1)(π)U𝒪(2)0.0\longrightarrow\mathcal{O}(-2)\longrightarrow\mathcal{O}(-1)\oplus\mathcal{O}(-1)\longrightarrow(\pi^{\prime})_{*}U\longrightarrow\mathcal{O}(-2)\longrightarrow 0.

Since the cokernel of the injective morphism 𝒪(2)𝒪(1)𝒪(1)\mathcal{O}(-2)\to\mathcal{O}(-1)\oplus\mathcal{O}(-1) must be isomorphic to 𝒪\mathcal{O}, we get (π)U𝒪(2)𝒪(\pi^{\prime})_{*}U\cong\mathcal{O}(-2)\oplus\mathcal{O} and hence (D1)y𝔽2(D_{1}^{\prime})_{y}\cong\mathbb{F}_{2}. This completes the proof of (5) and of the proposition. ∎

4. Singularities

In this section, we apply the results of §3 to the study of the singularities of the unstable varieties χZ1(0)\chi_{Z}^{-1}(0) and their deformations χZ:ZY^//W\chi_{Z}\colon Z\to\widehat{Y}{/\mkern-6.0mu/}W. We describe the singularities explicitly in §4.1, which are given in Theorem 4.1.3. In §4.2, we briefly sketch a minor variation on standard deformation theory (in which all deformation rings are graded by the character lattice of a torus) before stating and proving weighted miniversality of the deformations χZ\chi_{Z} (Theorem 4.2.9). Theorems 4.1.3 and 4.2.9 together include all the statements from Theorem 1.0.6 from the introduction.

4.1. Codimension 22 singularities of the locus of unstable bundles

Theorem 1.0.3 (and the more refined statements in Propositions 3.4.1 and 3.5.1) give very explicit descriptions of the families of normal crossings surfaces χ~Z1(0ΘY1)Y\tilde{\chi}_{Z}^{-1}(0_{\Theta_{Y}^{-1}})\to Y. We show in this section how these results can be used to give equally explicit descriptions of the unstable loci χZ1(0)\chi_{Z}^{-1}(0) (which give local models for the singularities of the unstable loci of BunG\mathrm{Bun}_{G}, since the maps ZBunGZ\to\mathrm{Bun}_{G} are slices). For the sake of simplicity, we will assume always that S=SpeckS=\operatorname{\mathrm{Spec}}k for some algebraically closed field kk.

We will see below that there is a dichotomy between the classical types (AlA_{l} for l>1l>1, BB, CC and DD) and the exceptional types (EE, FF, GG and A1A_{1}). In the exceptional types, the unstable varieties are always cones over elliptic curves, with unique isolated singularities. In the classical types, the unstable varieties have non-isolated singularities obtained via the following construction.

Construction 4.1.1.

Let π:XX\pi\colon X\to X^{\prime} be a degree 22 morphism between smooth, possibly stacky curves over kk, and let LL be a line bundle on XX. The surface obtained by gluing LL along π\pi is the affine stack over XX^{\prime} given by the spectrum of the fibre product

R{R}πn0Ln{\pi_{*}\bigoplus_{n\geq 0}L^{\otimes-n}}𝒪X{\mathcal{O}_{X^{\prime}}}π𝒪X,{\pi_{*}\mathcal{O}_{X},}

where the vertical arrow on the right is given by restriction of a function on the total space of LL to the zero section. Geometrically, SpecXR\operatorname{\mathrm{Spec}}_{X^{\prime}}R is the surface obtained by identifying points in the zero section of the total space of LL with the same image under π\pi.

Remark 4.1.2.

Assume that the characteristic of kk is not 22, let X′′X^{\prime\prime} be a surface obtained by Construction 4.1.1, and let pX′′p\in X^{\prime\prime} be a singular point. If pp does not lie over a branch point of π\pi, then the singularity at pp is of type AA_{\infty}, i.e., we can choose (formal) local coordinates xx, yy and zz at pp so that X′′X^{\prime\prime} has local equation xy=0xy=0. If pp does lie over a branch point, then the singularity is of type DD_{\infty}, i.e., we can choose local coordinates so that X′′X^{\prime\prime} has equation x2=y2zx^{2}=y^{2}z.

Theorem 4.1.3.

Assume that S=SpeckS=\operatorname{\mathrm{Spec}}k for kk an algebraically closed field and let (G,P,μ)(G,P,\mu) be a subregular Harder-Narasimhan class. Assume that (G,P,μ)(G,P,\mu) is not of type A1A_{1} (resp., (G,P,μ)(G,P,\mu) is of type A1A_{1} and kk does not have characteristic 22) and let Z=IndLG(Z0)BunG,rigZ=\mathrm{Ind}_{L}^{G}(Z_{0})\to\mathrm{Bun}_{G,rig} be the equivariant slice constructed in the proof of Theorem 2.2.6 (resp., Remark 2.2.10). Then the stack χZ1(0)Z\chi_{Z}^{-1}(0)\subseteq Z can be constructed as follows.

  1. (1)

    If (G,P,μ)(G,P,\mu) is of type AA (but not A1A_{1}), then then there are two line bundles L1L_{1} and L2L_{2} with degL1+degL2=l+1\deg L_{1}+\deg L_{2}=l+1 such that χZ1(0)\chi_{Z}^{-1}(0) is obtained by gluing the corresponding line bundle on EEE\sqcup E along the canonical map EEEE\sqcup E\to E. In particular, χZ1(0)\chi_{Z}^{-1}(0) has singularities of type AA_{\infty} only.

  2. (2)

    If (G,P,μ)(G,P,\mu) is of type BB (resp., CC, DD), then there exists a line bundle LL on EE of degree l6l-6 (resp., l4l-4, l8l-8) such that χZ1(0)\chi_{Z}^{-1}(0) is obtained by gluing LL along a degree 22 map E(1,2)E\to\mathbb{P}(1,2) (resp., E1E\to\mathbb{P}^{1}) branched over 33 (resp., 44) points. The singularities are of type AA_{\infty} at the non-branch points of (1,2)\mathbb{P}(1,2) (resp., 1\mathbb{P}^{1}) and, if the characteristic of kk is not 22, of type DD_{\infty} at the branch points.

  3. (3)

    If (G,P,μ)(G,P,\mu) is of type EE (resp., FF, GG, A1A_{1}), then χZ1(0)\chi_{Z}^{-1}(0) is the cone over EE obtained by contracting the zero section of a line bundle LL on EE of degree l9l-9 (resp., l5l-5, l3=1l-3=-1, 4-4) to a point. The singularity is simply elliptic of degree 9l9-l (resp., 5l5-l, 3l3-l, 44).

Proof.

We first prove (3). If (G,P,μ)(G,P,\mu) is of type A1A_{1}, then the claim is proved in Proposition 4.1.9 below. So assume that (G,P,μ)(G,P,\mu) is of type EE (resp., FF, GG).

By construction, χZ1(0)\chi_{Z}^{-1}(0) is affine, and the open subset

χZ1(0)reg=χZ1(0)×BunG,rigBunG,rigreg\chi_{Z}^{-1}(0)^{reg}=\chi_{Z}^{-1}(0)\times_{\mathrm{Bun}_{G,rig}}\mathrm{Bun}_{G,rig}^{reg}

is big, where BunG,rigregBunG,rig\mathrm{Bun}_{G,rig}^{reg}\subseteq\mathrm{Bun}_{G,rig} is the open substack of regular bundles of [davis19, Proposition 4.4.6]. So choosing any y:Speck0ΘY1y\colon\operatorname{\mathrm{Spec}}k\to 0_{\Theta_{Y}^{-1}}, we have

χZ1(0)=SpecH0(χZ1(0),𝒪)=SpecH0(χZ1(0)reg,𝒪)=SpecH0(χ~Z1(y)reg,𝒪),\chi_{Z}^{-1}(0)=\operatorname{\mathrm{Spec}}H^{0}(\chi_{Z}^{-1}(0),\mathcal{O})=\operatorname{\mathrm{Spec}}H^{0}(\chi_{Z}^{-1}(0)^{reg},\mathcal{O})=\operatorname{\mathrm{Spec}}H^{0}(\tilde{\chi}_{Z}^{-1}(y)^{reg},\mathcal{O}),

where χ~Z1(y)reg=χ~Z1(y)ψZ1(χZ1(0)reg)χZ1(y)reg\tilde{\chi}_{Z}^{-1}(y)^{reg}=\tilde{\chi}_{Z}^{-1}(y)\cap\psi_{Z}^{-1}(\chi_{Z}^{-1}(0)^{reg})\cong\chi_{Z}^{-1}(y)^{reg}. But by Proposition 3.4.1 (and the fact that αi\alpha_{i} is not a special root), χ~Z1(y)reg=(D1)yE\tilde{\chi}_{Z}^{-1}(y)^{reg}=(D_{1})_{y}\setminus E is the complement of the zero section in the line bundle L1=(D1)yL^{-1}=(D_{1})_{y} over E={y}×Pic0(E)E=\{y\}\times\mathrm{Pic}^{0}(E), which has (negative) degree l9l-9 (resp., l5l-5, l3l-3). So

χZ1(0)=SpecH0((D1)yE,𝒪)=Specn0H0(E,Ln)\chi_{Z}^{-1}(0)=\operatorname{\mathrm{Spec}}H^{0}((D_{1})_{y}\setminus E,\mathcal{O})=\operatorname{\mathrm{Spec}}\bigoplus_{n\geq 0}H^{0}(E,L^{\otimes n})

is a cone over EE of the asserted degree.

To prove (1) and (2), we argue as follows. Since ψZ,y𝒪=𝒪\psi_{Z,y}{\vphantom{p}}_{*}\mathcal{O}=\mathcal{O} by Proposition 4.1.8 below and χZ1(0)Z0\chi_{Z}^{-1}(0)\to Z_{0} is affine, we have

χZ1(0)=SpecZ0π𝒪D¯y=SpecZ0π𝒪Dy,\chi_{Z}^{-1}(0)=\operatorname{\mathrm{Spec}}_{Z_{0}}\pi_{*}\mathcal{O}_{\bar{D}_{y}}=\operatorname{\mathrm{Spec}}_{Z_{0}}\pi_{*}\mathcal{O}_{D_{y}},

for any choice of y:Speck0ΘY1y\colon\operatorname{\mathrm{Spec}}k\to 0_{\Theta_{Y}^{-1}}, where π:χ~Z1(y)Z0\pi\colon\tilde{\chi}_{Z}^{-1}(y)\to Z_{0} is the natural morphism and we write

D=Dαi(Z)+Dαj(Z)+Dαi+αj(Z)andD¯=χ~Z1(0ΘY1).D=D_{\alpha_{i}^{\vee}}(Z)+D_{\alpha_{j}^{\vee}}(Z)+D_{\alpha_{i}^{\vee}+\alpha_{j}^{\vee}}(Z)\quad\text{and}\quad\bar{D}=\tilde{\chi}_{Z}^{-1}(0_{\Theta_{Y}^{-1}}).

Using Theorem 1.0.3 and Propositions 3.4.1 and 3.6.1, it is easy to see that

π𝒪Dyπ𝒪(D1)y×π𝒪Eπ𝒪(D1)y,\pi_{*}\mathcal{O}_{D_{y}}\cong\pi_{*}\mathcal{O}_{(D_{1})_{y}}\times_{\pi_{*}\mathcal{O}_{E}}\pi_{*}\mathcal{O}_{(D_{1}^{\prime})_{y}},

where we have identified {y}×Pic0(E)=Dαi(Z)yDαj(Z)y\{y\}\times\mathrm{Pic}^{0}(E)=D_{\alpha_{i}^{\vee}}(Z)_{y}\cap D_{\alpha_{j}^{\vee}}(Z)_{y} with EE and by mild abuse of notation we have also written π\pi for the morphisms (D1)yZ0(D_{1})_{y}\to Z_{0}, (D1)yZ0(D_{1}^{\prime})_{y}\to Z_{0} and EZ0E\to Z_{0}. In type AA, L1=(D1)yL_{1}=(D_{1})_{y} and L2=(D1)yL_{2}=(D_{1}^{\prime})_{y} are line bundles satisfying degL1+degL2=l+1\deg L_{1}+\deg L_{2}=l+1 by Lemma 4.1.7 below, which proves (1). In type BB (resp., CC, DD), L=(D1)yL=(D_{1})_{y} is a line bundle on EE of the desired degree by Lemma 4.1.7, π𝒪(D1)y=𝒪Z0\pi_{*}\mathcal{O}_{(D_{1}^{\prime})_{y}}=\mathcal{O}_{Z_{0}}, and EZ0=(1,2)E\to Z_{0}=\mathbb{P}(1,2) (resp., 1\mathbb{P}^{1}) has degree 22 by Lemma 4.1.6. Since any degree 22 map E(1,2)E\to\mathbb{P}(1,2) (resp., E1E\to\mathbb{P}^{1}) is branched over 33 (resp., 44) points, (2) now follows. ∎

In order to prove the lemmas quoted in the proof of Theorem 4.1.3, we will appeal to the following formula for the canonical bundle of Bun~G,rig\widetilde{\mathrm{Bun}}_{G,rig}.

Proposition 4.1.4.

There exists a line bundle MM on BunG,rig\mathrm{Bun}_{G,rig} such that

KBun~G,rig/BunG,rigψM𝒪(μ𝕏(T)+(2+ρ,μ)Dμ).K_{\widetilde{\mathrm{Bun}}_{G,rig}/\mathrm{Bun}_{G,rig}}\cong\psi^{*}M\otimes\mathcal{O}\left(\sum_{\mu\in\mathbb{X}_{*}(T)_{+}}(-2+\langle\rho,\mu\rangle)D_{\mu}\right).
Proof.

This is an immediate consequence of [davis19a, Theorem 4.6.1]. ∎

Corollary 4.1.5.

For any slice ZBunG,rigZ\to\mathrm{Bun}_{G,rig}, there exists a line bundle MM on ZZ such that

KZ~/Z=ψZM𝒪(μ𝕏(T)+(2+ρ,μ)Dμ(Z)).K_{\tilde{Z}/Z}=\psi_{Z}^{*}M\otimes\mathcal{O}\left(\sum_{\mu\in\mathbb{X}_{*}(T)_{+}}(-2+\langle\rho,\mu\rangle)D_{\mu}(Z)\right).
Lemma 4.1.6.

Assume that (G,P,μ)(G,P,\mu) is of type BB, CC or DD. Then for any y:SpeckY=0ΘY1y\colon\operatorname{\mathrm{Spec}}k\to Y=0_{\Theta_{Y}^{-1}}, the morphism Pic0(E)={y}×Pic0(E)χ~Z1(y)Z0\mathrm{Pic}^{0}(E)=\{y\}\times\mathrm{Pic}^{0}(E)\subseteq\tilde{\chi}_{Z}^{-1}(y)\to Z_{0} has degree 22.

Proof.

By Corollary 4.1.5,

KZ~=ψZKZKZ~/Z=ψZM𝒪(Dαi(Z)Dαj(Z))K_{\tilde{Z}}=\psi_{Z}^{*}K_{Z}\otimes K_{\tilde{Z}/Z}=\psi_{Z}^{*}M\otimes\mathcal{O}(-D_{\alpha_{i}^{\vee}}(Z)-D_{\alpha_{j}^{\vee}}(Z))

for some line bundle MM on ZZ. So by adjunction, we have

KDαi(Z)y=(KZ~𝒪(Dαi(Z)))|Dαi(Z)y=ψZM|Dαi(Z)y𝒪(E),K_{D_{\alpha_{i}^{\vee}}(Z)_{y}}=(K_{\tilde{Z}}\otimes\mathcal{O}(D_{\alpha_{i}^{\vee}}(Z)))|_{D_{\alpha_{i}^{\vee}}(Z)_{y}}=\psi_{Z}^{*}M|_{D_{\alpha_{i}^{\vee}}(Z)_{y}}\otimes\mathcal{O}(-E), (4.1.1)

where we write E={y}×Pic0(E)Dαi(Z)yE=\{y\}\times\mathrm{Pic}^{0}(E)\subseteq D_{\alpha_{i}^{\vee}}(Z)_{y}. To compute the degree of the finite morphism EZ0E\to Z_{0}, choose a kk-point zZ0z\in Z_{0} disjoint from the images of θk(y)\theta_{k}^{\prime}(y) and the stacky point in type BB, and let Fzk1F_{z}\cong\mathbb{P}^{1}_{k} be the fibre of Dαi(Z)yZ0D_{\alpha_{i}^{\vee}}(Z)_{y}\to Z_{0} over zz. By (4.1.1) and adjunction, the degree is the intersection product

EFz=KDαi(Z)yFz=(KDαi(Z)y+Fz)Fz=degKFz=2,E\cdot F_{z}=-K_{D_{\alpha_{i}^{\vee}}(Z)_{y}}\cdot F_{z}=-(K_{D_{\alpha_{i}^{\vee}}(Z)_{y}}+F_{z})\cdot F_{z}=-\deg K_{F_{z}}=2,

which proves the lemma. ∎

Lemma 4.1.7.

Assume we are in the setup of Propositions 3.4.1 and 3.6.1 and fix a geometric point y:SpeckYy\colon\operatorname{\mathrm{Spec}}k\to Y. Then we have the following.

  1. (1)

    If (G,P,μ)(G,P,\mu) is of type AA (not A1A_{1}), then sum of the degrees of the line bundles (D1)y(D_{1})_{y} and (D1)y(D_{1}^{\prime})_{y} on Pic0(E)\mathrm{Pic}^{0}(E) is l+1l+1.

  2. (2)

    If (G,P,μ)(G,P,\mu) is not of type AA, then the degree of the line bundle (D1)y(D_{1})_{y} on Pic0(E)\mathrm{Pic}^{0}(E) is given in Table 3.

Type BB CC DD EE FF GG
deg(D1)y\deg(D_{1})_{y} l6l-6 l4l-4 l8l-8 l9l-9 l5l-5 l3l-3
Table 3. Degree of (D1)y(D_{1})_{y}
Proof.

To simplify the notation, identify Pic0(E)(D1)y\mathrm{Pic}^{0}(E)\subseteq(D_{1})_{y} with EE. The degree of the line bundle (D1)y(D_{1})_{y} is equal to the self-intersection number (E2)(D1)y(E^{2})_{(D_{1})_{y}} of EE on the surface (D1)y(D_{1})_{y}.

First note that by Proposition 3.4.1, Dαj(Z)yD_{\alpha_{j}^{\vee}}(Z)_{y} is the iterated blowup of (D1)y(D_{1})_{y} at n0+1n_{0}+1 points on EE, so we have

(E2)(D1)y=(E2)Dαj(Z)y+n0+1.(E^{2})_{(D_{1})_{y}}=(E^{2})_{D_{\alpha_{j}^{\vee}}(Z)_{y}}+n_{0}+1. (4.1.2)

Next, observe that we have

0=χ~Z1(0ΘY1)E=(dDαi(Z)+Dαj(Z)+Dαi+αj(Z))E=d(E2)Dαj(Z)y+(E2)Dαi(Z)y+1,0=\tilde{\chi}_{Z}^{-1}(0_{\Theta_{Y}^{-1}})\cdot E=(dD_{\alpha_{i}^{\vee}}(Z)+D_{\alpha_{j}^{\vee}}(Z)+D_{\alpha_{i}^{\vee}+\alpha_{j}^{\vee}}(Z))\cdot E=d(E^{2})_{D_{\alpha_{j}^{\vee}}(Z)_{y}}+(E^{2})_{D_{\alpha_{i}^{\vee}}(Z)_{y}}+1, (4.1.3)

where d=12(αi|αi)d=\frac{1}{2}(\alpha_{i}^{\vee}{\,|\,}\alpha_{i}^{\vee}) and we have used the fact that Dαi(Z)yDαj(Z)y=ED_{\alpha_{i}^{\vee}}(Z)_{y}\cap D_{\alpha_{j}^{\vee}}(Z)_{y}=E and that the exceptional curve of the final blowup Dαi+αj(Z)yDαj(Z)yD_{\alpha_{i}^{\vee}+\alpha_{j}^{\vee}}(Z)_{y}\cap D_{\alpha_{j}^{\vee}}(Z)_{y} meets EE transversely in a single point. Since Dαj(Z)yD_{\alpha_{j}^{\vee}}(Z)_{y} is the iterated blowup of the smooth surface (D1)y(D_{1}^{\prime})_{y} of Proposition 3.6.1 at NN points on EE, we have

(E2)Dαi(Z)=(E2)(D1)yN,(E^{2})_{D_{\alpha_{i}^{\vee}}(Z)}=(E^{2})_{(D_{1}^{\prime})_{y}}-N,

and hence (4.1.2) and (4.1.3) give

(E2)(D1)y=1d(N(E2)(D1)y1)+n0+1.(E^{2})_{(D_{1})_{y}}=\frac{1}{d}(N-(E^{2})_{(D_{1}^{\prime})_{y}}-1)+n_{0}+1. (4.1.4)

In type AA, d=1d=1 and N=n1+1N=n_{1}+1, so (4.1.4) is equivalent to

deg(D1)y+deg(D1)y=(E2)(D1)y+(E2)(D1)y=n0+n1+1=l+1,\deg(D_{1})_{y}+\deg(D_{1}^{\prime})_{y}=(E^{2})_{(D_{1})_{y}}+(E^{2})_{(D_{1}^{\prime})_{y}}=n_{0}+n_{1}+1=l+1,

which proves (1).

In types BB, CC and DD, we argue as follows. By Lemma 4.1.6, EE is a smooth elliptic curve contained in a (possibly stacky) Hirzebruch surface (D1)y(D_{1}^{\prime})_{y} mapping with degree 22 to the base Z0=(1,2)Z_{0}=\mathbb{P}(1,2) or 1\mathbb{P}^{1}. It follows from a straightforward adjunction calculation that EE is an anticanonical divisor on (D1)y(D_{1}^{\prime})_{y} and satisfies (E2)D1=6(E^{2})_{D_{1}^{\prime}}=6 in type BB and (E2)(D1)y=8(E^{2})_{(D_{1}^{\prime})_{y}}=8 in types CC and DD. Substituting into (4.1.4) gives the degrees in Table 3.

Finally, in types EE, FF and GG, note that by Corollary 4.1.5, we have

KZ~/Z=ψZM𝒪(Dαi(Z)Dαj(Z))K_{\tilde{Z}/Z}=\psi_{Z}^{*}M\otimes\mathcal{O}(-D_{\alpha_{i}^{\vee}}(Z)-D_{\alpha_{j}^{\vee}}(Z))

for some line bundle on MM on ZZ. Since Z𝔸l+3Z\cong\mathbb{A}^{l+3} is an affine space, every line bundle on ZZ is trivial, so

KZ~=KZ~/ZψZKZ𝒪(Dαi(Z)Dαj(Z)).K_{\tilde{Z}}=K_{\tilde{Z}/Z}\otimes\psi_{Z}^{*}K_{Z}\cong\mathcal{O}(-D_{\alpha_{i}^{\vee}}(Z)-D_{\alpha_{j}^{\vee}}(Z)).

By adjunction, we therefore have a linear equivalence

KDαi(Z)y(KZ~+Dαi(Z))|Dαi(Z)y=Dαi(Z)yDαj(Z)y=E.K_{D_{\alpha_{i}^{\vee}}(Z)_{y}}\sim(K_{\tilde{Z}}+D_{\alpha_{i}^{\vee}}(Z))|_{D_{\alpha_{i}^{\vee}}(Z)_{y}}=-D_{\alpha_{i}^{\vee}}(Z)_{y}\cap D_{\alpha_{j}^{\vee}}(Z)_{y}=-E.

So EDαi(Z)yE\subseteq D_{\alpha_{i}^{\vee}}(Z)_{y} is an anticanonical divisor, from which it follows that E(D1)yE\subseteq(D_{1}^{\prime})_{y} is also an anticanonical divisor in the blow down. So from the explicit identification of the surface (D1)y(D_{1}^{\prime})_{y} given in Proposition 3.6.1 as either a Hirzebruch surface or 2\mathbb{P}^{2}, we have

(E2)(D1)y=K(D1)y2={9,in typesEandG,8,in typeF.(E^{2})_{(D_{1}^{\prime})_{y}}=K_{(D_{1}^{\prime})_{y}}^{2}=\begin{cases}9,&\text{in types}\;E\;\text{and}\;G,\\ 8,&\text{in type}\;\;F.\end{cases}

Substituting the values of NN, n0n_{0} and dd into (4.1.4) in each of the different cases gives the desired expressions for (E2)(D1)y(E^{2})_{(D_{1})_{y}}. ∎

Proposition 4.1.8.

Fix any geometric point y:SpeckΘY1y\colon\operatorname{\mathrm{Spec}}k\to\Theta_{Y}^{-1}, and let

ψy:χ~1(y)χ1(y)\psi_{y}\colon\tilde{\chi}^{-1}(y)\longrightarrow\chi^{-1}(y)

be the pullback of the elliptic Grothendieck-Springer resolution. We have ψy𝒪=𝒪\psi_{y}{\vphantom{p}}_{*}\mathcal{O}=\mathcal{O}.

Proof.

Since χ1(y)\chi^{-1}(y) is a local complete intersection, hence Cohen-Macaulay, it is enough to prove the claim on an open substack of the target whose complement has codimension at least 22. We therefore reduce to proving that the map

ψZ,y:χ~Z1(y)χZ1(y)\psi_{Z,y}\colon\tilde{\chi}_{Z}^{-1}(y)\longrightarrow\chi_{Z}^{-1}(y)

satisfies ψZ,y𝒪=𝒪\psi_{Z,y}{\vphantom{p}}_{*}\mathcal{O}=\mathcal{O} for each of the slices ZBunG,rigZ\to\mathrm{Bun}_{G,rig} of Theorem 1.0.2.

Note that by the proof of Lemma 3.1.4, we have ψZ𝒪=𝒪\psi_{Z}^{\prime}{\vphantom{p}}_{*}\mathcal{O}=\mathcal{O}, where

ψZ:Z~Z×Y^//WΘY1\psi_{Z}^{\prime}\colon\tilde{Z}\longrightarrow Z\times_{\widehat{Y}{/\mkern-6.0mu/}W}\Theta_{Y}^{-1}

is the natural morphism induced by ψZ\psi_{Z}. Since the domain and codomain of ψZ\psi_{Z}^{\prime} are both flat over ΘY1\Theta_{Y}^{-1}, it is enough by base change to show that iψZ𝒪=0\mathbb{R}^{i}\psi_{Z}^{\prime}{\vphantom{p}}_{*}\mathcal{O}=0 for all i>0i>0. By equivariance, this will follow from the claim that iψZ,y𝒪=0\mathbb{R}^{i}\psi_{Z,y}{\vphantom{p}}_{*}\mathcal{O}=0 for all i>0i>0 and all y0ΘY1=Yy\in 0_{\Theta_{Y}^{-1}}=Y.

Since χZ1(0)Z0\chi_{Z}^{-1}(0)\to Z_{0} is affine by construction, it is enough to show that iπ𝒪=0\mathbb{R}^{i}\pi_{*}\mathcal{O}=0 for i>0i>0, where π:χ~Z1(y)Z0\pi\colon\tilde{\chi}_{Z}^{-1}(y)\to Z_{0} is the natural morphism. This holds by inspection for the fibre over yYy\in Y of the reduced normal crossings variety

D=Dαi(Z)+Dαj(Z)+Dαi+αj(Z),D=D_{\alpha_{i}^{\vee}}(Z)+D_{\alpha_{j}^{\vee}}(Z)+D_{\alpha_{i}^{\vee}+\alpha_{j}^{\vee}}(Z),

from the explicit descriptions of the components given by Theorem 1.0.3 and Proposition 3.6.1, using the fact that f𝒪=𝒪\mathbb{R}f_{*}\mathcal{O}=\mathcal{O} whenever ff is either a 1\mathbb{P}^{1}-bundle or the blow up of a smooth surface at a point. This proves the claim in types AA, BB and DD. In type CC, we claim that the morphism π𝒪D¯yπ𝒪Dy\mathbb{R}\pi_{*}\mathcal{O}_{\bar{D}_{y}}\to\mathbb{R}\pi_{*}\mathcal{O}_{D_{y}} is a quasi-isomorphism, where D¯=χ~Z1(0ΘY1)\bar{D}=\tilde{\chi}_{Z}^{-1}(0_{\Theta_{Y}^{-1}}), from which the desired vanishing follows. To see this, note that we have a short exact sequence

0𝒪(D)|Dαi(Z)𝒪D¯𝒪D0,0\longrightarrow\mathcal{O}(-D)|_{D_{\alpha_{i}^{\vee}}(Z)}\longrightarrow\mathcal{O}_{\bar{D}}\longrightarrow\mathcal{O}_{D}\longrightarrow 0,

so it is enough to show that iπ𝒪(D)|Dαi(Z)y=0\mathbb{R}^{i}\pi_{*}\mathcal{O}(-D)|_{D_{\alpha_{i}^{\vee}}(Z)_{y}}=0 for all ii. From the explicit description of Dαi(Z)yD_{\alpha_{i}^{\vee}}(Z)_{y} given in Proposition 3.6.1, it is enough to show that 𝒪(D)|Dαi(Z)y\mathcal{O}(-D)|_{D_{\alpha_{i}^{\vee}}(Z)_{y}} has degree 0 on the exceptional curve γ\gamma of the blowup and degree 1-1 on every irreducible fibre of D1Z0=1D_{1}^{\prime}\to Z_{0}=\mathbb{P}^{1}. But since ΘY\Theta_{Y} is trivial on Dαi(Z)yD_{\alpha_{i}^{\vee}}(Z)_{y}, we have a linear equivalence

2D|Dαi(Z)yDαj(Z)yDαi(Z)yDαi+αj(Z)yDαi(Z)y=Eγ,-2D|_{D_{\alpha_{i}^{\vee}}(Z)_{y}}\sim-D_{\alpha_{j}^{\vee}}(Z)_{y}\cap D_{\alpha_{i}^{\vee}}(Z)_{y}-D_{\alpha_{i}^{\vee}+\alpha_{j}^{\vee}}(Z)_{y}\cap D_{\alpha_{i}^{\vee}}(Z)_{y}=-E-\gamma,

from which the claim follows by Lemma 4.1.6. ∎

Proposition 4.1.9.

Assume that (G,P,μ)(G,P,\mu) is of type A1A_{1}, so that G=SL2G=SL_{2}, P=TP=T and ϖ1,μ=2\langle\varpi_{1},\mu\rangle=-2. Let SpeckBunTμ=Bun𝔾m2\operatorname{\mathrm{Spec}}k\to\mathrm{Bun}_{T}^{\mu}=\mathrm{Bun}_{\mathbb{G}_{m}}^{-2} be the slice classifying the line bundle 𝒪(2OE)\mathcal{O}(-2O_{E}) of Remark 2.2.10 and let Z=IndTSL2(Speck)BunSL2Z=\mathrm{Ind}_{T}^{SL_{2}}(\operatorname{\mathrm{Spec}}k)\to\mathrm{Bun}_{SL_{2}} be the induced equivariant slice. Then the unstable fibre χZ1(0)\chi_{Z}^{-1}(0) is isomorphic to the affine cone over EE obtained by contracting the zero section of a degree 4-4 line bundle to a point.

Proof.

If we identify BunSL2\mathrm{Bun}_{SL_{2}} with the stack of rank 22 vector bundles with trivial determinant, then the slice ZZ is nothing but the vector space

Z=Ext1(𝒪(2OE),𝒪(2OE))H1(E,𝒪(4OE)),Z=\mathrm{Ext}^{1}(\mathcal{O}(2O_{E}),\mathcal{O}(-2O_{E}))\cong H^{1}(E,\mathcal{O}(-4O_{E})),

with its tautological map to BunSL2\mathrm{Bun}_{SL_{2}}. To describe the unstable locus, note that a point zZz\in Z corresponds to an unstable extension

0𝒪(2OE)Vz𝒪(2OE)00\longrightarrow\mathcal{O}(-2O_{E})\longrightarrow V_{z}\longrightarrow\mathcal{O}(2O_{E})\longrightarrow 0

if and only if there exists a degree 11 line bundle LL on EE such that zz is in the (11-dimensional) kernel of the map

Ext1(𝒪(2OE),𝒪(2OE))Ext1(L,𝒪(2OE))\mathrm{Ext}^{1}(\mathcal{O}(2O_{E}),\mathcal{O}(-2O_{E}))\longrightarrow\mathrm{Ext}^{1}(L,\mathcal{O}(-2O_{E}))

induced by the unique (up to scale) nonzero morphism L𝒪(2OE)L\to\mathcal{O}(2O_{E}). We deduce that χZ1(0){0}\chi_{Z}^{-1}(0)\setminus\{0\} must be a 𝔾m\mathbb{G}_{m}-torsor over Pic1(E)E\mathrm{Pic}^{1}(E)\cong E, so the normal variety χZ1(0)\chi_{Z}^{-1}(0) must be an affine cone over EE as claimed.

To identify the degree, observe that since ZBunSL2Z\to\mathrm{Bun}_{SL_{2}} is an equivariant slice with equivariance group 𝔾m\mathbb{G}_{m} and weight 22 by Proposition 2.2.5, the morphism χZ:𝔸4ZY^//W𝔸2\chi_{Z}\colon\mathbb{A}^{4}\cong Z\to\widehat{Y}{/\mkern-6.0mu/}W\cong\mathbb{A}^{2} is equivariant with respect to the weight 11 action on 𝔸4\mathbb{A}^{4} and the weight 22 action on 𝔸2\mathbb{A}^{2}. So taking projectivisations, we deduce that the elliptic curve (χZ1(0){0})/𝔾m(\chi_{Z}^{-1}(0)\setminus\{0\})/\mathbb{G}_{m} is presented as an intersection of two quadric surfaces in 3\mathbb{P}^{3}, from which we deduce that the polarising line bundle has degree 44. The proposition now follows. ∎

4.2. Deformation theory

In this subsection, we study the deformation theory of the unstable varieties χZ1(0)\chi_{Z}^{-1}(0) of §4.1. As in the previous subsection, We will assume for simplicity that S=SpeckS=\operatorname{\mathrm{Spec}}k for some algebraically closed field kk.

Definition 4.2.1.

Let HH be a torus with character group 𝕏(H)\mathbb{X}^{*}(H), let 𝕏(H)+𝕏(H)\mathbb{X}^{*}(H)_{+}\subseteq\mathbb{X}^{*}(H) be a sub-monoid (without unit), and let XX be an algebraic stack over Speck\operatorname{\mathrm{Spec}}k with HH-action.

  1. (1)

    An 𝕏(H)+\mathbb{X}^{*}(H)_{+}-weighted deformation ring is an 𝕏(H)\mathbb{X}^{*}(H)-graded Noetherian kk-algebra

    R=λ𝕏(H)+{0}RλR=\bigoplus_{\lambda\in-\mathbb{X}^{*}(H)_{+}\cup\{0\}}R_{\lambda}

    such that R0=kR_{0}=k. Given such an RR, we write

    R^=λ𝕏(H)+{0}Rλ\widehat{R}=\prod_{\lambda\in-\mathbb{X}^{*}(H)_{+}\cup\{0\}}R_{\lambda}

    for the completion at the maximal ideal 𝔪R=λ𝕏(H)+Rλ\mathfrak{m}_{R}=\bigoplus_{\lambda\in-\mathbb{X}^{*}(H)_{+}}R_{\lambda}.

  2. (2)

    An 𝕏(H)+\mathbb{X}^{*}(H)_{+}-weighted deformation of XX over an 𝕏(H)+\mathbb{X}^{*}(H)_{+}-weighted deformation ring RR is a flat HH-equivariant morphism X¯SpfR^\bar{X}\to\operatorname{\mathrm{Spf}}\widehat{R} of formal stacks equipped with an HH-equivariant isomorphism X¯sX\bar{X}_{s}\cong X, where s:SpeckSpfR^s\colon\operatorname{\mathrm{Spec}}k\to\operatorname{\mathrm{Spf}}\widehat{R} is the unique (HH-fixed) point.

  3. (3)

    We say that an 𝕏(H)+\mathbb{X}^{*}(H)_{+}-weighted deformation X¯SpfR\bar{X}\to\operatorname{\mathrm{Spf}}R is versal if for every surjective (graded) homomorphism RR′′R^{\prime}\to R^{\prime\prime} of 𝕏(H)+\mathbb{X}^{*}(H)_{+}-weighted deformation rings, every homomorphism ϕ:RR′′\phi^{\prime}\colon R\to R^{\prime\prime} and every weighted deformation X¯RSpfR^\bar{X}_{R^{\prime}}\to\operatorname{\mathrm{Spf}}\widehat{R}^{\prime} with an isomorphism α:X¯R×SpfR^SpfR^′′X¯×SpfR^SpfR^′′\alpha\colon\bar{X}_{R^{\prime}}\times_{\operatorname{\mathrm{Spf}}\widehat{R}^{\prime}}\operatorname{\mathrm{Spf}}\widehat{R}^{\prime\prime}\cong\bar{X}\times_{\operatorname{\mathrm{Spf}}\widehat{R}}\operatorname{\mathrm{Spf}}\widehat{R}^{\prime\prime}, there exists a lift ϕ:RR\phi\colon R\to R^{\prime} and an isomorphism X¯RX¯×SpfR^SpfR^\bar{X}_{R^{\prime}}\cong\bar{X}\times_{\operatorname{\mathrm{Spf}}\widehat{R}}\operatorname{\mathrm{Spf}}\widehat{R}^{\prime} lifting α\alpha.

  4. (4)

    We say that a versal 𝕏(H)+\mathbb{X}^{*}(H)_{+}-weighted deformation X¯SpfR\bar{X}\to\operatorname{\mathrm{Spf}}R is miniversal (or semi-universal) if for all 𝕏(H)+\mathbb{X}^{*}(H)_{+}-weighted deformation rings RR^{\prime} and pairs ϕ,ϕ:RR\phi,\phi^{\prime}\colon R\to R^{\prime} of graded homomorphisms with X¯×SpfR^,ϕSpfR^X¯×SpfR^,ϕSpfR^\bar{X}\times_{\operatorname{\mathrm{Spf}}\widehat{R},\phi}\operatorname{\mathrm{Spf}}\widehat{R}^{\prime}\cong\bar{X}\times_{\operatorname{\mathrm{Spf}}\widehat{R},\phi^{\prime}}\operatorname{\mathrm{Spf}}\widehat{R}^{\prime}, the maps

    dϕ,dϕ:TsSpecRTsSpecRd\phi,d\phi^{\prime}\colon T_{s^{\prime}}\operatorname{\mathrm{Spec}}R^{\prime}\longrightarrow T_{s}\operatorname{\mathrm{Spec}}R

    on tangent spaces at fixed points are equal.

Remark 4.2.2.

Note that a versal (resp., miniversal) 𝕏(H)+\mathbb{X}_{*}(H)_{+}-weighted deformation need not be versal (resp., miniversal) as a plain unweighted deformation.

Weighted deformation theory in this sense works in more or less the same way as unweighted deformation theory for schemes. (See, for example, [hartshorne10] or [illusie71, Chapitre III] for the unweighted case.) For example, we have the following.

Proposition 4.2.3.

Let 𝕋X\mathbb{T}_{X} be the tangent complex to XX (the 𝒪X\mathcal{O}_{X}-linear dual to the cotangent complex). Let RRR^{\prime}\to R is a surjection of Artinian 𝕏(H)+\mathbb{X}_{*}(H)_{+}-weighted deformation rings with kernel II satisfying 𝔪RI=0\mathfrak{m}_{R^{\prime}}I=0, and let X¯SpfR^=SpecR\bar{X}\to\operatorname{\mathrm{Spf}}\widehat{R}=\operatorname{\mathrm{Spec}}R be a weighted deformation of XX.

  1. (1)

    There is an HH-invariant obstruction class ob(H2(X,𝕋X)I)H\mathrm{ob}\in(H^{2}(X,\mathbb{T}_{X})\otimes I)^{H} such that X¯\bar{X} lifts to a weighted deformation over RR^{\prime} if and only if ob=0\mathrm{ob}=0.

  2. (2)

    If lifts exist, then the set of isomorphism classes of HH-equivariant lifts form a torsor under the group (H1(X,𝕋X)I)HH1(X,𝕋X)I(H^{1}(X,\mathbb{T}_{X})\otimes I)^{H}\subseteq H^{1}(X,\mathbb{T}_{X})\otimes I.

As a consequence, we deduce the following via the usual argument for existence and behaviour of a miniversal deformation.

Proposition 4.2.4.

Assume that 𝕏(H)+\mathbb{X}^{*}(H)_{+}-weighted subspace H1(X,𝕋X)+H^{1}(X,\mathbb{T}_{X})_{+} of H1(X,𝕋X)H^{1}(X,\mathbb{T}_{X}) is finite dimensional. Then there exists a miniversal 𝕏(H)+\mathbb{X}_{*}(H)_{+}-weighted deformation X¯SpfR^\bar{X}\to\operatorname{\mathrm{Spf}}\widehat{R}, and SpfR^\operatorname{\mathrm{Spf}}\widehat{R} has tangent space H1(X,𝕋X)+H^{1}(X,\mathbb{T}_{X})_{+} at the fixed point. Moreover, if the 𝕏(H)+\mathbb{X}^{*}(H)_{+}-weighted subspace H2(X,𝕋X)+H^{2}(X,\mathbb{T}_{X})_{+} of H2(X,𝕋X)H^{2}(X,\mathbb{T}_{X}) vanishes, then

RSym(H1(X,𝕋X)+)R\cong\operatorname{\mathrm{Sym}}(H^{1}(X,\mathbb{T}_{X})_{+})^{\vee}

as 𝕏(H)\mathbb{X}^{*}(H)-graded rings.

Remark 4.2.5.

The obstruction class of Proposition 4.2.3 is the same as the obstruction for lifting ordinary (unweighted) deformations. It follows that if X¯SpfR^\bar{X}\to\operatorname{\mathrm{Spf}}\widehat{R} is an unweighted miniversal deformation such that RR is 𝕏(H)\mathbb{X}^{*}(H)-graded and the HH-action on XX lifts to a compatible action on X¯\bar{X}, then the restriction to SpfR^\operatorname{\mathrm{Spf}}\widehat{R}^{\prime} is an 𝕏(H)+\mathbb{X}^{*}(H)_{+}-graded miniversal deformation, where RR^{\prime} is the quotient of RR by the ideal generated by all weight spaces of the maximal ideal of RR with weights 𝕏(H)+\not\in\mathbb{X}_{*}(H)_{+}.

We now turn to the weighted deformation theory of the singularities of §4.1.

Lemma 4.2.6.

Assume we are in the setup of Construction 4.1.1, and assume moreover that either kk has characteristic not 22 or that the morphism π:XX\pi\colon X\to X^{\prime} is unramified. Let X′′X^{\prime\prime} be the surface obtained by gluing a line bundle LL on XX along π\pi. Let HH be a torus equipped with a sub-monoid 𝕏(H)+𝕏(H)\mathbb{X}^{*}(H)_{+}\subseteq\mathbb{X}^{*}(H) (not containing 0) acting on the line bundle LL (and hence on the surface X′′X^{\prime\prime}) in such a way that for every xXx\in X^{\prime}, the weights λ1\lambda_{1} and λ2\lambda_{2} at the two preimages of xx satisfy

λ1,λ2𝕏(H)+andλ1λ2,λ2λ1𝕏(H)+.\lambda_{1},\lambda_{2}\in\mathbb{X}^{*}(H)_{+}\quad\text{and}\quad\lambda_{1}-\lambda_{2},\lambda_{2}-\lambda_{1}\not\in\mathbb{X}^{*}(H)_{+}.

Then, in the notation of Proposition 4.2.4,

H1(X′′,𝕋X′′)+=H0(X,I)xram(π)Lx1Iπ(x),andH2(X′′,𝕋X′′)+=H1(X,I),H^{1}(X^{\prime\prime},\mathbb{T}_{X^{\prime\prime}})_{+}=H^{0}(X^{\prime},I^{\vee})\oplus\bigoplus_{x\in\mathrm{ram}(\pi)}L_{x}^{-1}\otimes I_{\pi(x)}^{\vee},\quad\text{and}\quad H^{2}(X^{\prime\prime},\mathbb{T}_{X^{\prime\prime}})_{+}=H^{1}(X^{\prime},I^{\vee}),

where ram(π)X\mathrm{ram}(\pi)\subseteq X is the set of ramification points of π\pi and

I=ker(Sym2π(L1)π(L2)).I=\ker(\operatorname{\mathrm{Sym}}^{2}\pi_{*}(L^{-1})\longrightarrow\pi_{*}(L^{-2})).
Proof.

First note that since HH acts on LL with weights in 𝕏(H)+\mathbb{X}_{*}(H)_{+} it follows that Hi(X′′,f𝕋X)H^{i}(X^{\prime\prime},f^{*}\mathbb{T}_{X^{\prime}}) has weights in 𝕏(H)+-\mathbb{X}^{*}(H)_{+} (and hence none in 𝕏(H)+\mathbb{X}_{*}(H)_{+}) for all ii, where f:X′′Xf\colon X^{\prime\prime}\to X^{\prime} is the structure map. So we can replace 𝕋X′′\mathbb{T}_{X^{\prime\prime}} with 𝕋X′′/X\mathbb{T}_{X^{\prime\prime}/X^{\prime}} in the statement of the lemma. We can also assume without loss of generality that XX^{\prime} is connected, so that the weights {λ1,λ2}\{\lambda_{1},\lambda_{2}\} of HH on LL restricted to π1(x)\pi^{-1}(x) are independent of xXx\in X^{\prime}.

By definition, we have X′′=SpecXX^{\prime\prime}=\operatorname{\mathrm{Spec}}_{X^{\prime}}\mathcal{R}, where \mathcal{R} is the sheaf of algebras

=πn0Ln×π𝒪X𝒪XSymπ(L1)Symπ(L1)I.\mathcal{R}=\pi_{*}\bigoplus_{n\geq 0}L^{-n}\times_{\pi_{*}\mathcal{O}_{X}}\mathcal{O}_{X^{\prime}}\cong\frac{\operatorname{\mathrm{Sym}}\pi_{*}(L^{-1})}{\operatorname{\mathrm{Sym}}\pi_{*}(L^{-1})\otimes I}.

So X′′X^{\prime\prime} is a local complete intersection over XX^{\prime} and the pushforwards of the relative tangent complex along the structure map f:X′′Xf\colon X^{\prime\prime}\to X^{\prime} is

f𝕋X′′/X=[Symπ(L1)π(L1)Symπ(L1)I],f_{*}\mathbb{T}_{X^{\prime\prime}/X^{\prime}}=[\operatorname{\mathrm{Sym}}\pi_{*}(L^{-1})\otimes\pi_{*}(L^{-1})^{\vee}\to\operatorname{\mathrm{Sym}}\pi_{*}(L^{-1})\otimes I^{\vee}],

concentrated in cohomological degrees 0 and 11. Since π(L1)\pi_{*}(L^{-1}) has weights λ1-\lambda_{1} and λ2-\lambda_{2} and II^{\vee} has weight λ1+λ2\lambda_{1}+\lambda_{2}, it follows that the 𝕏(H)+\mathbb{X}^{*}(H)_{+}-weight part of f𝕋X′′/Xf_{*}\mathbb{T}_{X^{\prime\prime}/X^{\prime}} is

(f𝕋X′′/X)+=I[1][π(L1)π(L1)I].(f_{*}\mathbb{T}_{X^{\prime\prime}/X^{\prime}})_{+}=I^{\vee}[-1]\oplus[\pi_{*}(L^{-1})^{\vee}\to\pi_{*}(L^{-1})\otimes I^{\vee}].

A local computation now shows that the second term has vanishing cohomology away from the ramification points. If kk does not have characteristic 22, then π\pi is given in formal local coordinates by xx2x\mapsto x^{2} near any ramification point, from which one can show that the natural map

[π(L1)π(L1)I]xram(π)Lx1Iπ(x)[1][\pi_{*}(L^{-1})^{\vee}\to\pi_{*}(L^{-1})\otimes I^{\vee}]\longrightarrow\bigoplus_{x\in\mathrm{ram}(\pi)}L^{-1}_{x}\otimes I^{\vee}_{\pi(x)}[-1]

is a quasi-isomorphism, where the terms of the direct sum on the right are interpreted as skyscraper sheaves at the branch points π(x)X\pi(x)\in X^{\prime}. The lemma now follows. ∎

Lemma 4.2.7.

Let (G,P,μ)(G,P,\mu) be a subregular Harder-Narasimhan class, and identify the torus Z(L)rigZ(L)_{rig} with 𝔾m\mathbb{G}_{m} (resp., 𝔾m×𝔾m\mathbb{G}_{m}\times\mathbb{G}_{m} in type AA) via the cocharacter ϖi:𝔾mZ(L)rig-\varpi_{i}^{\vee}\colon\mathbb{G}_{m}\to Z(L)_{rig} (resp., (ϖi,ϖi+1):𝔾m×𝔾mZ(L)rig(-\varpi_{i}^{\vee},-\varpi_{i+1})\colon\mathbb{G}_{m}\times\mathbb{G}_{m}\to Z(L)_{rig}), where ii is as in Notation 2.3.2. The weights of the Z(L)rigZ(L)_{rig}-action on the line bundles in Theorem 4.1.3 are as follows.

  1. (1)

    If (G,P,μ)(G,P,\mu) is of type AA (but not A1A_{1}), then the line bundles L1L_{1} and L2L_{2} have weights (1,0)(1,0) and (0,1)(0,1).

  2. (2)

    In all other cases, the line bundle LL has weight 11.

Proof.

We deduce this from the weights of the Z(L)rigZ(L)_{rig}-action on the affine space Y^//W\widehat{Y}{/\mkern-6.0mu/}W and the fibres of the affine space bundle ZZ0Z\to Z_{0}. By construction, the Z(L)rigZ(L)_{rig}-weights of Y^//W\widehat{Y}{/\mkern-6.0mu/}W are the canonical 𝔾m\mathbb{G}_{m}-weights multiplied by (μ|)(\mu{\,|\,}-). By a theorem of Looijenga [looijenga76, Theorem 3.4], these 𝔾m\mathbb{G}_{m} weights are 1,g1,,gl1,g_{1},\ldots,g_{l}, where gig_{i} are the coroot integers defined by

g1α1++glαl=α~,g_{1}\alpha_{1}^{\vee}+\cdots+g_{l}\alpha_{l}^{\vee}=\tilde{\alpha}^{\vee},

where α~\tilde{\alpha} is the highest root of GG. The Z(L)rigZ(L)_{rig}-weights on ZZ, on the other hand, can be computed using, for example, the formula in [davis19, Proposition 4.1.7] for the weight multiplicities in a parabolic induction. These are all given in Table 4 below.

Type 𝔾m\mathbb{G}_{m}-weights of Y^//W\widehat{Y}{/\mkern-6.0mu/}W (μ|)(\mu{\,|\,}-) Z(L)rigZ(L)_{rig}-weights of ZZ
A1A_{1} 121^{2} 22 141^{4}
Al,l>1A_{l},l>1 1l+11^{l+1} (1,1)(1,1) (1,0)1(0,1)1(1,1)l(1,0)^{1}(0,1)^{1}(1,1)^{l}
BlB_{l} 132l21^{3}2^{l-2} 11 152l31^{5}2^{l-3}
ClC_{l} 1l+11^{l+1} 22 122l1^{2}2^{l}
DlD_{l} 142l31^{4}2^{l-3} 11 162l41^{6}2^{l-4}
E5E_{5} 14221^{4}2^{2} 11 181^{8}
E6E_{6} 1323311^{3}2^{3}3^{1} 11 16231^{6}2^{3}
E7E_{7} 122232411^{2}2^{2}3^{2}4^{1} 11 1424321^{4}2^{4}3^{2}
E8E_{8} 1122324251611^{1}2^{2}3^{2}4^{2}5^{1}6^{1} 11 12233342511^{2}2^{3}3^{3}4^{2}5^{1}
F3F_{3} 13211^{3}2^{1} 22 12241^{2}2^{4}
F4F_{4} 122231^{2}2^{2}3 22 112331421^{1}2^{3}3^{1}4^{2}
G2G_{2} 12211^{2}2^{1} 33 1121331^{1}2^{1}3^{3}
Table 4. Weights of the subregular slices

In type AlA_{l}, l>1l>1, observe that the fixed loci of {1}×𝔾m\{1\}\times\mathbb{G}_{m} and 𝔾m×{1}Z(L)rig\mathbb{G}_{m}\times\{1\}\subseteq Z(L)_{rig} are necessarily contained in the zero fibre χZ1(0)\chi_{Z}^{-1}(0). Since these are both line bundles over E=Z0E=Z_{0}, they must be L1L_{1} and L2L_{2}. So the weights of these are (1,0)(1,0) and (0,1)(0,1) as claimed.

In the other types, assume for a contradiction that LL has weight w>1w>1. (Note that it cannot have negative weight, since all weights of ZZ are positive.) So the whole zero fibre χZ1(0)\chi_{Z}^{-1}(0) is contained in the fixed locus of μw\mu_{w}, and hence the images of the weight 1-1 generators of the polynomial ring Γ(Y^//W,𝒪)\Gamma(\widehat{Y}{/\mkern-6.0mu/}W,\mathcal{O}) span the weight 1-1 part of p𝒪Zp_{*}\mathcal{O}_{Z}, where p:ZZ0p\colon Z\to Z_{0} is the natural map. But in each case the multiplicity of the weight 1-1 in p𝒪Zp_{*}\mathcal{O}_{Z} is larger than in Γ(Y^//W,𝒪)\Gamma(\widehat{Y}{/\mkern-6.0mu/}W,\mathcal{O}), so this is a contradiction, and the lemma is proved. ∎

Lemma 4.2.8.

Let XX be a cone over EE of degree 1d41\leq d\leq 4, and assume that (char(k),d)(2,2),(3,3)(\mathrm{char}(k),d)\neq(2,2),(3,3). Then the miniversal >0\mathbb{Z}_{>0}-deformation space of XX is an affine space with the same weights as Y^//W\widehat{Y}{/\mkern-6.0mu/}W in type E9dE_{9-d} given in Table 4.

Proof.

Except for d=4d=4, this is pointed out in [grojnowski-shep19, Theorem 6.24]: the result is well-known in characteristic 0, and is due to M. Hirokado [hirokado04, Theorem 4.4] in positive characteristic.

For d=4d=4, we argue as follows. In this case, the cone XX is a complete intersection

X=Speck[x1,x2,x3,x4](f1,f2),X=\operatorname{\mathrm{Spec}}\frac{k[x_{1},x_{2},x_{3},x_{4}]}{(f_{1},f_{2})},

where f1,f2f_{1},f_{2} are homogeneous polynomials of degree 22. The deformation theory of XX is unobstructed, and weight dd part of the tangent space is the degree 2d2-d part of the cokernel of the 2×42\times 4 Jacobian matrix

A=(fixj)A=\left(\frac{\partial f_{i}}{\partial x_{j}}\right)

with entries in R=k[x1,x2,x3,x4]/(f1,f2)R=k[x_{1},x_{2},x_{3},x_{4}]/(f_{1},f_{2}). So the miniversal >0\mathbb{Z}_{>0}-weighted deformation space is an affine space with weights 14221^{4}2^{2} as long as intersection of the kernels of the two Hessian matrices

Hi=(2fixjxk)1j,k4,i=1,2H_{i}=\left(\frac{\partial^{2}f_{i}}{\partial x_{j}\partial x_{k}}\right)_{1\leq j,k\leq 4},\quad i=1,2

is zero.

If the characteristic of kk is not 22, then

fi(x)=12xtHix,anddfi=dxtHix.f_{i}(x)=\frac{1}{2}x^{t}H_{i}x,\quad\text{and}\quad df_{i}=dx^{t}H_{i}x.

So any nonzero vector vv in the intersection of the kernels gives a singular point in the curve E=Proj(R)E=\mathrm{Proj}(R), which is a contradiction.

If the characteristic of kk is 22, then (Hi)j,j=0(H_{i})_{j,j}=0 (so HiH_{i} is the matrix of an alternating form on k4k^{4}), and fif_{i} is of the form

fi(x)=1j<k4(Hi)j,kxjxk+j=14ai,jxj2f_{i}(x)=\sum_{1\leq j<k\leq 4}(H_{i})_{j,k}x_{j}x_{k}+\sum_{j=1}^{4}a_{i,j}x_{j}^{2}

for some vectors ai=(ai,1,ai,2,ai,3,ai,4)k4a_{i}=(a_{i,1},a_{i,2},a_{i,3},a_{i,4})\in k^{4}. The variety of all possible tuples (H1,H2,a1,a2)(H_{1},H_{2},a_{1},a_{2}) such that H1H_{1} and H2H_{2} have a common vector in their kernels is irreducible (it admits a surjection from a vector bundle over the projective space 3\mathbb{P}^{3}), and the subset of such tuples defining a smooth elliptic curve is an open subset; we will show that this subset must be empty.

Consider the open subset of tuples as above such that dimkerH1=dimkerH2=2\dim\ker H_{1}=\dim\ker H_{2}=2, dimkerH1kerH2=1\dim\ker H_{1}\cap\ker H_{2}=1, a1a_{1} and a2a_{2} are linearly independent, and some vector in the span of a1a_{1} and a2a_{2} does not lie in kerH1+kerH2\ker H_{1}+\ker H_{2}. For any such tuple, after performing an invertible linear transformation on f1f_{1} and f2f_{2}, and changing basis on k4k^{4}, we can arrange that

H1=(0000000100000100),H2=(0000000000010010),a1=(0001),a2=(abc0),H_{1}=\left(\begin{matrix}0&0&0&0\\ 0&0&0&1\\ 0&0&0&0\\ 0&1&0&0\end{matrix}\right),\quad H_{2}=\left(\begin{matrix}0&0&0&0\\ 0&0&0&0\\ 0&0&0&1\\ 0&0&1&0\end{matrix}\right),\quad a_{1}=\left(\begin{matrix}0\\ 0\\ 0\\ 1\end{matrix}\right),\quad a_{2}=\left(\begin{matrix}a\\ b\\ c\\ 0\end{matrix}\right),

for some a,b,cka,b,c\in k. A straightforward Jacobian calculation shows that Proj(R)\mathrm{Proj}(R) is singular in this case. So this nonempty open subset is disjoint from the open subset yielding smooth elliptic curves, so the latter must be empty by irreducibility, and we are done. ∎

Theorem 4.2.9.

Assume we are in the setup of Theorem 4.1.3 and, moreover, that kk does not have characteristic 22 if (G,P,μ)(G,P,\mu) is of type A1A_{1}, BB, CC, E7E_{7} or F3F_{3}, and that kk does not have characteristic 33 if (G,P,μ)(G,P,\mu) is of type E6E_{6}. Then (the formal completion of) the family χZ:ZY^//W\chi_{Z}\colon Z\to\widehat{Y}{/\mkern-6.0mu/}W is a miniversal >0(μ|)\mathbb{Z}_{>0}(\mu{\,|\,}-)-weighted deformation of χZ1(0)\chi_{Z}^{-1}(0) with respect to the action of the torus Z(L)rigZ(L)_{rig}.

Proof.

We first argue that there is no non-constant 𝔾m\mathbb{G}_{m}-orbit (equivalently, Z(L)rigZ(L)_{rig}-orbit) closure in Y^//W\widehat{Y}{/\mkern-6.0mu/}W on which the family χZ\chi_{Z} is equivariantly trivial over the completion at 0. To see this, note that every 𝔾m\mathbb{G}_{m}-orbit closure is of the form q(ΘY,y1)q(\Theta_{Y,y}^{-1}), where q:ΘY1Y^//Wq\colon\Theta_{Y}^{-1}\to\widehat{Y}{/\mkern-6.0mu/}W is the quotient map and ΘY,y1𝔸1\Theta_{Y,y}^{-1}\cong\mathbb{A}^{1} is the fibre of the line bundle ΘY1\Theta_{Y}^{-1} over a point yYy\in Y. If the pullback X𝔸1X\to\mathbb{A}^{1} of ZZ to this fibre is equivariantly trivial over the formal completion at 0, then it is trivial relative to Z0Z_{0} (since there are no deformations of the map χZ1(0)Z0\chi_{Z}^{-1}(0)\to Z_{0} of the relevant weights). Since XX is affine over Z0Z_{0}, the equivariant formal trivialisation therefore lifts uniquely to an equivariant isomorphism XχZ1(0)×𝔸1X\cong\chi_{Z}^{-1}(0)\times\mathbb{A}^{1}.

Now, there is a simultaneous log resolution π:X~=Z~yXχZ1(0)×𝔸1\pi\colon\tilde{X}=\tilde{Z}_{y}\to X\cong\chi_{Z}^{-1}(0)\times\mathbb{A}^{1} over 𝔸1\mathbb{A}^{1}, relative to the divisor 0𝔸10\in\mathbb{A}^{1}. Moreover, for any t𝔸1{t}t\in\mathbb{A}^{1}\setminus\{t\}, the map πt:X~tXt\pi_{t}\colon\tilde{X}_{t}\to X_{t} is an isomorphism over the dense open locus of points in XtX_{t} whose associated GG-bundle is regular, and has positive dimensional fibres over all other points. Since X~t\tilde{X}_{t} is smooth, it follows that Xt=χZ1(0)X_{t}=\chi_{Z}^{-1}(0) is regular in codimension 11. This contradicts Theorem 4.1.3 in the classical types AlA_{l} (l>1l>1), BB, CC and DD. In types A1A_{1}, EE, FF and GG, we instead note that Corollary 4.1.5 implies that X~t\tilde{X}_{t} has trivial canonical bundle (since Z0=SpeckZ_{0}=\operatorname{\mathrm{Spec}}k in these cases) for t0t\neq 0. In particular, from adjunction, every projective curve in X~t\tilde{X}_{t} is rational with self-intersection 2-2. But X~t\tilde{X}_{t} is a resolution of the elliptic cone χZ1(0)\chi_{Z}^{-1}(0) and is therefore birational (over χZ1(0)\chi_{Z}^{-1}(0)) to a line bundle over an elliptic curve. Since a birational map between smooth surfaces projective over a common base is a sequence of blowups at points and contractions of (1)(-1)-curves, this implies that X~t\tilde{X}_{t} must contain a projective curve whose normalisation is elliptic, which is a contradiction. So the deformation is formally nontrivial on all orbit closures as claimed.

Now let 𝔛Spf(R^)\mathfrak{X}\to\operatorname{\mathrm{Spf}}(\widehat{R}) be a miniversal >0(μ|)\mathbb{Z}_{>0}(\mu{\,|\,}-)-weighted deformation of χZ1(0)\chi_{Z}^{-1}(0). Then the completion of χZ:ZY^//W\chi_{Z}\colon Z\to\widehat{Y}{/\mkern-6.0mu/}W is the pullback of 𝔛\mathfrak{X} along some Z(L)rigZ(L)_{rig}-equivariant map Y^//WSpecR\widehat{Y}{/\mkern-6.0mu/}W\to\operatorname{\mathrm{Spec}}R such that the preimage of the origin is (set-theoretically) the fixed point. We will show that SpecR\operatorname{\mathrm{Spec}}R is an affine space with linear Z(L)rigZ(L)_{rig}-action of the same weights as Y^//W\widehat{Y}{/\mkern-6.0mu/}W, from which it follows that Y^//WSpecR\widehat{Y}{/\mkern-6.0mu/}W\to\operatorname{\mathrm{Spec}}R must be an isomorphism.

In type EE, the claim is proved in Lemma 4.2.8.

In type AlA_{l}, l>1l>1, in the notation of Lemma 4.2.7, Z(L)rig𝔾m×𝔾mZ(L)_{rig}\cong\mathbb{G}_{m}\times\mathbb{G}_{m} acts on L1L_{1} and L2L_{2} with weights (1,0)(1,0) and (0,1)(0,1) respectively and I=L11L21I=L_{1}^{-1}\otimes L_{2}^{-1} is a line bundle of degree l1-l-1. So applying Lemma 4.2.6 and Proposition 4.2.4, we have that the miniversal >02\mathbb{Z}_{>0}^{2}-weighted deformation is an affine space with weights (1,1)l+1(1,1)^{l+1}. Since (μ|)=(1,1)(\mu{\,|\,}-)=(1,1) in this presentation, this is also a miniversal >0(μ|)\mathbb{Z}_{>0}(\mu{\,|\,}-)-weighted deformation with weights (μ|)l+1(\mu{\,|\,}-)^{l+1} as required to prove.

In type BB, we identify the line bundle II on (1,2)\mathbb{P}(1,2) of Lemma 4.2.6 as follows. First, note that since line bundles on (1,2)\mathbb{P}(1,2) are rigid, we may assume without loss of generality that L=𝒪((l6)p)=π𝒪(1,2)(l6)L=\mathcal{O}((l-6)p)=\pi^{*}\mathcal{O}_{\mathbb{P}(1,2)}(l-6), where pEp\in E maps to the stacky point of (1,2)\mathbb{P}(1,2). So we have

πL1=π𝒪𝒪(6l)andπ(L2)=π𝒪𝒪(122l).\pi_{*}L^{-1}=\pi_{*}\mathcal{O}\otimes\mathcal{O}(6-l)\quad\text{and}\quad\pi_{*}(L^{-2})=\pi_{*}\mathcal{O}\otimes\mathcal{O}(12-2l).

Since π:E(1,2)\pi\colon E\to\mathbb{P}(1,2) is finite and flat of degree 22 (and the characteristic is not 22), we have π𝒪=𝒪𝒪(d)\pi_{*}\mathcal{O}=\mathcal{O}\oplus\mathcal{O}(d) for some dd\in\mathbb{Z}. Since h1((1,2),π𝒪)=h1(E,𝒪)=1h^{1}(\mathbb{P}(1,2),\pi_{*}\mathcal{O})=h^{1}(E,\mathcal{O})=1, we deduce that d=3d=-3 or 4-4. If d=4d=-4, then the μ2\mu_{2}-stabiliser of the stacky point of (1,2)\mathbb{P}(1,2) would act trivially on the fibre of π\pi, which contradicts the fact that EE is a scheme. So d=3d=-3. We deduce that

I=ker(Sym2(𝒪(6l)𝒪(3l))𝒪(122l)𝒪(92l))𝒪(62l).I=\ker(\operatorname{\mathrm{Sym}}^{2}(\mathcal{O}(6-l)\oplus\mathcal{O}(3-l))\to\mathcal{O}(12-2l)\oplus\mathcal{O}(9-2l))\cong\mathcal{O}(6-2l).

Since l3l\geq 3, 2l602l-6\geq 0, so H1((1,2),I)=0H^{1}(\mathbb{P}(1,2),I^{\vee})=0, and h1((1,2),I)=l2h^{1}(\mathbb{P}(1,2),I^{\vee})=l-2. Since Z(L)rig=𝔾mZ(L)_{rig}=\mathbb{G}_{m} acts on LL with weight 11 by Lemma 4.2.7 and hence on II^{\vee} with weight 22, we deduce from Lemma 4.2.6 and Proposition 4.2.4 that the miniversal weighted deformation space SpecR\operatorname{\mathrm{Spec}}R is an affine space with weights 132l21^{3}2^{l-2}, which are the same weights has Y^//W\widehat{Y}{/\mkern-6.0mu/}W from Table 4.

The proof in type DD is similar: comparing Euler characteristics, we see that the rank 22 bundles πL1\pi_{*}L^{-1} and πL2\pi_{*}L^{-2} on 1\mathbb{P}^{1} have degrees 6l6-l and 142l14-2l. It follows that the kernel of the surjection Sym2πL1πL2\operatorname{\mathrm{Sym}}^{2}\pi_{*}L^{-1}\to\pi_{*}L^{-2} is I=𝒪(4l)I=\mathcal{O}(4-l). Since l4l\geq 4, we have H1(1,I)=0H^{1}(\mathbb{P}^{1},I^{\vee})=0 and h0(1,I)=l3h^{0}(\mathbb{P}^{1},I^{\vee})=l-3. So the weighted miniversal deformation space SpecR\operatorname{\mathrm{Spec}}R is an affine space with weights 142l31^{4}2^{l-3}, which again agree with the weights of Y^//W\widehat{Y}{/\mkern-6.0mu/}W from Table 4.

For the remaining types, we note that in type A1A_{1} (resp., ClC_{l}, FlF_{l}, G2G_{2}), the unstable variety χZ1(0)\chi_{Z}^{-1}(0) is equivariantly isomorphic to the unstable variety for type E5E_{5} (resp., Dl+4D_{l+4}, El+4E_{l+4}, E8E_{8}), with (μ|)=2(\mu{\,|\,}-)=2 (resp., 22, 22, 33). So the miniversal >0(μ|)\mathbb{Z}_{>0}(\mu{\,|\,}-)-weighted deformation is just the μ2\mu_{2}- (resp., μ2\mu_{2}-, μ2\mu_{2}-, μ3\mu_{3}-)fixed part of the miniversal >0\mathbb{Z}_{>0}-weighted deformation. It follows from the cases proved above and inspection of Table 4 that this is an affine space with the desired weights. ∎

References