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

Level structures on parahoric torsors and complete integrability

Georgios Kydonakis and Lutian Zhao
(Date: 14 June 2025
2020 Mathematics Subject Classification: 14D15, 14H40, 14L15, 37J99, 37K10, 53D17.
Keywords: parahoric torsor, Higgs bundle, level structure, logarithmic Higgs field, moment map, Lagrangian fibration, completely integrable system)
Abstract.

For a smooth complex algebraic curve XX and a reduced effective divisor DD on XX, we introduce a notion of DD-level structure on parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsors over XX, for any connected complex reductive Lie group GG. A moduli space of parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsors equipped with a DD-level structure is constructed and we identify a canonical moment map with respect to the action of a level group on this moduli space. This action extends to a Poisson action on the cotangent, thus inducing a Poisson structure on the moduli space of logahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-Higgs torsors on XX. A study of the generic fibers of the parahoric Hitchin fibration of this moduli space identifies them as abelian torsors and introduces new algebraically completely integrable Hamiltonian Hitchin systems in this parahoric setting. We show that this framework generalizes, among other, the integrable system of Beauville and recovers the classical Gaudin model in its simplest form, the space of periodic KP elliptic solitons and the elliptic Calogero–Moser system, thus demonstrating that the logahoric Hitchin integrable system unifies many integrable systems with regular singularities under a single geometric framework.

1. Introduction

Algebraically completely integrable Hamiltonian systems on Poisson manifolds are characterized by the properties that the general level sets of the momentum map are isomorphic to an affine part of an abelian variety and the flows of the integrable vector fields are linearized by this isomorphism ([15], [34]). Many classical systems, such as the geodesic flow on an ellipsoid, the Korteweg–de Vries equation and its extensions, the elliptic Calogero–Moser system, various Euler–Arnold systems, or the Neumann system of evolution of a point on the sphere subject to a quadratic potential, can embed as symplectic leaves of certain spaces ([4], [22], [26], [30]).

Hitchin introduced in [20] an algebraically completely integrable Hamiltonian system of Poisson-commuting functions on the cotangent bundle of the moduli space of stable vector bundles on a smooth compact Riemann surface. This cotangent bundle parameterizes on the one hand Higgs bundles (vector bundles over the Riemann surface with a canonically twisted endomorphism) and on the other hand spectral data (generically line bundles on branched covers of the Riemann surface). Hitchin’s system is linearized on Jacobians of spectral curves and it is the lowest-rank symplectic leaf of a natural infinite-dimensional Poisson variety. In an extension to Hitchin’s result, Markman [26] studied families of Jacobians of spectral curves over algebraic curves of any genus obtained by twisting by any sufficiently positive line bundle. Markman proposed a geometric method via a deformation-theoretic construction, in which the Poisson structure on an open subset of the algebraically completely integrable Hamiltonian system is obtained via symplectic reduction from the cotangent of the moduli space 𝒰Σ(r,d,D)\mathcal{U}_{\Sigma}(r,d,D) of stable rank rr degree dd vector bundles EE over a Riemann surface Σ\Sigma equipped with a DD-level structure, namely, an isomorphism E|Di=1r𝒪DE|_{D}\to\bigoplus_{i=1}^{r}\mathcal{O}_{D}, for a fixed reduced effective divisor DD on Σ\Sigma. The concept of a DD-level structure was introduced earlier by Seshadri in [32], who also constructed a moduli space of stable vector bundles equipped with such structure as a smooth quasi-projective variety. The same result to [26] for the moduli space of twisted stable Higgs pairs was obtained independently by Bottacin in [9], who produced an explicit antisymmetric contravariant 2-tensor at the stable points. Using explicit computations, he proved that this defines a Poisson structure and checked the linearity of the flows.

The aim in this article is to introduce algebraically completely integrable Hamiltonian Hitchin systems for general complex reductive structure groups GG, as a way to provide a mathematical framework for describing the geometry of certain integrable systems in Physics with regular singularities. We do this by generalizing Markman’s approach, that is, by studying the symplectic leaves of the GG-Hitchin system over a complex algebraic curve with a divisor of finitely many distinct points. Various similar treatments have appeared in the literature with limitations on the parabolic structure and the framing imposed locally on the principal bundles (see, for instance, [3], [7], [25], [37]).

In this article, we use the language of parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsors and introduce a new notion of DD-level structure that both generalizes Seshadri’s concept and interacts coherently with the parahoric Bruhat–Tits data. This provides a more transparent framework for studying the Hitchin system and its symplectic leaves for general reductive groups. We next describe our results in more detail.

Let XX be a smooth complex projective curve, DD a reduced effective divisor on XX, and GG a connected complex reductive Lie group. For each point xDx\in D, fix a collection of weights 𝜽:={θx}xD\boldsymbol{\theta}:=\{\theta_{x}\}_{x\in D}, that is, a collection of points in the Bruhat–Tits apartment of G(((t)))G\left(\mathbb{C}((t))\right); each of these points determines a parahoric subgroup scheme 𝒢θx\mathcal{G}_{\theta_{x}} on the formal disk 𝔻x:=Spec𝒪^X,x\mathbb{D}_{x}:=\mathrm{Spec}\widehat{\mathcal{O}}_{X,x}. We introduce the following:

Definition 1.1 (Definition 3.1).

A DD-level structure of parahoric type (θx)xD(\theta_{x})_{x\in D} on a parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor \mathcal{E} over XX is a choice of section ηx:𝔻xGθx/Gθx+\eta_{x}:\mathbb{D}_{x}\to G_{\theta_{x}}/G_{\theta_{x}}^{+}, for each xDx\in D, where Gθx+G_{\theta_{x}}^{+} denotes the pro-unipotent radical of GθxG_{\theta_{x}}. We will denote a parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor equipped with a DD-level structure as a pair (,η)(\mathcal{E},\eta).

Note that the data of a DD-level structure refines the parahoric torsor by imposing a “flag-type” or “framing-type” condition at each point in DD. We call a DD-level structure on a parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor stable (resp. semistable) if the underlying parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor is RR-stable (resp. RR-semistable) in the sense introduced in [24]. A moduli space 𝒰(X,𝒢𝜽)\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}}) of stable pairs (,η)(\mathcal{E},\eta) of parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsors over XX with a DD-level structure is then constructed as an irreducible normal projective variety:

Theorem 1.2 (Theorem 3.7).

Let XX be a smooth complex projective curve and DXD\subset X a reduced effective divisor. Let 𝒢𝛉\mathcal{G}_{\boldsymbol{\theta}} be the parahoric Bruhat–Tits group scheme on XX corresponding to a collection 𝛉:={θx}xD\boldsymbol{\theta}:=\{\theta_{x}\}_{x\in D} of rational weights. Then the moduli functor which assigns to any scheme SS the set of SS-equivalent classes of semistable parahoric 𝒢𝛉\mathcal{G}_{\boldsymbol{\theta}}-torsors on XX with fixed DD-level structure is corepresented by an irreducible, normal, projective variety.

We analyze the deformations of the DD-level structure to show that the tangent space of the moduli space 𝒰(X,𝒢𝜽)\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}}) corresponds to the space of logarithmic Higgs fields H0(X,(𝔤)K(D))H^{0}(X,\mathcal{E}(\mathfrak{g})\otimes K(D)) (Proposition 4.1).

In order to identify a canonical moment map on the cotangent T𝒰(X,𝒢𝜽)T^{*}\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}}), we consider the action of a certain level group GDG_{D} on the moduli space 𝒰(X,𝒢𝜽)\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}}) (Definition 5.11). This action is free on the regularly stable locus and, in fact, extends to a Poisson action on the cotangent T𝒰(X,𝒢𝜽)T^{*}\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}}) (see Sections 2 and 5 for further notational explanations):

Theorem 1.3 (Theorem 5.16).

The group GDG_{D} acts Poisson on T𝒰(X,𝒢𝛉)T^{*}\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}}). Moreover, the canonical moment map

μ:T𝒰(X,𝒢𝜽)𝔤D\mu\colon T^{*}\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}})\;\longrightarrow\;\mathfrak{g}_{D}^{*}

is given by dualizing the infinitesimal action and can be explicitly described via coresidues at the divisor DD. Its image is the element of 𝔤D=j=1s𝔩^θj\mathfrak{g}_{D}^{*}=\bigoplus_{j=1}^{s}\hat{\mathfrak{l}}_{\theta_{j}}^{*} given by the direct sum of the coresidues at each point xjDx_{j}\in D:

μ([(,η)],φ)=j=1sCoResxj(φ).\mu([(\mathcal{E},\eta)],\varphi)=\bigoplus_{j=1}^{s}\operatorname{CoRes}_{x_{j}}(\varphi).

Explicitly, for any element Y=(Y1,,Ys)𝔤D=j=1s𝔩^θjY=(Y_{1},\dots,Y_{s})\in\mathfrak{g}_{D}=\bigoplus_{j=1}^{s}\hat{\mathfrak{l}}_{\theta_{j}}, the pairing is given by the sum of Killing form pairings over the divisor DD:

μ([(,η)],φ),Y=j=1s(Resxj(φ),Yj).\langle\mu([(\mathcal{E},\eta)],\varphi),Y\rangle=\sum_{j=1}^{s}(\operatorname{Res}_{x_{j}}(\varphi),Y_{j}).

Similarly to Markman’s original work, we deduce a Poisson structure on the moduli space H(X,𝒢𝜽)\mathcal{M}_{H}(X,\mathcal{G}_{\boldsymbol{\theta}}) of logahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-Higgs torsors via a forgetful morphism

l:LH(X,𝒢𝜽)H(X,𝒢𝜽),l:\mathcal{M}_{LH}(X,\mathcal{G}_{\boldsymbol{\theta}})\to\mathcal{M}_{H}(X,\mathcal{G}_{\boldsymbol{\theta}}),

from the coarse moduli space LH(X,𝒢𝜽)\mathcal{M}_{LH}(X,\mathcal{G}_{\boldsymbol{\theta}}) of triples (,φ,η)(\mathcal{E},\varphi,\eta) of logahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-Higgs torsors with DD-level structure to the moduli space of pairs (,φ)(\mathcal{E},\varphi). The cotangent T𝒰(X,𝒢𝜽)T^{*}\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}}) of the moduli space 𝒰(X,𝒢𝜽)\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}}) is an open subset of MLH(X,𝒢𝜽)M_{LH}(X,\mathcal{G}_{\boldsymbol{\theta}}) and the forgetful map ll is Poisson, thus inheriting with a Poisson structure the space H(X,𝒢𝜽)\mathcal{M}_{H}(X,\mathcal{G}_{\boldsymbol{\theta}}) (Theorem 5.20).

We then use the cameral covers of Donagi [14] to study the parahoric Hitchin fibration on the space H(X,𝒢𝜽)\mathcal{M}_{H}(X,\mathcal{G}_{\boldsymbol{\theta}}). We prove that its generic fibers are Lagrangian subvarieties with respect to the symplectic leaves of the Poisson moduli space H(X,𝒢𝜽)\mathcal{M}_{H}(X,\mathcal{G}_{\boldsymbol{\theta}}), thus concluding to our main result:

Theorem 1.4 (Theorem 6.17).

Let XX be a smooth complex algebraic curve and DD be a reduced effective divisor on XX. Let GG be a connected complex reductive group. The moduli space H(X,𝒢𝛉)\mathcal{M}_{H}(X,\mathcal{G}_{\boldsymbol{\theta}}) of logahoric 𝒢𝛉\mathcal{G}_{\boldsymbol{\theta}}-Higgs torsors over XX is Poisson and is fibered via a map h𝛉:H(X,𝒢𝛉)𝒜𝛉h_{\boldsymbol{\theta}}:\mathcal{M}_{H}(X,\mathcal{G}_{\boldsymbol{\theta}})\to\mathcal{A}_{\boldsymbol{\theta}} by abelian torsors. Moreover, h𝛉:H(X,𝒢𝛉)𝒜𝛉h_{\boldsymbol{\theta}}:\mathcal{M}_{H}(X,\mathcal{G}_{\boldsymbol{\theta}})\to\mathcal{A}_{\boldsymbol{\theta}} is an algebraically completely integrable Hamiltonian system.

We will call the algebraically completely integrable Hamiltonian system of Theorem 1.4, the logahoric Hitchin integrable system. Considering the image 𝒜𝜽+\mathcal{A}_{\boldsymbol{\theta}}^{+} under h𝜽h_{\boldsymbol{\theta}} of the strongly logahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-Higgs torsors, we show (Theorem 7.5) that there exists an isomorphism of affine varieties

𝒜𝜽/𝒜𝜽+𝔤D//GD,\mathcal{A}_{\boldsymbol{\theta}}/\mathcal{A}_{\boldsymbol{\theta}}^{+}\cong\mathfrak{g}_{D}^{*}//G_{D},

for the Hitchin fibration H~𝜽:LH(X,𝒢𝜽)𝒜𝜽{\widetilde{H}_{\boldsymbol{\theta}}}\colon\mathcal{M}_{LH}(X,\mathcal{G}_{\boldsymbol{\theta}})\to\mathcal{A}_{\boldsymbol{\theta}} on the space of leveled logahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-Higgs torsors. This allows us to describe the symplectic leaf foliation explicitly:

Theorem 1.5 (Corollary 7.6).

The foliation of the smooth locus H(X,𝒢𝛉)sm\mathcal{M}_{H}(X,\mathcal{G}_{\boldsymbol{\theta}})^{\mathrm{sm}} by its symplectic leaves refines the foliation by fibers of the map

qH~𝜽:H(X,𝒢𝜽)sm𝒜𝜽/𝒜𝜽+,q\circ\widetilde{H}_{\boldsymbol{\theta}}\;:\;\mathcal{M}_{H}(X,\mathcal{G}_{\boldsymbol{\theta}})^{\mathrm{sm}}\;\longrightarrow\;\mathcal{A}_{\boldsymbol{\theta}}/\mathcal{A}_{\boldsymbol{\theta}}^{+},

for the quotient map q:𝒜𝛉𝒜𝛉/𝒜𝛉+q:\mathcal{A}_{\boldsymbol{\theta}}\to\mathcal{A}_{\boldsymbol{\theta}}/\mathcal{A}_{\boldsymbol{\theta}}^{+}. Moreover, each fiber of qH~𝛉q\circ\widetilde{H}_{\boldsymbol{\theta}} contains a unique symplectic leaf of maximal dimension.

In the last part of the article, we demonstrate examples of algebraically completely integrable Hamiltonian systems that can be analyzed as logahoric Hitchin integrable systems. Namely, apart from the integrable systems studied, for instance, in [4], [26], [30] that can be studied using this approach in the absence of a parahoric structure, we show that over the Riemann sphere X=1X=\mathbb{P}^{1} the logahoric Hitchin integrable system generalizes the integrable system of Beauville [4]. Furthermore, still for X=1X=\mathbb{P}^{1} one can recover the classical Gaudin model for a specific choice of parahoric data corresponding to the simplest pole structure. For an elliptic base curve X=ΣX=\Sigma, and for group G=SL(r,)G=\mathrm{SL}(r,\mathbb{C}), we recover the KP hierarchy when taking the subvariety of LH(Σ,𝒢𝜽)\mathcal{M}_{LH}(\Sigma,\mathcal{G}_{\boldsymbol{\theta}}) consisting of triples (,η,φ)(\mathcal{E},\eta,\varphi), where the coresidue of the Higgs field is the coadjoint orbit of the element in 𝔰𝔩(r)\mathfrak{sl}(r)^{*} corresponding (via the Killing form) to the matrix diag(1,1,,1,r1)\mathrm{diag}(-1,-1,\dots,-1,r-1). Another example in this elliptic case that can be studied using the logahoric Hitchin integrable system is the elliptic Calogero–Moser system as demonstrated by Hurtubise and Markman in [22]. We note, lastly, that the framework built here, enables the study of the geometry of integrable systems for higher genus and/or non-trivial orbits than the ones exhibited here.

2. Parahoric Torsors

Let GG be a connected complex reductive Lie group, and fix a maximal torus TT in GG with corresponding Lie algebras 𝔱\mathfrak{t} and 𝔤\mathfrak{g}. We consider the character group X(T):=Hom(T,𝔾m)X(T):=\text{Hom}(T,\mathbb{G}_{m}) and the co-character group Y(T):=Hom(𝔾m,T)Y(T):=\text{Hom}(\mathbb{G}_{m},T), which is also the group of one-parameter subgroups of TT. The canonical pairing ,:Y(T)×X(T)\langle\cdot,\cdot\rangle:Y(T)\times X(T)\rightarrow\mathbb{Z} can be extended to \mathbb{R} by tensoring Y(T)Y(T) and X(T)X(T) with \mathbb{R}. We refer to co-characters with coefficients in \mathbb{R} and \mathbb{Q} as real weights and rational weights, respectively. In general, by a weight we will always consider a real weight unless stated otherwise.

Denote the root system with respect to the maximal torus TT as \mathcal{R}, and let +\mathcal{R}_{+}\subseteq\mathcal{R} be the set of positive roots. For a root rr\in\mathcal{R}, we have an isomorphism of Lie algebras Lie(𝔾a)(Lie(G))r{\rm Lie}(\mathbb{G}_{a})\rightarrow({\rm Lie}(G))_{r}, which induces a natural homomorphism of groups ur:𝔾aGu_{r}:\mathbb{G}_{a}\rightarrow G such that tur(a)t1=ur(r(t)a)tu_{r}(a)t^{-1}=u_{r}(r(t)a), for tTt\in T and a𝔾aa\in\mathbb{G}_{a}. We denote the image of the homomorphism uru_{r} by UrU_{r}, which is a closed subgroup of GG. The reductive group GG is generated by its subgroups TT and UrU_{r}, for rr\in\mathcal{R}. Namely, an element gg in GG can be written as a product g=gtrgrg=g_{t}\prod_{r\in\mathcal{R}}g_{r}, where gtTg_{t}\in T and grUrg_{r}\in U_{r}. Sometimes, we write gg as a tuple (gt,gr)r(g_{t},g_{r})_{r\in\mathcal{R}} for convenience.

2.1. Parahoric subgroups

Given a weight θ\theta, we can consider it as an element in 𝔱\mathfrak{t}, the Lie algebra of TT, under differentiation. We define the integer mr(θ):=r(θ)m_{r}(\theta):=\lceil-r(\theta)\rceil, where \lceil\cdot\rceil is the ceiling function and r(θ):=θ,rr(\theta):=\langle\theta,r\rangle. We introduce the following definition:

Definition 2.1.

Let R:=[[z]]R:=\mathbb{C}[[z]] and K:=((z))K:=\mathbb{C}((z)). With respect to the above data, we define the parahoric subgroup Gθ(K)G_{\theta}(K) of G(K)G(K) as

Gθ(K):=T(R),Ur(zmr(θ)R),r.\displaystyle G_{\theta}(K):=\langle T(R),U_{r}(z^{m_{r}(\theta)}R),r\in\mathcal{R}\rangle.

Denote by 𝒢θ\mathcal{G}_{\theta} the corresponding group scheme of Gθ(K)G_{\theta}(K), which is called the parahoric group scheme.

The parahoric subgroup of G(K)G(K) determined by θ\theta can be alternatively defined as

Gθ(K):={g(z)G(K) | zθg(z)zθ has a limit as z0},\displaystyle G^{\prime}_{\theta}(K):=\{g(z)\in G(K)\text{ }|\text{ }z^{\theta}g(z)z^{-\theta}\text{ has a limit as $z\rightarrow 0$}\},

where zθ:=eθlnzz^{\theta}:=e^{\theta\ln z}. This definition is of a more analytic nature; the equivalence of these two definitions can be found in our previous paper [21, Lemma 2.2]. We will thus utilize either definition of parahoric subgroups, Gθ(K)G_{\theta}(K) or Gθ(K)G^{\prime}_{\theta}(K), in the study of connected complex reductive groups and their representations.

2.2. Parahoric Torsors

For GG as introduced earlier, let XX be a smooth projective curve over \mathbb{C}, and let

D={x1,,xs}XD=\{x_{1},\dots,x_{s}\}\subset X

be a reduced effective divisor. Denote the complement by XD:=XDX_{D}:=X\setminus D. For every xDx\in D, fix a weight θx\theta_{x}, namely, a point of the Bruhat–Tits apartment of G(((t)))G(\mathbb{C}((t))). This point determines a facet of the Bruhat–Tits apartment and hence a parahoric subgroup scheme GθxG_{\theta_{x}} on the formal disc 𝔻x:=Spec𝒪X,x^\mathbb{D}_{x}:=\operatorname{Spec}\widehat{\mathcal{O}_{X,x}}. θx\theta_{x}. Let us also write 𝜽:={θx}xD\boldsymbol{\theta}:=\{\theta_{x}\}_{x\in D}, for a collection of weights for each point in DD.

Following [2, Section 2] we now glue together

𝒢𝜽|XD=G×XD,𝒢𝜽|𝔻x=Gθx(xD)\mathcal{G}_{\boldsymbol{\theta}}\bigl{|}_{X_{D}}=G\times X_{D},\qquad\mathcal{G}_{\boldsymbol{\theta}}\bigl{|}_{\mathbb{D}_{x}}=G_{\theta_{x}}\;(x\in D)

to obtain the parahoric Bruhat–Tits group scheme

𝒢𝜽X.\mathcal{G}_{\boldsymbol{\theta}}\longrightarrow X.

This scheme is smooth, affine, of finite type, and flat over XX [10, Lemma 3.18].

Definition 2.2 (Generically split torsor [2, Section 3]).

A 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor X\mathcal{E}\to X is called generically split if its restriction to the generic point q=Spec(X)q=\operatorname{Spec}\mathbb{C}(X) becomes trivial:

|q𝒢𝜽|qG×q.\mathcal{E}\bigl{|}_{q}\;\cong\;\mathcal{G}_{\boldsymbol{\theta}}\bigl{|}_{q}\;\simeq\;G\times q.
Remark 2.3.

Except when we explicitly state otherwise (see Example 2.7), we shall abuse terminology throughout the article and call a generically split 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor simply a parahoric 𝒢𝛉\mathcal{G}_{\boldsymbol{\theta}}-torsor.

Local description and gluing.

Fix a parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor \mathcal{E} on XDX_{D} and, for every xDx\in D, fix a Gθx(K)G_{\theta_{x}}(K)-torsor x\mathcal{E}_{x} on the punctured disc 𝔻x×:=SpecK\mathbb{D}_{x}^{\times}:=\operatorname{Spec}K, where K=((t))K=\mathbb{C}((t)). A parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor \mathcal{E} on XX is given by gluing \mathcal{E} and the {x}\{\mathcal{E}_{x}\} through isomorphisms

Θx:x|𝔻x×|𝔻x×,\Theta_{x}:\mathcal{E}_{x}\bigl{|}_{\mathbb{D}_{x}^{\times}}\;\longrightarrow\;\mathcal{E}\bigl{|}_{\mathbb{D}_{x}^{\times}}, (2.1)

one for each xDx\in D. Each Θx\Theta_{x} corresponds to a loop gx(t)G(K)g_{x}(t)\in G(K).

Two sets of data (,{x},{Θx})(\mathcal{E},\{\mathcal{E}_{x}\},\{\Theta_{x}\}) and (,{x},{Θx})(\mathcal{E}^{\prime},\{\mathcal{E}_{x}^{\prime}\},\{\Theta_{x}^{\prime}\}) determine isomorphic parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsors provided there are isomorphisms for every xDx\in D

ψ: and ψx:xx,\psi:\mathcal{E}\rightarrow\mathcal{E}^{\prime}\text{ and }\psi_{x}:\mathcal{E}_{x}\rightarrow\mathcal{E}_{x}^{\prime},

such that ψΘx=Θxψx\psi\circ\Theta_{x}=\Theta_{x}^{\prime}\circ\psi_{x} for every xDx\in D:

x{\mathcal{E}_{x}}x{\mathcal{E}_{x}^{\prime}}{\mathcal{E}}.{\mathcal{E}^{\prime}.}Θx\scriptstyle{\Theta_{x}}ψx\scriptstyle{\psi_{x}}Θx\scriptstyle{\Theta_{x}^{\prime}}ψ\scriptstyle{\psi}
Remark 2.4 (Beauville–Laszlo gluing for generically split torsors [5, 2]).

The Beauville–Laszlo gluing construction provides a method to construct global torsors on the curve XX. In the specific setting described above, consider the case of generically split torsors. If we begin with the trivial 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor on XD=XDX_{D}=X\setminus D and the trivial Gθx(K)G_{\theta_{x}}(K)-torsor on each punctured disc 𝔻x×\mathbb{D}_{x}^{\times} (for xDx\in D), then a given collection of gluing isomorphisms Θx\Theta_{x} (which correspond to loops gx(t)G(K)g_{x}(t)\in G(K)) defines a parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor \mathcal{E} on XX. The existence and uniqueness of this resulting torsor \mathcal{E} are ensured by the general principles of torsor gluing, as detailed in [2, Remark 5.2.3]. This result for torsors is a consequence of the more general Beauville–Laszlo framework [5], and is analogous to the gluing of group schemes themselves as described in [2, Lemma 5.2.2]. Thus, for generically split data, the gluing process is well-defined and yields a unique torsor on XX.

Remark 2.5 (Uniformization for generically split torsors).

Continuing from Remark 2.4, the isomorphism classes of generically split parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsors, constructed via the Beauville–Laszlo gluing of trivial torsors, can be classified using the uniformization theorem of Heinloth [19]. A generically split 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor X\mathcal{E}\to X (as per Definition 2.2 and Remark 2.3) is trivial at the generic point q=Spec(X)q=\operatorname{Spec}\mathbb{C}(X), i.e., |qG×q\mathcal{E}|_{q}\cong G\times q. Locally, \mathcal{E} is described via the Beauville–Laszlo gluing construction: it is obtained by gluing a GG-torsor |XD\mathcal{E}|_{X_{D}} on XDX_{D} with 𝒢θx\mathcal{G}_{\theta_{x}}-torsors x\mathcal{E}_{x} on each 𝔻x\mathbb{D}_{x}, using isomorphisms

Θx:x|𝔻x×|𝔻x×,\Theta_{x}:\mathcal{E}_{x}|_{\mathbb{D}_{x}^{\times}}\to\mathcal{E}|_{\mathbb{D}_{x}^{\times}},

where 𝔻x×=Spec((t))\mathbb{D}_{x}^{\times}=\operatorname{Spec}\mathbb{C}((t)), and each Θx\Theta_{x} corresponds to a gluing element gx(t)G(K)g_{x}(t)\in G(K). Since \mathcal{E} is generically split, we may assume (after choosing a trivialization) that |XDG×XD\mathcal{E}|_{X_{D}}\cong G\times X_{D}, making |𝔻x×G×𝔻x×\mathcal{E}|_{\mathbb{D}_{x}^{\times}}\cong G\times\mathbb{D}_{x}^{\times}, so gx(t)g_{x}(t) defines the transition from the trivial torsor on 𝔻x\mathbb{D}_{x}.

Two sets of gluing data g={gxG(K)}xDg_{\bullet}=\{g_{x}\in G(K)\}_{x\in D} and g={gxG(K)}xDg_{\bullet}^{\prime}=\{g_{x}^{\prime}\in G(K)\}_{x\in D} define isomorphic 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsors if there exist isomorphisms ψ:|XD|XD\psi:\mathcal{E}|_{X_{D}}\to\mathcal{E}^{\prime}|_{X_{D}} and ψx:xx\psi_{x}:\mathcal{E}_{x}\to\mathcal{E}_{x}^{\prime} satisfying ψΘx=Θxψx\psi\circ\Theta_{x}=\Theta_{x}^{\prime}\circ\psi_{x}, for all xDx\in D. In the trivial case, ψ\psi corresponds to an element h𝒢𝜽(XD)=G(XD)h\in\mathcal{G}_{\boldsymbol{\theta}}(X_{D})=G(X_{D}), and each ψx\psi_{x} corresponds to kx𝒢θx(𝔻x)=Gθx(K)k_{x}\in\mathcal{G}_{\theta_{x}}(\mathbb{D}_{x})=G_{\theta_{x}}(K). The compatibility condition translates to:

gx=h|𝔻x×gxkx,for all xD,g_{x}^{\prime}=h|_{\mathbb{D}_{x}^{\times}}\,g_{x}\,k_{x},\quad\text{for all }x\in D,

where h|𝔻x×h|_{\mathbb{D}_{x}^{\times}} is the restriction of hG(XD)h\in G(X_{D}) to 𝔻x×\mathbb{D}_{x}^{\times}, an element of G(K)G(K), and kxGθx(K)k_{x}\in G_{\theta_{x}}(K) acts on the right.

Thus, the isomorphism classes of generically split 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsors are parameterized by the quotient

Bun𝒢𝜽G(XD)\xD(G(K)/Gθx(K)),\mathrm{Bun}_{\mathcal{G}_{\boldsymbol{\theta}}}\cong G(X_{D})\backslash\prod_{x\in D}\left(G(K)/G_{\theta_{x}}(K)\right),

where:

  • G(K)/Gθx(K)G(K)/G_{\theta_{x}}(K) is the affine Grassmannian (or partial affine flag variety) at xx, denoted by Gr𝒢,x\mathrm{Gr}_{\mathcal{G},x},

  • G(XD)G(X_{D}) acts on the left via the natural maps G(XD)G(K)G(X_{D})\to G(K), for each xDx\in D,

  • The product xDGr𝒢,x\prod_{x\in D}\mathrm{Gr}_{\mathcal{G},x} is the multi-point affine Grassmannian Gr𝒢,D\mathrm{Gr}_{\mathcal{G},D}.

This classification is a direct application of the uniformization theorem for the stack Bun𝒢𝜽\mathrm{Bun}_{\mathcal{G}_{\boldsymbol{\theta}}} of 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsors by Heinloth [19] and is explicitly stated by Damiolini and Hong in [12, Section 2.1].

Remark 2.6 (Beyond the generically split case).

For parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsors that are not necessarily generically split, Damiolini and Hong in [11, Theorem 6.2.2] present a construction for general parahoric Bruhat–Tits group schemes, which also applies to their torsors, accommodating non-generically split cases. They establish that any such parahoric Bruhat–Tits group scheme can be realized from a (Γ,G)(\Gamma,G)-bundle associated with a finite, tamely ramified cover X~X\widetilde{X}\to X. Within this framework, the gluing procedure and the classification of isomorphism classes are addressed through a Γ\Gamma-equivariant adaptation of the uniformization principle. The structure of these torsors, especially their local aspects at points xDx\in D (which may correspond to ramified points in the cover), is then classified using non-abelian group cohomology, typically involving classes in H1(Γx,G(Kx))H^{1}(\Gamma_{x},G(K_{x})), where Γx\Gamma_{x} represents the local Galois group (or a relevant inertia subgroup) at xx.

Example 2.7.

We now list a few fundamental examples of parahoric 𝒢θ\mathcal{G}_{\theta}-torsors:

  1. (1)

    Parabolic principal GG-bundles.

    For each xDx\in D, let αx𝔱\alpha_{x}\in\mathfrak{t} be a parabolic weight defining a parabolic subgroup PαxGP_{\alpha_{x}}\subseteq G containing a Borel subgroup. Set the parahoric weight θx=αx\theta_{x}=\alpha_{x}. The parahoric subgroup scheme GθxG_{\theta_{x}} over 𝔻x=Spec[[tx]]\mathbb{D}_{x}=\operatorname{Spec}\mathbb{C}[[t_{x}]] is defined as Gθx=ev01(Pαx)G_{\theta_{x}}=\mathrm{ev}_{0}^{-1}(P_{\alpha_{x}}), where ev0:G([[tx]])G()\mathrm{ev}_{0}:G(\mathbb{C}[[t_{x}]])\to G(\mathbb{C}) maps g(tx)g(0)g(t_{x})\mapsto g(0). Thus, sections of GθxG_{\theta_{x}} are maps g:𝔻xGg:\mathbb{D}_{x}\to G with g(0)Pαxg(0)\in P_{\alpha_{x}}.

    A 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor X\mathcal{E}\to X is a principal GG-bundle over XDX\setminus D and a GθxG_{\theta_{x}}-torsor over each 𝔻x\mathbb{D}_{x}, locally isomorphic to Gθx×𝔻xG_{\theta_{x}}\times\mathbb{D}_{x}. A generically split 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor corresponds to a principal GG-bundle 𝒫\mathcal{P} over XX such that, in any local trivialization over 𝔻x\mathbb{D}_{x}, sections s:𝔻x𝒫s:\mathbb{D}_{x}\to\mathcal{P} correspond to g(tx)G([[tx]])g(t_{x})\in G(\mathbb{C}[[t_{x}]]) with g(0)Pαxg(0)\in P_{\alpha_{x}}. This defines a parabolic structure at each xDx\in D, recovering the classical notion of parabolic GG-bundles.

  2. (2)

    Parabolic vector bundles.

    For G=GLnG=\mathrm{GL}_{n}, a parabolic vector bundle of rank nn on XX is a vector bundle VXV\to X equipped with a flag of subspaces in the fiber VxV_{x} for each xDx\in D:

    Vx=Fx1Fx2Fxk{0},V_{x}=F^{1}_{x}\supset F^{2}_{x}\supset\dots\supset F^{k}_{x}\supset\{0\},

    defining a parabolic subgroup PxGLnP_{x}\subset\mathrm{GL}_{n} as the stabilizer of the flag. Let αx𝔱\alpha_{x}\in\mathfrak{t} (the Lie algebra of diagonal matrices) be a parabolic weight corresponding to PxP_{x}.

    The frame bundle Fr(V)\mathrm{Fr}(V) is a principal GLn\mathrm{GL}_{n}-bundle over XX. The flags induce parabolic structures on Fr(V)\mathrm{Fr}(V) with parabolic subgroups PxP_{x}. Setting θx=αx\theta_{x}=\alpha_{x}, then Fr(V)\mathrm{Fr}(V) becomes a generically split parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor, where, over 𝔻x\mathbb{D}_{x}, the sections satisfy the flag condition via PxP_{x}.

2.3. Parahoric Lie Algebra

Given a weight θ\theta, then associated to Gθ(K)G_{\theta}(K), we have the parahoric Lie algebra

𝔤θ(K)=𝔱(R)rzmr(θ)𝔤r(R),\mathfrak{g}_{\theta}(K)=\mathfrak{t}(R)\oplus\bigoplus_{r\in\mathcal{R}}z^{m_{r}(\theta)}\mathfrak{g}_{r}(R),

with the Killing form providing the identification

𝔤θ(K)=𝔱(R)rzmr(θ)𝔤r(R).\mathfrak{g}_{\theta}(K)^{\perp}=\mathfrak{t}(R)\oplus\bigoplus_{r\in\mathcal{R}}z^{-m_{r}(\theta)}\mathfrak{g}_{-r}(R).

Recall also from [27, Section 2.6]) that each parahoric subgroup Gθ(K)G(K)G_{\theta}(K)\subset G(K) admits an exact sequence of group schemes. There exists an exact sequence

1Gθ+(K)Gθ(K)Lθ1,1\to G_{\theta}^{+}(K)\to G_{\theta}(K)\to L_{\theta}\to 1,

where Gθ+(K)G_{\theta}^{+}(K) is the pro-unipotent radical of Gθ(K)G_{\theta}(K) and LθL_{\theta} is the reductive Levi quotient. Since Gθ+(K)G_{\theta}^{+}(K) is a normal subgroup, this sequence defines the group LθL_{\theta} abstractly as the quotient Gθ(K)/Gθ+(K)G_{\theta}(K)/G_{\theta}^{+}(K). The Lie algebra of the pro-unipotent radical, 𝔤θ+=Lie(Gθ+(K))\mathfrak{g}_{\theta}^{+}=\text{Lie}(G_{\theta}^{+}(K)), is an ideal in 𝔤θ(K)\mathfrak{g}_{\theta}(K) given by

𝔤θ+=z𝔱(R)rzmr(θ)+1𝔤r(R).\mathfrak{g}_{\theta}^{+}=z\mathfrak{t}(R)\oplus\bigoplus_{r\in\mathcal{R}}z^{\lceil-m_{r}(\theta)\rceil+1}\mathfrak{g}_{r}(R). (2.1)

The Lie algebra of the Levi quotient, which we will denote by 𝔩^θ\hat{\mathfrak{l}}_{\theta}, is the quotient of the corresponding Lie algebras

𝔩^θ:=𝔤θ/𝔤θ+.\hat{\mathfrak{l}}_{\theta}:=\mathfrak{g}_{\theta}/\mathfrak{g}_{\theta}^{+}.

This is a reductive Lie algebra over the residue field kk. The Levi decomposition theorem states that this abstract quotient can be realized as a concrete subgroup. Let LθL_{\theta} be the finite-dimensional reductive group over the residue field kk whose root data corresponds to the integer eigenspaces of adθ\text{ad}_{\theta}:

Lθ=CG(e2πiθ)=T(k×),Ur(k)r,mr(θ).L_{\theta}=C_{G}(e^{2\pi i\theta})=\langle T(k^{\times}),U_{r}(k)\mid r\in\mathcal{R},m_{r}(\theta)\in\mathbb{Z}\rangle. (2.2)

The concrete Levi subgroup L^θG(K)\hat{L}_{\theta}\subset G(K) is obtained by conjugating LθL_{\theta}:

L^θ=zθLθzθ=T(k×),Ur(zmr(θ)k)r,mr(θ).\hat{L}_{\theta}=z^{-\theta}L_{\theta}z^{\theta}=\langle T(k^{\times}),U_{r}(z^{-m_{r}(\theta)}k)\mid r\in\mathcal{R},m_{r}(\theta)\in\mathbb{Z}\rangle.

The corresponding Lie algebra 𝔩^θ=Lie(L^θ)\hat{\mathfrak{l}}_{\theta}=\text{Lie}(\hat{L}_{\theta}) is a subalgebra of the loop algebra 𝔤(K)\mathfrak{g}(K) given by

𝔩^θ=Ad(zθ)(Lie(Lθ))=𝔱(k)r,mr(θ)zmr(θ)𝔤r(k).\hat{\mathfrak{l}}_{\theta}=\text{Ad}(z^{-\theta})(\text{Lie}(L_{\theta}))=\mathfrak{t}(k)\oplus\bigoplus_{r\in\mathcal{R},\,m_{r}(\theta)\in\mathbb{Z}}z^{-m_{r}(\theta)}\mathfrak{g}_{r}(k).

The Lie algebra 𝔤θ(K)\mathfrak{g}_{\theta}(K) admits a decomposition (the Levi decomposition) as a semidirect product:

𝔤θ(K)=𝔩^θ𝔤θ+.\mathfrak{g}_{\theta}(K)=\hat{\mathfrak{l}}_{\theta}\ltimes\mathfrak{g}_{\theta}^{+}.

The Lie algebra 𝔩^θ\hat{\mathfrak{l}}_{\theta} is isomorphic to the finite-dimensional reductive Lie algebra 𝔩θ=Lie(Lθ)\mathfrak{l}_{\theta}=\text{Lie}(L_{\theta}) (cf. [8, Section 2.2]), which is the subalgebra of 𝔤(k)\mathfrak{g}(k):

𝔩θ=𝔱(k)r,mr(θ)𝔤r(k).\mathfrak{l}_{\theta}=\mathfrak{t}(k)\oplus\bigoplus_{r\in\mathcal{R},\,m_{r}(\theta)\in\mathbb{Z}}\mathfrak{g}_{r}(k).

The isomorphism is given by the evaluation map ι:𝔩^θ𝔩θ\iota:\hat{\mathfrak{l}}_{\theta}\to\mathfrak{l}_{\theta} defined by setting z=1z=1 after conjugating back by zθz^{\theta}:

ι(X(z))=Ad(zθ)(X(z))|z=1.\iota(X(z))=\left.\text{Ad}(z^{\theta})(X(z))\right|_{z=1}.

We thus have canonical isomorphisms:

LθL^θ,𝔩θ𝔩^θ.L_{\theta}\cong\hat{L}_{\theta},\quad\mathfrak{l}_{\theta}\cong\hat{\mathfrak{l}}_{\theta}.

This allows us to treat the abstract Levi quotient as the concrete and familiar reductive Lie group LθL_{\theta} of the group GG, and its Lie algebra as the corresponding Lie algebra 𝔩θ\mathfrak{l}_{\theta} over kk.

Definition 2.8 (Parahoric adjoint sheaf).

For a parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor \mathcal{E}, we define the adjoint sheaf (𝔤)\mathcal{E}(\mathfrak{g}) as the sheaf of infinitesimal automorphisms of \mathcal{E}. Its structure is best understood by describing its local sections:

  • On any open set UXDU\subset X\setminus D, \mathcal{E} is a standard GG-bundle, and the sheaf (𝔤)\mathcal{E}(\mathfrak{g}) restricts to the standard adjoint vector bundle (|U)×G𝔤(\mathcal{E}|_{U})\times_{G}\mathfrak{g}.

  • Over the formal disk Spec(𝒪^X,x)\mathrm{Spec}(\hat{\mathcal{O}}_{X,x}) at a point xDx\in D, sections of (𝔤)\mathcal{E}(\mathfrak{g}) are identified with sections of the parahoric Lie algebra 𝔤θx\mathfrak{g}_{\theta_{x}}.

Thus, (𝔤)\mathcal{E}(\mathfrak{g}) is a coherent sheaf of 𝒪X\mathcal{O}_{X}-modules whose generic fiber is the complex Lie algebra 𝔤\mathfrak{g}, but whose stalks over the points xiDx_{i}\in D are the full parahoric Lie algebras 𝔤θxi(𝒪X,xi)\mathfrak{g}_{\theta_{x_{i}}}(\mathcal{O}_{X,x_{i}}).

The local Lie algebra decompositions at the points xiDx_{i}\in D allow us to define another sheaf that inherits this structure: The pro-unipotent radical sheaf, denoted by (𝔤+)\mathcal{E}(\mathfrak{g}^{+}), is the coherent subsheaf of (𝔤)\mathcal{E}(\mathfrak{g}) whose sections over the formal disk at each xiDx_{i}\in D correspond to elements of the pro-unipotent radical Lie algebra 𝔤θxi+(𝒪X,xi)\mathfrak{g}_{\theta_{x_{i}}}^{+}(\mathcal{O}_{X,x_{i}}).

2.4. Parahoric Degree

Let GG be a connected complex reductive Lie group. Fix a maximal torus TGT\subset G and let

θY(T)\theta\in Y(T)\otimes_{\mathbb{Z}}\mathbb{Q}

be a rational weight. Denote by Gθ(K)G(K)G_{\theta}(K)\subseteq G(K) the parahoric subgroup corresponding to θ\theta. Recall that a parabolic subgroup PGP\subset G (with Lie algebra 𝔭\mathfrak{p}) is determined by a subset of roots P\mathcal{R}_{P}\subseteq\mathcal{R}. In this context, we define the following subgroup of P(K)P(K)

Pθ(K):=T(A),Ur(zmr(θ)A),rP.P_{\theta}(K):=\Big{\langle}T(A),\;U_{r}\Big{(}z^{m_{r}(\theta)}A\Big{)},\;r\in\mathcal{R}_{P}\Big{\rangle}.

Let 𝒫θ\mathcal{P}_{\theta} be the corresponding group scheme on the formal disc 𝔻=Spec(R)\mathbb{D}=\operatorname{Spec}(R) (see, e.g., [18] for further details on this construction.) Furthermore, if

Gθ(K)GG_{\theta}(K)\longrightarrow G

is the evaluation map, then its image is a parabolic subgroup PθGP_{\theta}\subset G, whose inverse image is exactly Pθ(K)P_{\theta}(K).

We now describe the global situation. Let XX be a smooth projective curve over \mathbb{C} and fix a reduced effective divisor DXD\subset X. For each point xDx\in D, let θx\theta_{x} be a weight and denote by 𝒫θx\mathcal{P}_{\theta_{x}} the corresponding local group scheme. One defines a global group scheme 𝒫𝜽\mathcal{P}_{\boldsymbol{\theta}} on XX by gluing the local data:

{𝒫𝜽|XDP×(XD),𝒫𝜽|𝔻x=𝒫θx,xD.\begin{cases}\mathcal{P}_{\boldsymbol{\theta}}|_{X\setminus D}\cong P\times(X\setminus D),\\[2.84526pt] \mathcal{P}_{\boldsymbol{\theta}}|_{\mathbb{D}_{x}}=\mathcal{P}_{\theta_{x}},\quad x\in D.\end{cases}

Accordingly, we next define the pairing that will be used in the definition of our parahoric degree on the Lie algebra of the torus (Definition 2.9). The case in which 𝒫𝜽|𝔻x𝒫θx\mathcal{P}_{\boldsymbol{\theta}}\big{|}_{\mathbb{D}_{x}}\;\cong\;\mathcal{P}_{\theta_{x}} is handled in exactly the same way; the only difference is that the parahoric degree can be then defined after conjugating into the Lie algebra of a maximal torus.

By [10, Lemma 3.18], the group scheme 𝒫𝜽\mathcal{P}_{\boldsymbol{\theta}} is smooth, affine of finite type, and flat over XX. Moreover, one has an inclusion

𝒫𝜽𝒢𝜽,\mathcal{P}_{\boldsymbol{\theta}}\subset\mathcal{G}_{\boldsymbol{\theta}},

where 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}} is the Bruhat–Tits group scheme associated to 𝜽\boldsymbol{\theta}.

Let \mathcal{E} be a 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor on XX. A reduction of structure group of \mathcal{E} to 𝒫𝜽\mathcal{P}_{\boldsymbol{\theta}} is given by a section

ς:X/𝒫𝜽.\varsigma:X\longrightarrow\mathcal{E}/\mathcal{P}_{\boldsymbol{\theta}}.

Denote by ς\mathcal{E}_{\varsigma} the corresponding 𝒫𝜽\mathcal{P}_{\boldsymbol{\theta}}-torsor obtained from the Cartesian diagram

ς{\mathcal{E}_{\varsigma}}{\mathcal{E}}X{X}/𝒫𝜽.{\mathcal{E}/\mathcal{P}_{\boldsymbol{\theta}}.}ς\scriptstyle{\varsigma}

Let κ:𝒫𝜽𝔾m\kappa:\mathcal{P}_{\boldsymbol{\theta}}\to\mathbb{G}_{m} be a group scheme morphism (a character of 𝒫𝜽\mathcal{P}_{\boldsymbol{\theta}}). There is a natural one-to-one correspondence (see [24, Lemma 4.2])

Hom(𝒫𝜽,𝔾m)Hom(P,).\operatorname{Hom}(\mathcal{P}_{\boldsymbol{\theta}},\mathbb{G}_{m})\cong\operatorname{Hom}(P,\mathbb{C}^{*}).

Denote by χ:P\chi:P\to\mathbb{C}^{*} the character corresponding to κ\kappa. For any weight θ\theta, we define the pairing

θ,κ:=θ,χ.\langle\theta,\kappa\rangle:=\langle\theta,\chi\rangle.

Returning to the parahoric torsor ς\mathcal{E}_{\varsigma}, the pushforward via κ\kappa defines a line bundle on XX, which we denote by

L(ς,κ):=κ(ς).L(\varsigma,\kappa):=\kappa_{*}(\mathcal{E}_{\varsigma}).

In the special case when P=GP=G (and if the reduction ς:X/𝒢𝜽\varsigma:X\to\mathcal{E}/\mathcal{G}_{\boldsymbol{\theta}} is trivial), one writes L(κ):=κL(\kappa):=\kappa_{*}\mathcal{E}.

The following notion of parahoric degree was introduced in [24].

Definition 2.9.

[24, Definition 4.2] Let \mathcal{E} be a 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor on XX, and let ς\varsigma be a reduction of structure group to 𝒫𝜽\mathcal{P}_{\boldsymbol{\theta}}. For a character κ:𝒫𝜽𝔾m\kappa:\mathcal{P}_{\boldsymbol{\theta}}\to\mathbb{G}_{m}, the parahoric degree of \mathcal{E} with respect to ς\varsigma and κ\kappa is defined by

parhdeg(ς,κ)=degL(ς,κ)+xDθx,κ.\operatorname{parhdeg}\mathcal{E}(\varsigma,\kappa)=\deg L(\varsigma,\kappa)+\sum_{x\in D}\langle\theta_{x},\kappa\rangle.

If ς\varsigma is a trivial reduction of \mathcal{E} to 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}, we define

parhdeg(κ)=degL(κ)+xDθx,κ.\operatorname{parhdeg}\mathcal{E}(\kappa)=\deg L(\kappa)+\sum_{x\in D}\langle\theta_{x},\kappa\rangle.

Using this notion of parahoric degree, we define stability for parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsors as follows:

Definition 2.10.

[24, Definition 4.3] A parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor \mathcal{E} on XX is called stable (resp. semistable) if for every proper parabolic subgroup PGP\subset G, every reduction of structure group

ς:X/𝒫𝜽\varsigma:X\to\mathcal{E}/\mathcal{P}_{\boldsymbol{\theta}}

and every nontrivial anti-dominant character κ:𝒫𝜽𝔾m\kappa:\mathcal{P}_{\boldsymbol{\theta}}\to\mathbb{G}_{m}, one has

parhdeg(ς,κ)>0(resp. 0).\operatorname{parhdeg}\mathcal{E}(\varsigma,\kappa)>0\quad(\text{resp. }\geq 0).
Remark 2.11.

A moduli space of semistable parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsors on XX is constructed in [2], where 𝜽=(θx)xD\boldsymbol{\theta}=(\theta_{x})_{x\in D} involves rational weights θxY(T)\theta_{x}\in Y(T)\otimes_{\mathbb{Z}}\mathbb{Q}. The choice of rational weights is generally sufficient for the theory of these moduli spaces. This is because the isomorphism class of the local parahoric group scheme 𝒢θx\mathcal{G}_{\theta_{x}} over Spec([[zx]])\operatorname{Spec}(\mathbb{C}[\![z_{x}]\!]) (which in turn determines the structure of 𝒫θx\mathcal{P}_{\theta_{x}}) depends only on the open facet of the affine apartment (in the Bruhat–Tits building associated with G(Kx)G(K_{x})) to which θx\theta_{x} belongs. Since each such facet necessarily contains rational points, any distinct parahoric group scheme structure defined by this framework can be represented by a rational weight lying within the same facet. Consequently, the moduli theory developed using rational weights comprehensively addresses the range of these algebro-geometric structures.

2.5. Stable Parahoric Torsors and Associated Vector Bundles

Let GG be a connected complex reductive group and

ρ:GGL(V)\rho\colon G\to\mathrm{GL}(V)

be a rational representation on a finite-dimensional complex vector space VV. Suppose that VV decomposes into a direct sum of irreducible GG-modules

V=V1Vr,V=V_{1}\oplus\cdots\oplus V_{r},

so that ρ=ρ1ρr\rho=\rho_{1}\oplus\cdots\oplus\rho_{r}, where ρi:GGL(Vi)\rho_{i}\colon G\to\mathrm{GL}(V_{i}) is the representation on ViV_{i}. Following [29, Section 3], we define

S:=ρ(G)GL(V),andS:=\rho(G)\subset\mathrm{GL}(V),\quad\text{and}
C:={(λ1ρ1(g),,λrρr(g))GL(V)λi,gG}.C:=\{(\lambda_{1}\rho_{1}(g),\dots,\lambda_{r}\rho_{r}(g))\in\mathrm{GL}(V)\mid\lambda_{i}\in\mathbb{C}^{*},\;g\in G\}.

Since CC is the image of the rational homomorphism ((λi),g)(λ1ρ1(g),,λrρr(g))((\lambda_{i}),g)\mapsto(\lambda_{1}\rho_{1}(g),\dots,\lambda_{r}\rho_{r}(g)) from r×G\mathbb{C}^{*r}\times G to GL(V)\mathrm{GL}(V), it is a constructible, hence locally closed, subset of GL(V)\mathrm{GL}(V). We denote by C¯\overline{C} its Zariski closure in End(V)\mathrm{End}(V). The sets S,C,S,C, and C¯\overline{C} are invariant under left and right multiplication by elements of SS.

Let \mathcal{E} be a parahoric torsor for a parahoric group scheme 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}} over a smooth projective variety XX. Via the representation ρ\rho, we obtain an associated vector bundle (V)=×𝒢𝜽V\mathcal{E}(V)=\mathcal{E}\times^{\mathcal{G}_{\boldsymbol{\theta}}}V. For each irreducible component ViV_{i}, (Vi)\mathcal{E}(V_{i}) is the corresponding vector subbundle. Similarly, using the adjoint representation Ad:GGL(𝔤)\mathrm{Ad}:G\to\mathrm{GL}(\mathfrak{g}), where 𝔤=Lie(G)\mathfrak{g}=\mathrm{Lie}(G), we obtain the Lie algebra bundle (𝔤)\mathcal{E}(\mathfrak{g}).

We next extend classical arguments for principal GG-bundles, particularly those of Ramanathan [29], to the setting of parahoric torsors. We analyze the structure of certain endomorphisms, stabilizers of subspaces, and establish slope inequalities using the parahoric degree. These results culminate in a uniqueness property for homomorphisms between associated vector bundles (Proposition 2.16) and a characterization of global sections of the adjoint bundle associated to a stable parahoric torsor (Proposition 2.18).

Lemma 2.12.

Let ρ:GGL(V)\rho\colon G\to\mathrm{GL}(V) be an irreducible representation. Let

𝒞ρ:={μρ(g)μ,gG}¯End(V).\mathcal{C}_{\rho}:=\overline{\{\mu\,\rho(g)\mid\mu\in\mathbb{C}^{*},\,g\in G\}}\subset\mathrm{End}(V).

Any endomorphism T𝒞ρT\in\mathcal{C}_{\rho} is of the form

λρ(k1)pρ(k2),\lambda\,\rho(k_{1})\,p_{\mathcal{R}}\,\rho(k_{2}),

for some λ\lambda\in\mathbb{C}, elements k1,k2Gk_{1},k_{2}\in G, and a projection pp_{\mathcal{R}} onto V:=αD()VαV_{\mathcal{R}}:=\bigoplus_{\alpha\in D(\mathcal{R})}V_{\alpha}. Here, D()D(\mathcal{R}) is the set of weights α=Λαimiαi\alpha=\Lambda-\sum_{\alpha_{i}\in\mathcal{R}}m_{i}\alpha_{i} (where Λ\Lambda is the highest weight of VV, mi0m_{i}\geq 0 integers, and {αj}\{\alpha_{j}\} is a fixed system of simple roots of 𝔤\mathfrak{g}), for some subset \mathcal{R} of the simple roots. Conversely, every endomorphism of this form lies in 𝒞ρ\mathcal{C}_{\rho}.

Proof.

Let T𝒞ρT\in\mathcal{C}_{\rho}. Then T=limnμnρ(gn)T=\lim_{n\to\infty}\mu_{n}\,\rho(g_{n}), for some μn\mu_{n}\in\mathbb{C}^{*} and gnGg_{n}\in G. Let KK be a maximal compact subgroup of GG. By the Cartan decomposition G=KAKG=KAK, we can write gn=k1,nank2,ng_{n}=k_{1,n}a_{n}k_{2,n} where k1,n,k2,nKk_{1,n},k_{2,n}\in K and an=exp(hn)a_{n}=\exp(h_{n}) for hniLie(A0)h_{n}\in i\mathrm{Lie}(A_{0}), where A0A_{0} is a maximal \mathbb{R}-split torus whose Lie algebra is 𝔞0\mathfrak{a}_{0}. By passing to a subsequence, we may assume that k1,nk1Kk_{1,n}\to k_{1}\in K and k2,nk2Kk_{2,n}\to k_{2}\in K. Then T=ρ(k1)(limnμnρ(an))ρ(k2)T=\rho(k_{1})\left(\lim_{n\to\infty}\mu_{n}\rho(a_{n})\right)\rho(k_{2}). Let hni𝔞h_{n}\in i\mathfrak{a} (where 𝔞\mathfrak{a} is the Lie algebra of a maximal torus AKA\subset K, so i𝔞i\mathfrak{a} is a Cartan subalgebra of 𝔤\mathfrak{g}^{\mathbb{C}} up to conjugation) such that exp(hn)\exp(h_{n}) is conjugate to ana_{n}. By passing to a further subsequence and conjugating AA if necessary, we can assume all hnh_{n} lie in the closure of a fixed Weyl chamber, and αj(hn)\alpha_{j}(h_{n}) (for simple roots αj\alpha_{j}) have defined limiting behaviors. The operator limμnρ(an)\lim\mu_{n}\rho(a_{n}) acts on each weight space VαV_{\alpha} by multiplication by Cα=limμneα(hn)C_{\alpha}=\lim\mu_{n}e^{\alpha(h_{n})}. Let Λ\Lambda be the highest weight of VV. Then Cα=limμne(αΛ)(hn)eΛ(hn)C_{\alpha}=\lim\mu_{n}e^{(\alpha-\Lambda)(h_{n})}e^{\Lambda(h_{n})}. For CΛ=limμneΛ(hn)C_{\Lambda}=\lim\mu_{n}e^{\Lambda(h_{n})} to be finite and non-zero (or CΛ=0C_{\Lambda}=0 if λ=0\lambda=0), we analyze (αΛ)(hn)(\alpha-\Lambda)(h_{n}). There is a subset of simple roots {α1,,αl}\mathcal{R}\subseteq\{\alpha_{1},\dots,\alpha_{l}\} such that (αΛ)(hn)(\alpha-\Lambda)(h_{n}) tends to a finite limit if α=Λαimiαi\alpha=\Lambda-\sum_{\alpha_{i}\in\mathcal{R}}m_{i}\alpha_{i}, for mi0m_{i}\geq 0, and tends to -\infty if α\alpha involves simple roots not in \mathcal{R} with negative coefficients in Λα\Lambda-\alpha. Thus, limμnρ(an)\lim\mu_{n}\rho(a_{n}) becomes λp\lambda^{\prime}p_{\mathcal{R}}, for some scalar λ\lambda^{\prime}, and the projection pp_{\mathcal{R}} onto V=αD()VαV_{\mathcal{R}}=\bigoplus_{\alpha\in D(\mathcal{R})}V_{\alpha}. Therefore, T=λρ(k1)pρ(k2)T=\lambda\rho(k_{1})p_{\mathcal{R}}\rho(k_{2}^{\prime}), for some k2k_{2}^{\prime} (after absorbing ρ(k1)1ρ(k1)\rho(k_{1})^{-1}\rho(k_{1}) and adjusting k2k_{2}).

For the converse, as in [29, Lemma 3.1], choose hmi𝔞h_{m}\in i\mathfrak{a} such that αi(hm)=0\alpha_{i}(h_{m})=0 for αi\alpha_{i}\in\mathcal{R} and αj(hm)=m\alpha_{j}(h_{m})=m for αj\alpha_{j}\notin\mathcal{R}. Let μm=λeΛ(hm)\mu_{m}=\lambda e^{-\Lambda(h_{m})}. Then μmρ(exphm)\mu_{m}\rho(\exp h_{m}) converges to λp\lambda p_{\mathcal{R}}. Pre- and post-multiplying by ρ(k1)\rho(k_{1}) and ρ(k2)\rho(k_{2}) gives the desired form. ∎

Lemma 2.13.

Let VV_{\mathcal{R}} be a subspace as defined in Lemma 2.12, corresponding to an irreducible representation ρ\rho and a subset of simple roots \mathcal{R}. The stabilizer P=StabG(V)P_{\mathcal{R}}=\mathrm{Stab}_{G}(V_{\mathcal{R}}) is a parabolic subgroup of GG. If Z0Z_{0} is the identity component of the center of GG, and P=Z0PP_{\mathcal{R}}=Z_{0}P^{\prime}_{\mathcal{R}}, then the character χ\chi_{\mathcal{R}} of PP_{\mathcal{R}} given by its action on det(V)\det(V_{\mathcal{R}}) has its restriction to PP^{\prime}_{\mathcal{R}} dominant. Moreover, if V0V_{\mathcal{R}}\neq 0 and VVV_{\mathcal{R}}\neq V, then χ\chi_{\mathcal{R}} (or a related character used for stability) is non-trivial in the appropriate sense for stability arguments.

Proof.

The Lie algebra of StabG(V)\mathrm{Stab}_{G}(V_{\mathcal{R}}) contains the Borel subalgebra 𝔟\mathfrak{b} (corresponding to the choice of simple roots) plus all root spaces 𝔤α\mathfrak{g}^{\alpha} such that α\alpha is a sum of simple roots where those not in \mathcal{R} appear with non-negative coefficients. This defines a parabolic subalgebra, so PP_{\mathcal{R}} is parabolic. The assertion concerning the character χ\chi_{\mathcal{R}} follows by considering the PP_{\mathcal{R}}-action on the line dimVV\bigwedge^{\dim V_{\mathcal{R}}}V_{\mathcal{R}}, whose weight is a sum of weights in D()D(\mathcal{R}), and is dominant for PP^{\prime}_{\mathcal{R}}. If VV_{\mathcal{R}} is a proper non-zero PP_{\mathcal{R}}-invariant subspace, then PP_{\mathcal{R}} is a proper parabolic, and such characters are non-trivial. ∎

Definition 2.14.

Let ρ:GGL(V)\rho:G\to\mathrm{GL}(V) be an irreducible representation, and \mathcal{R} be a subset of simple roots. A vector subbundle W(V)W\subset\mathcal{E}(V) is said to be of type \mathcal{R} if for a generic point xXx\in X, the fiber Wx(V)xVW_{x}\subset\mathcal{E}(V)_{x}\cong V is GG-conjugate to the subspace V=αD()VαV_{\mathcal{R}}=\bigoplus_{\alpha\in D(\mathcal{R})}V_{\alpha}.

Lemma 2.15.

Let \mathcal{E} be a stable (resp. semistable) parahoric 𝒢𝛉\mathcal{G}_{\boldsymbol{\theta}}-torsor on XX, and let ρ:GGL(V)\rho\colon G\to\mathrm{GL}(V) be an irreducible representation. If W(V)W\subset\mathcal{E}(V) is a proper, non-zero subbundle of type \mathcal{R} for some {α1,,αl}\mathcal{R}\subset\{\alpha_{1},\dots,\alpha_{l}\}, then

parhdegWrkW<parhdeg(V)rkV(resp. parhdegWrkWparhdeg(V)rkV).\frac{\operatorname{parhdeg}W}{\mathrm{rk}W}<\frac{\operatorname{parhdeg}\mathcal{E}(V)}{\mathrm{rk}V}\quad\left(\text{resp. }\frac{\operatorname{parhdeg}W}{\mathrm{rk}W}\leq\frac{\operatorname{parhdeg}\mathcal{E}(V)}{\mathrm{rk}V}\right).
Proof.

A subbundle WW of type \mathcal{R} corresponds to a section σ:X(G/P)\sigma\colon X\to\mathcal{E}(G/P_{\mathcal{R}}), where P=StabG(V)P_{\mathcal{R}}=\mathrm{Stab}_{G}(V_{\mathcal{R}}) is a proper parabolic subgroup (Lemma 2.13). This section σ\sigma defines a reduction of the structure group of \mathcal{E} from 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}} to a parahoric subgroup scheme 𝒫𝜽,\mathcal{P}_{\boldsymbol{\theta},\mathcal{R}} whose generic fiber is PP_{\mathcal{R}}. Let χ\chi be the character of PP_{\mathcal{R}} given by its action on det(V)\det(V_{\mathcal{R}}). We form a character μ\mu of PP_{\mathcal{R}}, trivial on Z0(G)Z_{0}(G) and dominant on PP^{\prime}_{\mathcal{R}} (e.g., μ=(det|V)rkV(det|V)rkV\mu=(\det|_{V_{\mathcal{R}}})^{\mathrm{rk}V}\otimes(\det|_{V})^{-\mathrm{rk}V_{\mathcal{R}}}, restricted to PP^{\prime}_{\mathcal{R}} and extended trivially on Z0(G)Z_{0}(G)). The definition of (semi)stability for the parahoric torsor \mathcal{E} implies that the parahoric degree of the associated line bundle μ=σ(μ)\mathcal{L}_{\mu}=\sigma^{*}\mathcal{E}(\mu) satisfies parhdegμ<0\operatorname{parhdeg}\mathcal{L}_{\mu}<0 (resp. 0\leq 0). The parahoric degree of μ\mathcal{L}_{\mu} is given by

parhdegμ=(rkV)parhdegW(rkW)parhdeg(V),\operatorname{parhdeg}\mathcal{L}_{\mu}=(\mathrm{rk}V)\operatorname{parhdeg}W-(\mathrm{rk}W)\operatorname{parhdeg}\mathcal{E}(V_{\mathcal{R}}),

where VV_{\mathcal{R}} is identified with WW via the section. More directly,

parhdegμ=(rkV)(rkW)(parhdegWrkWparhdeg(V)rkV).\operatorname{parhdeg}\mathcal{L}_{\mu}=(\mathrm{rk}V)(\mathrm{rk}W)\left(\frac{\operatorname{parhdeg}W}{\mathrm{rk}W}-\frac{\operatorname{parhdeg}\mathcal{E}(V)}{\mathrm{rk}V}\right).

The inequality on parhdegμ\operatorname{parhdeg}\mathcal{L}_{\mu} then translates directly to the stated slope inequality for WW. ∎

Proposition 2.16.

Let \mathcal{E} and \mathcal{E}^{\prime} be parahoric 𝒢𝛉\mathcal{G}_{\boldsymbol{\theta}}-torsors on XX having the same topological type, that is, parhdeg(Vi)=parhdeg(Vi)\operatorname{parhdeg}\mathcal{E}(V_{i})=\operatorname{parhdeg}\mathcal{E}^{\prime}(V_{i}), for any representation ViV_{i}. Suppose \mathcal{E} is stable and \mathcal{E}^{\prime} is semistable. Let

s=(s1,,sr)i=1rH0(X,Hom((Vi),(Vi)))s=(s_{1},\dots,s_{r})\in\bigoplus_{i=1}^{r}H^{0}\Bigl{(}X,\,\mathrm{Hom}\bigl{(}\mathcal{E}(V_{i}),\mathcal{E}^{\prime}(V_{i})\bigr{)}\Bigr{)}

be a section such that for each xXx\in X, s(x)C¯s(x)\in\overline{C} (when viewed as an endomorphism in End(V)\mathrm{End}(V) after trivializing x(V)\mathcal{E}_{x}(V) and x(V)\mathcal{E}^{\prime}_{x}(V)). Then, for each ii, the induced homomorphism si:(Vi)(Vi)s_{i}\colon\mathcal{E}(V_{i})\to\mathcal{E}^{\prime}(V_{i}) is either identically zero or an isomorphism. Moreover, if si0s_{i}\neq 0 for every ii and if ρ=ρi\rho=\bigoplus\rho_{i} is faithful, then there exist scalars λi\lambda_{i}\in\mathbb{C}^{*}, 1ir1\leq i\leq r, such that (λ1s1,,λrsr)(\lambda_{1}s_{1},\dots,\lambda_{r}s_{r}) is induced by an isomorphism of parahoric torsors \mathcal{E}\stackrel{{\scriptstyle\sim}}{{\to}}\mathcal{E}^{\prime}.

Proof.

Suppose that for some ii, the homomorphism si:(Vi)(Vi)s_{i}\colon\mathcal{E}(V_{i})\to\mathcal{E}^{\prime}(V_{i}) is not identically zero and not an isomorphism. Then its image Wi=Im(si)W_{i}=\mathrm{Im}(s_{i}) and its kernel U1,i=Ker(si)U_{1,i}=\mathrm{Ker}(s_{i}) are proper, non-zero subbundles (on a suitable open set).

Since si(x)C¯is_{i}(x)\in\overline{C}_{i}, the image subbundle Wi(Vi)W_{i}\subset\mathcal{E}^{\prime}(V_{i}) is of type i\mathcal{R}_{i} for some subset of simple roots i\mathcal{R}_{i}. As \mathcal{E}^{\prime} is semistable, Lemma 2.15 implies that μpar(Wi)μpar((Vi))\mu_{\text{par}}(W_{i})\leq\mu_{\text{par}}(\mathcal{E}^{\prime}(V_{i})). From the short exact sequence 0U1,i(Vi)Wi00\to U_{1,i}\to\mathcal{E}(V_{i})\to W_{i}\to 0 and the fact that μpar((Vi))=μpar((Vi))\mu_{\text{par}}(\mathcal{E}(V_{i}))=\mu_{\text{par}}(\mathcal{E}^{\prime}(V_{i})), this leads to the inequality

μpar(U1,i)μpar((Vi)).\mu_{\text{par}}(U_{1,i})\geq\mu_{\text{par}}(\mathcal{E}(V_{i})).

Now, we analyze the kernel subbundle U1,i(Vi)U_{1,i}\subset\mathcal{E}(V_{i}). For any point xXx\in X, the fiber (U1,i)x(U_{1,i})_{x} is the kernel of the endomorphism si(x)C¯is_{i}(x)\in\overline{C}_{i}. By Lemma 2.12, an endomorphism in C¯i\overline{C}_{i} has a kernel that is GG-conjugate to a standard subspace of the form VV_{\mathcal{R}^{\prime}} (the sum of weight spaces not in the image of the projection pp_{\mathcal{R}}). Therefore, the kernel subbundle U1,iU_{1,i} is also of type \mathcal{R}^{\prime} for some \mathcal{R}^{\prime}.

Since \mathcal{E} is stable, we can apply Lemma 2.15 directly to the proper, non-zero subbundle U1,i(Vi)U_{1,i}\subset\mathcal{E}(V_{i}). This gives the strict slope inequality

μpar(U1,i)<μpar((Vi)).\mu_{\text{par}}(U_{1,i})<\mu_{\text{par}}(\mathcal{E}(V_{i})).

This is a direct contradiction to our earlier finding that μpar(U1,i)μpar((Vi))\mu_{\text{par}}(U_{1,i})\geq\mu_{\text{par}}(\mathcal{E}(V_{i})). Therefore, each homomorphism sis_{i} must be either identically zero or an isomorphism.

If sis_{i} is a generic isomorphism, then det(si)\det(s_{i}) is a holomorphic section of det((Vi))det((Vi))\det(\mathcal{E}(V_{i})^{*})\otimes\det(\mathcal{E}^{\prime}(V_{i})). Let ki=rkVik_{i}=\mathrm{rk}V_{i}, and then this line bundle is ki(Vi)ki(Vi)\bigwedge^{k_{i}}\mathcal{E}(V_{i})^{*}\otimes\bigwedge^{k_{i}}\mathcal{E}^{\prime}(V_{i}). Since \mathcal{E} and \mathcal{E}^{\prime} have the same topological type, this line bundle has parahoric degree zero. A non-zero holomorphic section of a degree-zero line bundle over a projective variety XX must be nowhere vanishing (if H0(X,𝒪X)=H^{0}(X,\mathcal{O}_{X})=\mathbb{C}). Thus, det(si)\det(s_{i}) is nowhere zero, and sis_{i} is an isomorphism. Therefore, each sis_{i} is either zero or an isomorphism.

For the second part, assume that each si0s_{i}\neq 0, so each sis_{i} is an isomorphism. Then s(X)(C)s(X)\subset\mathcal{E}(C) (where (C)\mathcal{E}(C) denotes sections whose values at each point xx lie in CC relative to the fiber x(V)\mathcal{E}_{x}(V)). This means that, for each xXx\in X, there exist λi(x)\lambda_{i}(x)\in\mathbb{C}^{*} and g(x)Gg(x)\in G such that sj(x)=λj(x)ρj(g(x))s_{j}(x)=\lambda_{j}(x)\rho_{j}(g(x)), for all j=1,,rj=1,\dots,r. The argument from [29, p. 136] involving characters χk\chi_{k} on r\mathbb{C}^{*r} shows that χk(λ1(x),,λr(x))\chi_{k}(\lambda_{1}(x),\dots,\lambda_{r}(x)) are constant. This allows finding constants λj\lambda_{j}\in\mathbb{C}^{*} such that (λ1s1,,λrsr)(\lambda_{1}s_{1},\dots,\lambda_{r}s_{r}) takes values in (S)\mathcal{E}(S). If ρ\rho is faithful, this implies that (λ1s1,,λrsr)(\lambda_{1}s_{1},\dots,\lambda_{r}s_{r}) is induced by an isomorphism of parahoric torsors \mathcal{E}\stackrel{{\scriptstyle\sim}}{{\to}}\mathcal{E}^{\prime}. ∎

Remark 2.17.

The condition s(X)(C¯)s(X)\subset\mathcal{E}(\overline{C}) means that for any xXx\in X, s(x)s(x), viewed as an element of Hom(V,V)\mathrm{Hom}(V,V) via trivializations of x(V)\mathcal{E}_{x}(V) and x(V)\mathcal{E}^{\prime}_{x}(V), lies in C¯\overline{C}. If si(x)s_{i}(x) is not an isomorphism, it corresponds to an element in C¯iCi\overline{C}_{i}\setminus C_{i}. If s(X)s(X) were not contained in (C¯)\mathcal{E}(\overline{C}), then the arguments relying on Lemma 2.12 (about the structure of elements in C¯\overline{C}) would not directly apply. Note that the assumption s(X)(C¯)s(X)\subset\mathcal{E}(\overline{C}) is also used in [29, Proposition 3.1] for the case of principal GG-bundles.

Proposition 2.18.

Let \mathcal{E} be a stable parahoric 𝒢𝛉\mathcal{G}_{\boldsymbol{\theta}}-torsor on XX. Then the space of global sections of the adjoint bundle H0(X,(𝔤))H^{0}\bigl{(}X,\mathcal{E}(\mathfrak{g})\bigr{)} is isomorphic to the center 𝔷\mathfrak{z} of the Lie algebra 𝔤\mathfrak{g}. In particular, if GG is semisimple (so that 𝔷=0\mathfrak{z}=0), then H0(X,(𝔤))=0H^{0}\bigl{(}X,\mathcal{E}(\mathfrak{g})\bigr{)}=0, and consequently, the group of automorphisms Aut()\mathrm{Aut}(\mathcal{E}) that are extensions of the identity on XX is discrete (and finite if XX is projective).

Proof.

Let AutX()\operatorname{Aut}_{X}(\mathcal{E}) be the group of automorphisms of \mathcal{E} covering the identity on XX. Its Lie algebra is Lie(AutX())H0(X,(𝔤))\mathrm{Lie}(\operatorname{Aut}_{X}(\mathcal{E}))\cong H^{0}(X,\mathcal{E}(\mathfrak{g})). Let ρ=i=1rρi:GGL(V=Vi)\rho=\bigoplus_{i=1}^{r}\rho_{i}:G\to\mathrm{GL}(V=\bigoplus V_{i}) be a faithful representation, where each ViV_{i} is irreducible. An automorphism αAutX()\alpha\in\operatorname{Aut}_{X}(\mathcal{E}) induces isomorphisms si(α):(Vi)(Vi)s_{i}(\alpha):\mathcal{E}(V_{i})\to\mathcal{E}(V_{i}) for each ii. By Proposition 2.16 (with =\mathcal{E}^{\prime}=\mathcal{E}, which is stable and thus semistable), each si(α)s_{i}(\alpha) is an isomorphism. As in [29, Proof of Prop 3.2], the induced sections (s1(α),,sr(α))(s_{1}(\alpha),\dots,s_{r}(\alpha)) define a map from AutX()\operatorname{Aut}_{X}(\mathcal{E}) to a product of projective spaces of sections (after projectivizing and considering H0(X,(EndVi))H^{0}(X,\mathcal{E}(\mathrm{End}V_{i}))). The image of this map is finite. More precisely, there is a homomorphism Φ:AutX()I\Phi:\operatorname{Aut}_{X}(\mathcal{E})\to I, where II is a finite group. This implies that the connected component of the identity, AutX0()\operatorname{Aut}_{X}^{0}(\mathcal{E}), is contained in KerΦ\text{Ker}\Phi. Thus Lie(AutX())=Lie(AutX0())Lie(KerΦ)\mathrm{Lie}(\operatorname{Aut}_{X}(\mathcal{E}))=\mathrm{Lie}(\operatorname{Aut}_{X}^{0}(\mathcal{E}))\subseteq\mathrm{Lie}(\text{Ker}\Phi). An element hH0(X,(𝔤))h\in H^{0}(X,\mathcal{E}(\mathfrak{g})) corresponds to a 1-parameter subgroup exp(th)\exp(th) in AutX0()\operatorname{Aut}_{X}^{0}(\mathcal{E}). If exp(th)KerΦ\exp(th)\in\text{Ker}\Phi, then for each ii, the induced isomorphism si(exp(th))s_{i}(\exp(th)) on (Vi)\mathcal{E}(V_{i}) must correspond to the identity element in the relevant factor of II. The isomorphism si(exp(th))s_{i}(\exp(th)) is given by exp(tρi(h))\exp(t\cdot\rho_{i}(h)) acting on the fibers (where ρi(h)\rho_{i}(h) here denotes the action derived from the adjoint action, i.e., (ρi(h))\mathcal{E}(\rho_{i}(h))). For exp(tρi(h))\exp(t\cdot\rho_{i}(h)) to effectively be trivial in II for all tt, ρi(h)\rho_{i}(h) must correspond to a scalar endomorphism for each ii. Since ρi\rho_{i} is an irreducible representation of GG, if ρi(h)\rho_{i}(h) is a scalar matrix, and h𝔤h\in\mathfrak{g}, then by Schur’s Lemma, ρi(h)=ciIdVi\rho_{i}(h)=c_{i}\cdot\operatorname{Id}_{V_{i}} for cic_{i}\in\mathbb{C}. If hh is in 𝔤\mathfrak{g}, then ρi(h)\rho_{i}(h) is an endomorphism that comes from 𝔤\mathfrak{g}. For ρi(h)\rho_{i}(h) to be scalar for all ii in a faithful representation ρ=ρi\rho=\bigoplus\rho_{i}, hh must lie in the center 𝔷\mathfrak{z} of 𝔤\mathfrak{g}. Thus, H0(X,(𝔤))=Lie(AutX0())𝔷H^{0}(X,\mathcal{E}(\mathfrak{g}))=\mathrm{Lie}(\operatorname{Aut}_{X}^{0}(\mathcal{E}))\subseteq\mathfrak{z}. Conversely, if h𝔷h\in\mathfrak{z}, then exp(th)\exp(th) is a 1-parameter subgroup of Z(G)Z(G) (center of GG). These act as global automorphisms of \mathcal{E} and correspond to sections in H0(X,(𝔷))H0(X,𝒪X)𝔷𝔷H^{0}(X,\mathcal{E}(\mathfrak{z}))\cong H^{0}(X,\mathcal{O}_{X})\otimes\mathfrak{z}\cong\mathfrak{z} (if H0(X,𝒪X)=H^{0}(X,\mathcal{O}_{X})=\mathbb{C}). Therefore, H0(X,(𝔤))𝔷H^{0}(X,\mathcal{E}(\mathfrak{g}))\cong\mathfrak{z}.

If GG is semisimple, then 𝔷=0\mathfrak{z}=0, and so H0(X,(𝔤))=0H^{0}(X,\mathcal{E}(\mathfrak{g}))=0. This implies that AutX0()\operatorname{Aut}_{X}^{0}(\mathcal{E}) is trivial, so AutX()\operatorname{Aut}_{X}(\mathcal{E}) is discrete. If XX is projective, it is known that such an automorphism group is finite (e.g., it is an algebraic group scheme). ∎

3. DD-level Structure on Parahoric Torsors

In this section we introduce the concept of a DD-level structure on a parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor over a curve, generalizing the notion of DD-level structure on vector bundles given by Seshadri in [32]. We then construct their moduli space as an irreducible, normal, projective variety.

3.1. DD-level structures and stability

As in the previous section, let XX be a smooth complex projective curve and let KXK_{X} represent the holomorphic cotangent bundle of XX. We consider a reduced effective divisor DD on XX and a connected complex reductive Lie group GG.

A parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor π:X\pi:\mathcal{E}\to X over XX is a holomorphic fiber bundle over XX equipped with a holomorphic right-action of the group 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}, q:×𝒢𝜽q:\mathcal{E}\times\mathcal{G}_{\boldsymbol{\theta}}\to\mathcal{E}, satisfying:

π(q(z,g))=π(z), for all (z,g)×𝒢𝜽.\pi(q(z,g))=\pi(z),\mbox{ for all }(z,g)\in\mathcal{E}\times\mathcal{G}_{\boldsymbol{\theta}}.

For notational convenience, the point q(z,g)q(z,g)\in\mathcal{E} will be denoted by zgz\cdot g. For any xXx\in X, the fiber π1(x)\pi^{-1}(x)\subset\mathcal{E} will be also denoted by x\mathcal{E}_{x}.

The following definition extends to the case of parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsors the fundamental notion of a DD-level structure on vector bundles introduced by Seshadri in [32, Quatrième Partie].

Definition 3.1.

(DD-level structure) A DD-level structure of parahoric type (θx)xD(\theta_{x})_{x\in D} on a parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor \mathcal{E} over XX is a choice of section ηx:𝔻xGθx/Gθx+\eta_{x}:\mathbb{D}_{x}\to G_{\theta_{x}}/G_{\theta_{x}}^{+}, for each xDx\in D. We will denote a parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor equipped with a DD-level structure as a pair (,η)(\mathcal{E},\eta).

We say that two pairs (1,η1)(\mathcal{E}_{1},\eta_{1}) and (2,η2)(\mathcal{E}_{2},\eta_{2}) are equivalent if there exists an isomorphism h:12h:\mathcal{E}_{1}\to\mathcal{E}_{2} that sends G1,θx/G1,θx+G_{1,\theta_{x}}/G_{1,\theta_{x}}^{+} to G2,θx/G2,θx+G_{2,\theta_{x}}/G_{2,\theta_{x}}^{+} and such that hη1,x=η2,xh\circ\eta_{1,x}=\eta_{2,x}.

Note that for each point xDx\in D, the DD-level structure defines locally over a formal disk 𝔻x\mathbb{D}_{x} a reduction of structure group of GθxG_{\theta_{x}}-torsor to a Gθx+G_{\theta_{x}}^{+}-torsor. For each xDx\in D, there is a natural quotient map qx:GθxGθx/Gθx+q_{x}:G_{\theta_{x}}\to G_{\theta_{x}}/G_{\theta_{x}}^{+}. Let also

EGθx+:=qx1(ηx(𝔻x))GθxE_{G_{\theta_{x}}^{+}}:=q_{x}^{-1}(\eta_{x}(\mathbb{D}_{x}))\subset G_{\theta_{x}}

as a principal Gθx+G_{\theta_{x}}^{+}-bundle, and we define Gθx+(𝔤θx+)x(𝔤θx)\mathcal{E}_{G_{\theta_{x}}^{+}}(\mathfrak{g}_{\theta_{x}}^{+})\subset\mathcal{E}_{x}(\mathfrak{g}_{\theta_{x}}). Locally, a choice of ηx\eta_{x} chooses a Lie subalgebra 𝔤θx+\mathfrak{g}_{\theta_{x}}^{+} of 𝔤θx\mathfrak{g}_{\theta_{x}}. As so we have the short exact sequence

0(𝔤)η(𝔤)xD(𝔤θx/𝔤θx+)0.0\to\mathcal{E}(\mathfrak{g})_{\eta}\to\mathcal{E}(\mathfrak{g})\to\bigoplus_{x\in D}\mathcal{E}(\mathfrak{g}_{\theta_{x}}/\mathfrak{g}_{\theta_{x}}^{+})\to 0.

Here (𝔤)η\mathcal{E}(\mathfrak{g})_{\eta} is the associated adjoint sheaf to a new torsor, η\mathcal{E}_{\eta}, which is obtained from \mathcal{E} by reducing the structure group from 𝒢θx\mathcal{G}_{\theta_{x}} to 𝒢θx+\mathcal{G}_{\theta_{x}}^{+} at each point xDx\in D.

Remark 3.2.

In [36, Section 4.2], Yun defines bundles with a notion of parahoric level structure at a varying point as a reduction to the Borel subgroup. This notion was introduced in order to introduce parahoric versions of the Hitchin stacks.

Definition 3.3.

We call a DD-level structure on a parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor stable (resp. semistable) if the underlying parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor is stable (resp. semistable).

3.2. From Parahoric to Equivariant Level Structures

In order to establish the construction of a moduli space of stable pairs (,η)(\mathcal{E},\eta), we first show the equivalence of the notion of DD-level structure on a parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor to a notion of level structure on an equivariant principal GG-bundle with respect to a cyclic group action; we refer to [2], [24] for more details on the correspondence between parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsors and equivariant GG-bundles.

Let YY be a smooth algebraic curve over \mathbb{C} and let Γ\Gamma be a cyclic group of order dd together with an action on YY. Denote by RR a finite set of points of YY (also a divisor), such that the stabilizer group Γy\Gamma_{y} is nontrivial for any yRy\in R. Let GG be a connected complex reductive group with maximal torus TT and root system \mathcal{R}. Fix a rational cocharacter

θY(T)\theta\in Y(T)\otimes\mathbb{Q}

and let d>0d>0 be the smallest integer so that dθY(T).d\,\theta\in Y(T). Define

Δ:=dθY(T).\Delta:=d\,\theta\in Y(T).

Assume that the group Γ\Gamma acts on the formal disc 𝔻y\mathbb{D}_{y} via

γω=ζω,\gamma\cdot\omega=\zeta\,\omega,

with ζ\zeta a primitive ddth root of unity, and that the representation

ρ:ΓTG,ρ(γ)=Δ(ζ)\rho:\Gamma\to T\subset G,\quad\rho(\gamma)=\Delta(\zeta)

has been defined. As so we may define the subgroup of GG fixed by the Γ\Gamma-action,

GΓ:=CG(ρ(γ))={gGρ(γ)gρ(γ)1=g}.G^{\Gamma}:=C_{G}\bigl{(}\rho(\gamma)\bigr{)}=\{g\in G\mid\rho(\gamma)\,g\,\rho(\gamma)^{-1}=g\}.

In what follows the associated parahoric subgroup GθG_{\theta} (with pro-unipotent radical Gθ+G_{\theta}^{+}) has (generalized) Levi subgroup

Lθ=T(k),Ur(k),rθ.L_{\theta}=\langle T(k),U_{r}(k),r\in\mathcal{R}_{\theta}\rangle.

Moreover, set

θ:={rr(θ)}.\mathcal{R}_{\theta}:=\{r\in\mathcal{R}\mid r(\theta)\in\mathbb{Z}\}.

When θ\theta is small, i.e. |r(θ)|<1|r(\theta)|<1 for all rr, one recovers the classical Levi subgroup.

Lemma 3.4 (Centralizer and Levi Factor).

We have

CG(ρ(γ))=Lθ=T(k),Ur(k)rθ.C_{G}\bigl{(}\rho(\gamma)\bigr{)}=L_{\theta}=\langle T(k),\,U_{r}(k)\mid r\in\mathcal{R}_{\theta}\rangle.
Proof.

An element gGg\in G commutes with ρ(γ)\rho(\gamma) if and only if

ρ(γ)gρ(γ)1=g.\rho(\gamma)\,g\,\rho(\gamma)^{-1}=g.

Since ρ(γ)T\rho(\gamma)\in T and TT is abelian, every tTt\in T lies in CG(ρ(γ))C_{G}\bigl{(}\rho(\gamma)\bigr{)}. For a unipotent element ur(x)Ur(k)u_{r}(x)\in U_{r}(k), the conjugation formula yields

ρ(γ)ur(x)ρ(γ)1=ur(ζdr(θ)x).\rho(\gamma)\,u_{r}(x)\,\rho(\gamma)^{-1}=u_{r}\Bigl{(}\zeta^{d\,r(\theta)}\,x\Bigr{)}.

Thus, ur(x)u_{r}(x) commutes with ρ(γ)\rho(\gamma) if and only if ζdr(θ)=1\zeta^{d\,r(\theta)}=1. Since ζ\zeta is a primitive ddth root of unity, this holds if and only if

dr(θ)dr(θ).d\,r(\theta)\in d\,\mathbb{Z}\quad\Longleftrightarrow\quad r(\theta)\in\mathbb{Z}.

Hence,

CG(ρ(γ))=T(k),Ur(k)rθ,C_{G}\bigl{(}\rho(\gamma)\bigr{)}=\langle T(k),\,U_{r}(k)\mid r\in\mathcal{R}_{\theta}\rangle,

which by definition is equal to LθL_{\theta}. ∎

Theorem 3.5 (Equivalence of DD-Level Structures and Equivariant Maps).

Let xDx\in D be a point on a smooth projective curve XX with formal disc 𝔻x\mathbb{D}_{x}. Suppose that \mathcal{E} is a parahoric 𝒢𝛉\mathcal{G}_{\boldsymbol{\theta}}-torsor equipped with a DD-level structure at xx, i.e. a section

ηx:𝔻xGθx/Gθx+.\eta_{x}:\mathbb{D}_{x}\to G_{\theta_{x}}/G_{\theta_{x}}^{+}.

Let

p:𝔻y𝔻xp:\mathbb{D}_{y}\to\mathbb{D}_{x}

be the Galois covering with group Γ\Gamma (acting by γω=ζω\gamma\cdot\omega=\zeta\,\omega). Then, the level structure ηx\eta_{x} (which, via the identification Gθx/Gθx+LθxG_{\theta_{x}}/G_{\theta_{x}}^{+}\cong L_{\theta_{x}}, takes values in LθxL_{\theta_{x}}) is equivalent to a Γ\Gamma-invariant map

η~:𝔻yG\tilde{\eta}:\mathbb{D}_{y}\to G

satisfying

η~(γy)=η~(y),for all γΓ,\tilde{\eta}(\gamma\cdot y)=\tilde{\eta}(y),\quad\text{for all }\gamma\in\Gamma,

so that the image of η~\tilde{\eta} lies in GΓLθG^{\Gamma}\cong L_{\theta}.

Proof.

Since Gθx/Gθx+LθxG_{\theta_{x}}/G_{\theta_{x}}^{+}\cong L_{\theta_{x}} by definition (see Section 2.3), a DD-level structure ηx\eta_{x} provides a reduction of the structure group of \mathcal{E} over 𝔻x\mathbb{D}_{x} to LθxL_{\theta_{x}}. By standard descent theory for the Galois covering p:𝔻y𝔻xp:\mathbb{D}_{y}\to\mathbb{D}_{x}, such a reduction is equivalent to the existence of a unique Γ\Gamma-invariant lift

η~:𝔻yG\tilde{\eta}:\mathbb{D}_{y}\to G

satisfying η~(γy)=η~(y)\tilde{\eta}(\gamma\cdot y)=\tilde{\eta}(y), for all γΓ\gamma\in\Gamma. Conversely, any such Γ\Gamma-invariant map descends to a section ηx:𝔻xGΓLθx\eta_{x}:\mathbb{D}_{x}\to G^{\Gamma}\cong L_{\theta_{x}}, thereby defining a DD-level structure. ∎

In the light of the previous theorem, we now introduce the following:

Definition 3.6 (Equivariant DD-level structure).

We shall call a Γ\Gamma-equivariant DD-level structure on a Γ\Gamma-principal bundle to be a Γ\Gamma-invariant map

η~:𝔻yG\tilde{\eta}:\mathbb{D}_{y}\to G

satisfying

η~(γy)=η~(y),for all γΓ,\tilde{\eta}(\gamma\cdot y)=\tilde{\eta}(y),\quad\text{for all }\gamma\in\Gamma,

so that the image of η~\tilde{\eta} lies in GΓLθxG^{\Gamma}\cong L_{\theta_{x}}.

3.3. Moduli space

There is a correspondence between a (semi)stable parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor with DD-level structure and a (semi)stable equivariant (Γ,G)(\Gamma,G)-bundle equipped with an equivariant DD-level structure. Indeed, in the light of Definition 3.3, the (semi)stability of \mathcal{E} equipped with a DD-level structure η\eta amounts to the (semi)stability of \mathcal{E}, and from [24, Theorem 4.12] (following [2, Theorem 6.3.5]) we have that a parahoric torsor is (semi)stable if and only if the corresponding equivariant GG-bundle is.

The local type of EE at yy is defined to be the conjugacy class (in GG) of the representation ρy\rho_{y}. We denote by 𝝉\boldsymbol{\tau} the collection of such local types at all points of YY lying over the branch locus of the covering YXY\to X.

A standard approach (following Balaji–Seshadri [2, Section 8]) is to first fix a faithful representation

GGL(n),G\hookrightarrow\mathrm{GL}(n),

and then consider a parameter scheme Q(Γ,GL(n))𝝉Q^{\boldsymbol{\tau}}_{(\Gamma,\mathrm{GL}(n))} which classifies Γ\Gamma-equivariant vector bundles on a suitable ramified cover YY (of XX) that are semistable and of fixed local type 𝝉\boldsymbol{\tau}. In particular, the points of Q(Γ,GL(n))𝝉Q^{\boldsymbol{\tau}}_{(\Gamma,\mathrm{GL}(n))} correspond to Γ\Gamma-semistable principal (Γ,GL(n))(\Gamma,\mathrm{GL}(n))-bundles.

Next, one defines

Q(Γ,G)𝝉Q(Γ,GL(n))𝝉Q^{\boldsymbol{\tau}}_{(\Gamma,G)}\subset Q^{\boldsymbol{\tau}}_{(\Gamma,\mathrm{GL}(n))}

to be the subscheme parameterizing those bundles which admit a reduction of structure group to GG. This subscheme has the local universal property for families of semistable (Γ,G)(\Gamma,G)-bundles of local type 𝝉\boldsymbol{\tau}.

The extra data of a DD-level structure is now imposed as follows. For each xDx\in D, consider the formal disc 𝔻y\mathbb{D}_{y} over yy and the fixed equivariant map

η~x:𝔻yG\tilde{\eta}_{x}:\mathbb{D}_{y}\to G

(which, by definition, has image in GΓxG^{\Gamma_{x}}, where Γx\Gamma_{x} denotes the Galois group acting for the particular parahoric point xx). The section η~x\tilde{\eta}_{x} being a trivialization over 𝔻y\mathbb{D}_{y} of the universal family of (Γ,G)(\Gamma,G)-bundles is a closed condition. Hence, one obtains a closed (or locally closed) subscheme

Q(Γ,G,D)𝝉Q(Γ,G)𝝉Q^{\boldsymbol{\tau}}_{(\Gamma,G,D)}\subset Q^{\boldsymbol{\tau}}_{(\Gamma,G)}

parameterizing those (Γ,G)(\Gamma,G)-bundles together with the prescribed DD-level structure.

Finally, one forms the good quotient (in the sense of Geometric Invariant Theory) by a suitable reductive group \mathcal{H} acting on Q(Γ,G,D)𝝉Q^{\boldsymbol{\tau}}_{(\Gamma,G,D)} to obtain the coarse moduli space

𝒰(X,𝒢𝜽):=Q(Γ,G,D)𝝉//.\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}}):=Q^{\boldsymbol{\tau}}_{(\Gamma,G,D)}//\mathcal{H}.

Remember that the moduli space of parahoric 𝒢θ\mathcal{G}_{\theta}-torsors was constructed in [2, Section 8] via the correspondence to Γ\Gamma-equivariant GG-bundles. Since the extra DD-level structure on a parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor (equivalently, the equivariant DD-level structure on a Γ\Gamma-equivariant GG-bundle) is a rigid condition, all the standard arguments (regarding the existence of the quotient, its irreducibility, normality, and projectivity) carry through unchanged.

Let S×X\mathcal{E}\to S\times X be a flat family of Γ\Gamma-equivariant principal GG-bundles on XX, each equipped with its prescribed Γ\Gamma-equivariant DD-level structure η~\tilde{\eta}. Suppose there is a reference (Γ,G)(\Gamma,G)-bundle 0\mathcal{E}_{0} (with the same level DD-level structure η~\tilde{\eta}) such that, for a dense open subset USU\subset S, every fiber s\mathcal{E}_{s} with sUs\in U is Γ\Gamma-equivariantly isomorphic to 0\mathcal{E}_{0} in a way compatible with the given level structure. Then we say that \mathcal{E} is S-equivalent to 0\mathcal{E}_{0}.

We summarize the previous analysis to the following:

Theorem 3.7 (Existence of Moduli Space).

Let XX be a smooth projective curve over \mathbb{C} and DXD\subset X a reduced effective divisor. Let 𝒢𝛉\mathcal{G}_{\boldsymbol{\theta}} be the parahoric Bruhat–Tits group scheme on XX corresponding to a collection 𝛉:={θx}xD{\boldsymbol{\theta}}:=\{\theta_{x}\}_{x\in D} of rational weights. Then, the moduli functor which assigns to any scheme SS the set of SS-equivalent classes of semistable parahoric 𝒢𝛉\mathcal{G}_{\boldsymbol{\theta}}-torsors on XX with fixed DD-level structure is corepresented by an irreducible, normal, projective variety.

We thus introduce the following

Definition 3.8.

We will denote by 𝒰(X,𝒢𝜽),\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}}), the moduli space of stable parahoric 𝒢𝛉\mathcal{G}_{\boldsymbol{\theta}}-torsors over XX with a DD-level structure in the sense of Definition 3.3.

Remark 3.9.

One can also see how the definition of a parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor with DD-level structure reduces to Seshadri’s definition of a DD-level structure on a vector bundle from [32, Définition 1, p. 92]. Firstly, in the absence of a parahoric structure and for structure group G=GL(n,)G=\mathrm{GL}(n,\mathbb{C}), a parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor reduces to a vector bundle. Now, a DD-level structure at a point xDx\in D is a section ηx:𝔻xGθx/Gθx+\eta_{x}:\mathbb{D}_{x}\to G_{\theta_{x}}/G_{\theta_{x}}^{+}, for each xDx\in D. Restricting to the fiber over the point xx, then a level structure is just the trivialization of the fiber over xx. This is exactly the notion as in Seshadri’s definition as an isomorphism η:E|Di=1r𝒪D\eta:E|_{D}\to\oplus_{i=1}^{r}\mathcal{O}_{D} for a rank rr vector bundle EE over XX. Indeed, locally for each xDx\in D, when GθxG_{\theta_{x}} is GL([[t]])\mathrm{GL}(\mathbb{C}[\![t]\!]), for local coordinate tt, and Gθx+G_{\theta_{x}}^{+} is trivial, then the section ηx\eta_{x} is a section of GL([[t]])\mathrm{GL}(\mathbb{C}[\![t]\!]), thus after extending to a formal disk 𝔻x\mathbb{D}_{x}, the section ηx\eta_{x} is just an isomorphism of the fiber above xx.

4. Deformations of DD-level structures and singularities of parahoric torsors

The deformation theory of DD-level structures is studied in this Section. We study the tangent space of the moduli space 𝒰(X,𝒢𝜽)\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}}) and its singular points.

4.1. Deformation of DD-level structure

We start with the following proposition.

Proposition 4.1.

Let 𝒰(X,𝒢𝛉)\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}}) be the moduli space of stable parahoric 𝒢𝛉\mathcal{G}_{\boldsymbol{\theta}}-torsors with a DD-level structure. For any representative [(,η)]𝒰(X,𝒢θ)[(\mathcal{E},\eta)]\in\mathcal{U}(X,\mathcal{G}_{\theta}), the tangent space is

T[(,η)]𝒰(X,𝒢𝜽)H1(X,(𝔤)η).T_{[(\mathcal{E},\eta)]}\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}})\cong H^{1}(X,\mathcal{E}(\mathfrak{g})_{\eta}).
Proof.

Choose a sufficiently fine open cover {Ui}iI\{U_{i}\}_{i\in I} of XX where \mathcal{E} trivializes on each UiU_{i}. If UiU_{i} contains a marked point xjDx_{j}\in D, we need to ensure that the trivialization near xjx_{j} is compatible with the parahoric reduction ηxj\eta_{x_{j}}. Concretely, over UiU_{i}, we have |UiUi×G\mathcal{E}|_{U_{i}}\cong U_{i}\times G. On each non-empty intersection UiUjU_{i}\cap U_{j}, the torsor is determined by transition functions gij:UiUjGg_{ij}\colon U_{i}\cap U_{j}\to G satisfying gijgjk=gikg_{ij}g_{jk}=g_{ik} on triple overlaps.

An infinitesimal deformation of \mathcal{E} involves deforming gijg_{ij} to first order in a parameter ε\varepsilon with ε2=0\varepsilon^{2}=0 as

gijgij(Id+εαij),g_{ij}\mapsto g_{ij}(\mathrm{Id}+\varepsilon\,\alpha_{ij}),

where each αij\alpha_{ij} is a section of the adjoint bundle (𝔤)\mathcal{E}(\mathfrak{g}) over UiUjU_{i}\cap U_{j}. Trivializing (𝔤)\mathcal{E}(\mathfrak{g}) on UiU_{i} identifies these with elements of Γ(Ui,𝔤)\Gamma(U_{i},\mathfrak{g}). The cocycle condition on triple overlaps becomes

αij+Adgij(αjk)=αik,\alpha_{ij}+\mathrm{Ad}_{g_{ij}}(\alpha_{jk})=\alpha_{ik},

so {αij}\{\alpha_{ij}\} forms a Čech 1-cocycle in Zˇ1({Ui},(𝔤))\check{Z}^{1}(\{U_{i}\},\mathcal{E}(\mathfrak{g})).

Two cocycles {αij}\{\alpha_{ij}\} and {αij}\{\alpha^{\prime}_{ij}\} differ by an infinitesimal gauge transformation if there exists a Čech 0-cochain {βi}\{\beta_{i}\} in (𝔤)\mathcal{E}(\mathfrak{g}) such that

αij=αij+βjAdgij(βi).\alpha^{\prime}_{ij}=\alpha_{ij}+\beta_{j}-\mathrm{Ad}_{g_{ij}}(\beta_{i}).

In Čech terms, {αij}={αij}+δ({βi})\{\alpha^{\prime}_{ij}\}=\{\alpha_{ij}\}+\delta(\{\beta_{i}\}), where δ\delta is the Čech differential. The infinitesimal deformations of \mathcal{E} are classified by

Hˇ1({Ui},(𝔤))H1(X,(𝔤)).\check{H}^{1}(\{U_{i}\},\mathcal{E}(\mathfrak{g}))\cong H^{1}(X,\mathcal{E}(\mathfrak{g})).

To incorporate the DD-level structure, each marked point xjDx_{j}\in D imposes a parahoric reduction near xjx_{j}. Gauge transformations must preserve this reduction, so if xjUix_{j}\in U_{i}, then βi\beta_{i} must lie in the subalgebra 𝔤θxj+\mathfrak{g}_{\theta_{x_{j}}}^{+} stabilizing the parahoric flag. Globally, this is encoded by the subsheaf (𝔤)η(𝔤)\mathcal{E}(\mathfrak{g})_{\eta}\subset\mathcal{E}(\mathfrak{g}), whose sections preserve the parahoric structure at all points of DD. The allowed infinitesimal gauge transformations form Cˇ0({Ui},(𝔤)η)\check{C}^{0}(\{U_{i}\},\mathcal{E}(\mathfrak{g})_{\eta}).

Thus, the space of infinitesimal deformations of a pair (,η)(\mathcal{E},\eta) is given by Čech 1-cocycles in (𝔤)\mathcal{E}(\mathfrak{g}) modulo these restricted 0-cochains:

Hˇ1({Ui},(𝔤)η){αij}Zˇ1((𝔤))δ({βi}) where {βi}Cˇ0((𝔤)η).\check{H}^{1}(\{U_{i}\},\mathcal{E}(\mathfrak{g})_{\eta})\cong\frac{\{\alpha_{ij}\}\in\check{Z}^{1}(\mathcal{E}(\mathfrak{g}))}{\delta(\{\beta_{i}\})\text{ where }\{\beta_{i}\}\in\check{C}^{0}(\mathcal{E}(\mathfrak{g})_{\eta})}.

Since {Ui}\{U_{i}\} is sufficiently fine, this is identified with H1(X,(𝔤)η)H^{1}(X,\mathcal{E}(\mathfrak{g})_{\eta}). By the standard correspondence between first-order deformations and the tangent space, we have

T[(,η)]𝒰(X,𝒢𝜽)H1(X,(𝔤)η),T_{[(\mathcal{E},\eta)]}\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}})\cong H^{1}(X,\mathcal{E}(\mathfrak{g})_{\eta}), (4.1)

as claimed. ∎

Taking dual vector spaces, observe that locally at a formal disk around each point xDx\in D, it is

(𝔤)η|𝔻x𝔤θ+=𝔤θ𝒪(x).\mathcal{E}(\mathfrak{g})_{\eta}|_{\mathbb{D}_{x}}\cong\mathfrak{g}_{\theta}^{+}=\mathfrak{g}_{\theta}^{\perp}\otimes\mathcal{O}(-x).

Dualizing, we obtain

(𝔤)η|𝔻x𝔤θ𝒪(x),\mathcal{E}(\mathfrak{g})_{\eta}^{\vee}|_{\mathbb{D}_{x}}\cong\mathfrak{g}_{\theta}\otimes\mathcal{O}(x),

and by Serre duality we have

H1(X,(𝔤)η)H0(X,(𝔤)K(D)).H^{1}(X,\mathcal{E}(\mathfrak{g})_{\eta})^{\vee}\cong H^{0}(X,\mathcal{E}(\mathfrak{g})\otimes K(D)).

Therefore, the tangent space corresponds to the space of logarithmic Higgs fields as introduced in [24]:

H0(X,(𝔤)K(D)).H^{0}(X,\mathcal{E}(\mathfrak{g})\otimes K(D)).

We recall the following:

Definition 4.2.

[24, Definition 3.1] A logahoric 𝒢𝛉\mathcal{G}_{\boldsymbol{\theta}}-Higgs torsor on a smooth complex algebraic curve XX is defined as a pair (,φ)(\mathcal{E},\varphi), where

  • \mathcal{E} is a parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor on XX;

  • φH0(X,(𝔤)K(D))\varphi\in H^{0}(X,\mathcal{E}(\mathfrak{g})\otimes K(D)) is a section called a logarithmic Higgs field.

In [24], a moduli space of logahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-Higgs torsors over a smooth complex algebraic curve was constructed as a quasi-projective variety. We will denote this moduli space by H(X,𝒢𝜽)\mathcal{M}_{H}(X,\mathcal{G}_{\boldsymbol{\theta}}).

4.2. Regular Stability and Singularities for Parahoric Torsors

We next study the singular points of the moduli space 𝒰(X,𝒢𝜽)\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}}). We begin with the following:

Definition 4.3 (Regularly stable parahoric torsor).

Let GG be a connected complex reductive Lie group with center Z(G)Z(G). A stable parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor \mathcal{E} over XX is said to be regularly stable if Aut()=Z(G).\mathrm{Aut}(\mathcal{E})=Z(G). Equivalently, a stable parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor with DD-level structure is regularly stable precisely when the underlying torsor has no extra automorphisms apart from the ones in Z(G)Z(G). We shall denote the moduli of regularly stable parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsors with DD-level structures by 𝒰rs(X,𝒢𝜽)\mathcal{U}^{rs}(X,\mathcal{G}_{\boldsymbol{\theta}}). This is an open subvariety of 𝒰(X,𝒢𝜽)\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}}).

Remark 4.4.

Note that the notion of a regularly stable parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor extends the notion of a δ\delta-stable vector bundle of Seshadri [32], as well as the notion of a regularly stable principal GG-bundle of Biswas–Hoffmann [6].

Proposition 4.5.

Let [(,η)][(\mathcal{E},\eta)] represent an isomorphism class of stable but not regularly stable 𝒢𝛉\mathcal{G}_{\boldsymbol{\theta}}-torsors on XX with DD-level structure. Then the corresponding point [(,η)]𝒰(X,𝒢𝛉)[(\mathcal{E},\eta)]\;\in\;\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}}) is a singular point.

Proof.

The deformation theory for parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsors, combined with Luna’s étale slice theorem, identifies a Zariski neighborhood of [(,η)][(\mathcal{E},\eta)] in 𝒰(X,𝒢𝜽)\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}}) with the GIT quotient

H1(X,(𝔤𝜽)η)/Aut()H^{1}\bigl{(}X,\,\mathcal{E}({\mathfrak{g}_{\boldsymbol{\theta}}})_{\eta}\bigr{)}\;\big{/}\;\mathrm{Aut}(\mathcal{E})

near the origin.

Because \mathcal{E} is stable but not regularly stable, there exists a non-trivial automorphism of finite order

fAut()Z(G).f\in\mathrm{Aut}(\mathcal{E})\setminus Z(G).

The local structure of the moduli space is therefore a quotient of the vector space V=H1(X,(𝔤𝜽)η)V=H^{1}\bigl{(}X,\,\mathcal{E}({\mathfrak{g}_{\boldsymbol{\theta}}})_{\eta}\bigr{)} by the action of the finite non-trivial group f\langle f\rangle.

We apply the Chevalley–Shephard–Todd singularity criterion: Let a finite group HH act linearly on a complex vector space VV. If for some non-trivial hHh\in H, the fixed-point subspace VhV^{h} has codimension 2\geq 2, then the affine quotient V/HV/H is singular at the origin.

The automorphism ff acts on the torsor \mathcal{E}. This induces a semi-simple linear action on the fiber of the associated Lie algebra bundle over any generic point xXDx\in X\setminus D, which is the vector space 𝔤=Lie(G)\mathfrak{g}=\mathrm{Lie}(G). This action on 𝔤\mathfrak{g} induces an eigenspace decomposition 𝔤=𝔤1𝔤1\mathfrak{g}=\mathfrak{g}_{1}\oplus\mathfrak{g}_{\neq 1}, where 𝔤1\mathfrak{g}_{1} is the trivial eigenspace (the subspace fixed by ff). Because fZ(G)f\notin Z(G), its adjoint action on 𝔤\mathfrak{g} is non-trivial. The root spaces of 𝔤\mathfrak{g} on which ff acts non-trivially come in pairs (corresponding to a root and its negative). This pairing ensures that the vector space 𝔤1\mathfrak{g}_{\neq 1} has dimension at least 2, i.e., dim(𝔤1)2\dim_{\mathbb{C}}(\mathfrak{g}_{\neq 1})\geq 2.

Since ff is an automorphism of the parahoric torsor, its action preserves the parahoric structure at each point xiDx_{i}\in D. Consequently, the decomposition of 𝔤\mathfrak{g} extends to a decomposition of the entire sheaf of Lie algebras 𝔤𝜽{\mathfrak{g}_{\boldsymbol{\theta}}} associated with the group scheme 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}, thus giving a direct sum of sheaves

𝔤𝜽=𝔤𝜽1𝔤𝜽1.{\mathfrak{g}_{\boldsymbol{\theta}}}={\mathfrak{g}_{\boldsymbol{\theta}}}_{1}\oplus{\mathfrak{g}_{\boldsymbol{\theta}}}_{\neq 1}.

This decomposition lifts to the associated coherent sheaf on XX

(𝔤𝜽)η=(𝔤𝜽1)η(𝔤𝜽1)η,\mathcal{E}({\mathfrak{g}_{\boldsymbol{\theta}}})_{\eta}=\mathcal{E}({\mathfrak{g}_{\boldsymbol{\theta}}}_{1})_{\eta}\oplus\mathcal{E}({\mathfrak{g}_{\boldsymbol{\theta}}}_{\neq 1})_{\eta},

which in turn induces a direct sum decomposition on the first cohomology group

H1(X,(𝔤𝜽)η)H1(X,(𝔤𝜽1)η)H1(X,(𝔤𝜽1)η).H^{1}\bigl{(}X,\,\mathcal{E}({\mathfrak{g}_{\boldsymbol{\theta}}})_{\eta}\bigr{)}\cong H^{1}\bigl{(}X,\,\mathcal{E}({\mathfrak{g}_{\boldsymbol{\theta}}}_{1})_{\eta}\bigr{)}\oplus H^{1}\bigl{(}X,\,\mathcal{E}({\mathfrak{g}_{\boldsymbol{\theta}}}_{\neq 1})_{\eta}\bigr{)}.

The action of ff is trivial on the first summand and non-trivial on the second. Thus, the fixed-point subspace is precisely VfH1(X,(𝔤𝜽1)η)V^{f}\cong H^{1}\bigl{(}X,\,\mathcal{E}({\mathfrak{g}_{\boldsymbol{\theta}}}_{1})_{\eta}\bigr{)}.

The codimension of this fixed-point subspace is therefore dimH1(X,(𝔤𝜽1)η)\dim H^{1}\bigl{(}X,\,\mathcal{E}({\mathfrak{g}_{\boldsymbol{\theta}}}_{\neq 1})_{\eta}\bigr{)}. Let us denote the sheaf :=(𝔤𝜽1)η\mathcal{F}:=\mathcal{E}({\mathfrak{g}_{\boldsymbol{\theta}}}_{\neq 1})_{\eta}. The rank of \mathcal{F} (the dimension of its stalk at a generic point of XX) is dim(𝔤1)2\dim_{\mathbb{C}}(\mathfrak{g}_{\neq 1})\geq 2.

Since the torsor \mathcal{E} is stable, the associated sheaf (𝔤𝜽)\mathcal{E}({\mathfrak{g}_{\boldsymbol{\theta}}}) is semistable and has parahoric degree 0. This property is inherited by its direct summands, so \mathcal{F} is a semistable coherent sheaf of rank 2\geq 2 and degree 0. For such a sheaf on a projective curve, it is a standard result that its first cohomology group has dimension at least 2. A more detailed analysis using Serre duality shows that dimH1(X,)=dimH0(X,KX)\dim H^{1}(X,\mathcal{F})=\dim H^{0}(X,\mathcal{F}^{*}\otimes K_{X}), and the properties of KX\mathcal{F}^{*}\otimes K_{X} guarantee that this dimension is at least 2 across all genera for a non-regularly stable torsor.

Therefore, the codimension of the fixed subspace VfV^{f} in VV is at least 2. The singularity criterion applies, and we conclude that the quotient is singular at the origin. Hence, the point [(,η)][(\mathcal{E},\eta)] is a singular point of the moduli space 𝒰(X,𝒢𝜽)\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}}). ∎

5. Poisson action and moment map

The first part of this Section includes the necessary preliminaries from Symplectic and Poisson geometry over smooth algebraic varieties that will be useful for establishing our main results. We then introduce a level group and study its action on the moduli space 𝒰(X,𝒢𝜽)\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}}) of stable parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsors with DD-level structure. This action is shown to be inducing a Poisson action on the cotangent T𝒰(X,𝒢𝜽)T^{*}\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}}), thus providing a canonical moment map.

5.1. Hamiltonian group actions on smooth algebraic varieties

In this subsection, we gather those basic notions from Poisson geometry and completely integrable systems which will be used in the rest of the article. Standard references in the context of smooth algebraic varieties include [34, Chapter I] or [15, Section 2.3].

Definition 5.1.

Let XX be a smooth algebraic variety of dimension nn. A symplectic form (or symplectic structure) on XX is an algebraic 2-form ωΓ(X,ΩX2)\omega\in\Gamma(X,\Omega_{X}^{2}) such that:

  1. (1)

    ω\omega is closed: dω=0d\omega=0, where dd is the exterior derivative.

  2. (2)

    ω\omega is non-degenerate: for every point xXx\in X, the map TxXTxXT_{x}X\rightarrow T_{x}^{*}X defined by vωx(v,)v\mapsto\omega_{x}(v,\cdot) is an isomorphism of vector spaces. This implies that nn must be even.

A smooth algebraic variety XX equipped with a symplectic form ω\omega is called a symplectic algebraic variety (X,ω)(X,\omega).

Definition 5.2.

Let XX be a smooth algebraic variety. A Poisson bracket on XX is a Lie bracket

{ , }:𝒪X×𝒪X𝒪X\displaystyle\left\{\text{ },\text{ }\right\}:\mathcal{O}_{X}\times\mathcal{O}_{X}\rightarrow\mathcal{O}_{X}

satisfying the Leibniz rule {f,gh}={f,g}h+g{f,h}\left\{f,gh\right\}=\left\{f,g\right\}h+g\left\{f,h\right\}, for f,g,h𝒪Xf,g,h\in\mathcal{O}_{X}. Poisson brackets bijectively correspond to bi-vector fields

ΠΓ(2TX)\Pi\in\Gamma\left({{\wedge}^{2}}TX\right)

such that

[Π,Π]=0.\left[\Pi,\Pi\right]=0.

The Poisson bracket {,}Π{{\left\{,\right\}}_{\Pi}} that corresponds to such a bi-vector field Π\Pi is given by {f,g}Π=Π(df,dg){{\left\{f,g\right\}}_{\Pi}}=\Pi\left(df,dg\right).

Remark 5.3.

Every symplectic algebraic variety (X,ω)(X,\omega) is naturally a Poisson algebraic variety. The non-degenerate 2-form ω\omega induces an isomorphism ω:TXΩX1\omega^{\sharp}:TX\rightarrow\Omega_{X}^{1} (the cotangent sheaf). Its inverse (ω)1:ΩX1TX(\omega^{\sharp})^{-1}:\Omega_{X}^{1}\rightarrow TX allows us to define the Poisson bi-vector ΠΓ(X,2TX)\Pi\in\Gamma(X,\wedge^{2}TX) by Π(α,β)=ω((ω)1(α),(ω)1(β))\Pi(\alpha,\beta)=\omega((\omega^{\sharp})^{-1}(\alpha),(\omega^{\sharp})^{-1}(\beta)), for local sections α,β\alpha,\beta of ΩX1\Omega_{X}^{1}. The Poisson bracket is then given by {f,g}=Π(df,dg)=ω(Vf,Vg)\{f,g\}=\Pi(df,dg)=\omega(V_{f},V_{g}), where VfV_{f} is the Hamiltonian vector field of ff (see Definition 5.6 below). The condition dω=0d\omega=0 ensures that [Π,Π]=0[\Pi,\Pi]=0.

Example 5.4 (Cotangent Bundles).

Let QQ be a smooth algebraic variety of dimension mm. Its cotangent bundle X=TQX=T^{*}Q is a smooth algebraic variety of dimension 2m2m. Let π:TQQ\pi:T^{*}Q\to Q be the canonical projection. There exists a canonical 1-form θΓ(TQ,ΩTQ1)\theta\in\Gamma(T^{*}Q,\Omega_{T^{*}Q}^{1}), called the Liouville form (or tautological 1-form). If (q1,,qm)(q_{1},\ldots,q_{m}) are local coordinates on an open subset UQU\subseteq Q, and (p1,,pm)(p_{1},\ldots,p_{m}) are the corresponding fiber coordinates on TUU×kmT^{*}U\cong U\times k^{m}, then θ=i=1mpidqi\theta=\sum_{i=1}^{m}p_{i}dq_{i}. The 2-form ω=dθ=i=1mdqidpi\omega=-d\theta=\sum_{i=1}^{m}dq_{i}\wedge dp_{i} is a symplectic form on TQT^{*}Q. This is known as the canonical symplectic structure on the cotangent bundle (cf. [1]). Thus, (TQ,ω)(T^{*}Q,\omega) is a symplectic algebraic variety.

Example 5.5 (Kostant–Kirillov structures).

For a Lie group GG with Lie algebra 𝔤\mathfrak{g}, there exists a canonical Poisson structure on the dual vector space 𝔤\mathfrak{g}^{*} called the Kostant–Kirillov Poisson structure defined by the bracket

{F,G}(ξ):=ξ,[dξF,dξG],\{F,G\}(\xi):=\langle\xi,[d_{\xi}F,d_{\xi}G]\rangle,

for F,GC(𝔤)F,G\in C^{\infty}(\mathfrak{g}^{*}). This is obtained by extending on 𝔤\mathfrak{g}^{*} the symplectic structures on coadjoint orbits of 𝔤\mathfrak{g}. Note here that dξFd_{\xi}F is identified with an element of 𝔤=𝔤\mathfrak{g}=\mathfrak{g}^{**}. On the dual 𝔤\mathfrak{g}^{*} the symplectic leaves of the Kostant–Kirillov Poisson structure are precisely the coadjoint orbits. Moreover, the rank of 𝔤\mathfrak{g} is equal to the smallest codimension of a coadjoint orbit.

Definition 5.6.

Let (X,{,})(X,\{\cdot,\cdot\}) be a Poisson algebraic variety (or (X,ω)(X,\omega) be a symplectic algebraic variety). For a regular function h𝒪Xh\in\mathcal{O}_{X}, called a Hamiltonian function, the Hamiltonian vector field VhV_{h} is the unique vector field such that Vh(f)={f,h}V_{h}(f)=\{f,h\}, for all f𝒪Xf\in\mathcal{O}_{X}. If (X,ω)(X,\omega) is symplectic, then VhV_{h} is equivalently defined by the condition iVhω=dhi_{V_{h}}\omega=-dh, where iVhωi_{V_{h}}\omega is the interior product of VhV_{h} with ω\omega.

Definition 5.7.

Let GG be an algebraic group with Lie algebra 𝔤=TeG\mathfrak{g}=T_{e}G. Let GG act on a symplectic algebraic variety (X,ω)(X,\omega) via a morphism Φ:G×XX\Phi:G\times X\rightarrow X. For each A𝔤A\in\mathfrak{g}, the action induces a fundamental vector field AXΓ(X,TX)A_{X}\in\Gamma(X,TX) defined at xXx\in X by AX(x)=ddt|t=0(exp(tA)x)A_{X}(x)=\frac{d}{dt}|_{t=0}(\exp(-tA)\cdot x), or more algebraically, as the image of AA under the map 𝔤Γ(X,TX)\mathfrak{g}\to\Gamma(X,TX) induced by the action.

The action of GG on (X,ω)(X,\omega) is called Hamiltonian if:

  1. (1)

    For every A𝔤A\in\mathfrak{g}, the fundamental vector field AXA_{X} is Hamiltonian. That is, there exists a regular function hA𝒪Xh_{A}\in\mathcal{O}_{X} such that AX=VhAA_{X}=V_{h_{A}} (or, equivalently, iAXω=dhAi_{A_{X}}\omega=-dh_{A}).

  2. (2)

    There exists a GG-equivariant morphism μ:X𝔤\mu:X\rightarrow\mathfrak{g}^{*} (where 𝔤\mathfrak{g}^{*} is the dual of the Lie algebra) called the moment map (or momentum map) such that hA(x)=μ(x),Ah_{A}(x)=\langle\mu(x),A\rangle, for all A𝔤A\in\mathfrak{g} and xXx\in X. The pairing ,\langle\cdot,\cdot\rangle is the natural pairing between 𝔤\mathfrak{g}^{*} and 𝔤\mathfrak{g}, and GG-equivariance here means that μ(gx)=Adg(μ(x))\mu(g\cdot x)=\mathrm{Ad}_{g}^{*}(\mu(x)), for all gG,xXg\in G,x\in X, where Ad\mathrm{Ad}^{*} is the coadjoint action of GG on 𝔤\mathfrak{g}^{*}.

Often, an additional condition is imposed: the map AhAA\mapsto h_{A} from 𝔤\mathfrak{g} to (𝒪X,{,})(\mathcal{O}_{X},\{\cdot,\cdot\}) is a Lie algebra homomorphism, i.e., {hA,hB}=h[A,B]\{h_{A},h_{B}\}=h_{[A,B]}, for all A,B𝔤A,B\in\mathfrak{g}.

Example 5.8 (Lifted Action on Cotangent Bundles).

Let GG be an algebraic group acting on a smooth algebraic variety QQ. This action can be lifted to an action on the cotangent bundle TQT^{*}Q. Let ρq:GQ\rho_{q}:G\to Q be the orbit map ggqg\mapsto g\cdot q, for qQq\in Q. For each αTqQ\alpha\in T_{q}^{*}Q, the induced action on (q,α)TQ(q,\alpha)\in T^{*}Q admits a moment map

μ:TQ𝔤\mu:T^{*}Q\longrightarrow\mathfrak{g}^{*}

characterized by

μ(q,α),A=α,AQ(q),for all A𝔤,\langle\mu(q,\alpha),A\rangle=\langle\alpha,A_{Q}(q)\rangle,\quad\text{for all }A\in\mathfrak{g},

where AQA_{Q} is the fundamental vector field on QQ generated by AA. Equivalently, using the differential of the orbit map dρq:𝔤TqQd\rho_{q}:\mathfrak{g}\to T_{q}Q, the moment map can be written as

μ(q,α)=(dρq)(α)𝔤.\mu(q,\alpha)=(d\rho_{q})^{*}(\alpha)\in\mathfrak{g}^{*}.

This lifted action on TQT^{*}Q (equipped with its canonical symplectic form) is automatically Hamiltonian.

Definition 5.9.

Let (X,ω)(X,\omega) be a symplectic algebraic variety of dimension 2m2m. A smooth subvariety LXL\subset X is called:

  • Isotropic if for every xLx\in L, the tangent space TxLT_{x}L is an isotropic subspace of TxXT_{x}X, that is, ωx(v,w)=0\omega_{x}(v,w)=0, for all v,wTxLv,w\in T_{x}L. This is equivalent to saying that the pullback of ω\omega to LL, ω|L\omega|_{L}, is zero.

  • Lagrangian if it is isotropic and its dimension is m=12dimXm=\frac{1}{2}\dim X.

For a Poisson algebraic variety (X,Π)(X,\Pi), an irreducible subvariety YXY\subset X is Lagrangian if it is generically a Lagrangian subvariety of a symplectic leaf of XX. More precisely, YY is contained in the closure of a symplectic leaf SXS\subset X, and YSY\cap S is a Lagrangian subvariety of SS (where SS is equipped with the symplectic structure induced by the bi-vector field Π\Pi).

Definition 5.10.

Let XX be a smooth Poisson algebraic variety. An algebraically completely integrable Hamiltonian system (often referred to as an algebraic integrable system) is typically given by a proper morphism H:XBH:X\to B to an algebraic variety BB of dimension m=(dimX)/2m=(\dim X)/2, such that:

  1. (1)

    The components H1,,HmH_{1},\dots,H_{m} of HH (if BkmB\subset k^{m}) are in involution, i.e., {Hi,Hj}=0\{H_{i},H_{j}\}=0, for all i,ji,j.

  2. (2)

    The generic fibers H1(b)H^{-1}(b), for bBb\in B, are Lagrangian subvarieties of XX.

  3. (3)

    These generic fibers are (open subsets of) abelian varieties, and the Hamiltonian vector fields VHiV_{H_{i}} are tangent to the fibers and correspond to translation-invariant vector fields on these abelian varieties.

More generally, an algebraically completely integrable Hamiltonian system structure on a family of abelian varieties H:XBH:X\to B is a Poisson structure on XX with respect to which H:XBH:X\to B is a Lagrangian fibration (meaning its generic fibers are Lagrangian). If XX is a smooth algebraic variety, BB an algebraic variety, ABA\subset B a proper closed subvariety, and H:XBH:X\to B a proper morphism such that the fibers over BAB\setminus A are (isomorphic to) abelian varieties, then a Poisson structure on XX defines an algebraically completely integrable Hamiltonian system if H:XBH:X\to B is a Lagrangian fibration over BAB\setminus A.

5.2. Action of GDG_{D} on Local Trivializations

To provide a more explicit understanding of a group action on the moduli space 𝒰(X,𝒢𝜽)\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}}) of parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsors with DD-level structure, we analyze the local description using open covers (patches) of the curve XX.

Let D={x1,x2,,xs}D=\{x_{1},x_{2},\ldots,x_{s}\} be the reduced effective divisor on XX, where each xix_{i} is a distinct point. Choose an open cover {Ui}i=0s\{U_{i}\}_{i=0}^{s} of XX such that:

  • U0=XDU_{0}=X\setminus D is the complement of the divisor DD.

  • For each 1js1\leq j\leq s, UjU_{j} is a small open disc 𝔻xj\mathbb{D}_{x_{j}} around the point xjDx_{j}\in D, equipped with a local coordinate zjz_{j} centered at xjx_{j}, i.e., zj(xj)=0z_{j}(x_{j})=0.

Over each 𝔻xj\mathbb{D}_{x_{j}}, the parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor \mathcal{E} trivializes

|𝔻xjGθxj.\mathcal{E}\big{|}_{\mathbb{D}_{x_{j}}}\cong G_{\theta_{x_{j}}}.

The transition functions on the overlaps UiUjU_{i}\cap U_{j} (1i,js1\leq i,j\leq s) encode the gluing data, respecting the parahoric reductions at each xjDx_{j}\in D.

Definition 5.11 (Level group).

Given the data introduced above, we define the level group

GD=Lθ1××LθsZ,G_{D}=\frac{L_{\theta_{1}}\times\cdots\times L_{\theta_{s}}}{Z},

where each LθjL_{\theta_{j}} is the Levi subgroup as in (2.2) at xjDx_{j}\in D, j=1,,sj=1,...,s and ZZ is the center of GG.

An element gGDg\in G_{D} can be represented by a tuple (g1,g2,,gs)Lθ1×Lθ2××Lθs(g_{1},g_{2},\ldots,g_{s})\in L_{\theta_{1}}\times L_{\theta_{2}}\times\cdots\times L_{\theta_{s}}, modulo the diagonal action of the center ZZ.

Proposition 5.12.

The action of gGDg\in G_{D} on the moduli space 𝒰(X,𝒢𝛉)\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}}) is induced locally by the action of each gjLθjg_{j}\in L_{\theta_{j}} on the corresponding local trivialization 𝔻xj×G\mathbb{D}_{x_{j}}\times G. Specifically, in the local coordinates zjz_{j} around each xjDx_{j}\in D, the action is given by:

g(ej(zj),φj(zj))=(gjej(zj),Ad(gj)φj(zj)),g\cdot(e_{j}(z_{j}),\varphi_{j}(z_{j}))=\left(g_{j}\cdot e_{j}(z_{j}),\,\mathrm{Ad}(g_{j})\cdot\varphi_{j}(z_{j})\right),

where:

  • ej(zj)e_{j}(z_{j}) represents a local section of the parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor over 𝔻xj\mathbb{D}_{x_{j}}.

  • φj(zj)\varphi_{j}(z_{j}) is a local section of (𝔤)K(D)\mathcal{E}(\mathfrak{g})\otimes K(D) over 𝔻xj\mathbb{D}_{x_{j}}, viewed as an element of the cotangent space T[,η]𝒰(X,𝒢𝜽)T^{*}_{[\mathcal{E},\eta]}\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}}).

  • Ad(gj)\mathrm{Ad}(g_{j}) denotes the adjoint action of gjg_{j} on 𝔤\mathfrak{g}.

Proof.

The action of GDG_{D} on the moduli space 𝒰(X,𝒢𝜽)\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}}) is defined globally by modifying the level structures at each point xjDx_{j}\in D. Locally, near each xjx_{j}, this corresponds to acting by the element gjLθjg_{j}\in L_{\theta_{j}} on the trivialization 𝔻xj×G\mathbb{D}_{x_{j}}\times G.

Given a local trivialization, an element gjLθjg_{j}\in L_{\theta_{j}} acts on a local section ej(zj)Ge_{j}(z_{j})\in G by multiplication:

ej(zj)gjej(zj).e_{j}(z_{j})\mapsto g_{j}\cdot e_{j}(z_{j}).

For the cotangent vectors, which are represented by Higgs fields φj(zj)(𝔤)K(D)\varphi_{j}(z_{j})\in\mathcal{E}(\mathfrak{g})\otimes K(D), the action is via the adjoint representation:

φj(zj)Ad(gj)φj(zj).\varphi_{j}(z_{j})\mapsto\mathrm{Ad}(g_{j})\cdot\varphi_{j}(z_{j}).

Since the action preserves the parahoric structure, gjg_{j} lies in the Levi subgroup LθjL_{\theta_{j}}, as so the level structure at xjx_{j} remains intact.

Finally, because GDG_{D} is defined modulo the center ZZ, the overall action respects the identification under ZZ. This completes the local description. ∎

We examine the infinitesimal action of the Lie algebra 𝔤D:=Lie(GD)\mathfrak{g}_{D}:=\text{Lie}(G_{D}) on the cotangent space.

Lemma 5.13.

Let (X1,X2,,Xs)𝔤D(X_{1},X_{2},\ldots,X_{s})\in\mathfrak{g}_{D}, where each Xj𝔩θjX_{j}\in\mathfrak{l}_{\theta_{j}} is an element of the Levi subalgebra corresponding to xjDx_{j}\in D. The infinitesimal action of (X1,X2,,Xs)(X_{1},X_{2},\ldots,X_{s}) on a cotangent vector φH0(X,(𝔤)K(D))\varphi\in H^{0}(X,\mathcal{E}(\mathfrak{g})\otimes K(D)) is given locally near each xjx_{j} by:

δXφj(zj)=[Xj,φj(zj)].\delta_{X}\varphi_{j}(z_{j})=[X_{j},\varphi_{j}(z_{j})].
Proof.

The infinitesimal action of Xj𝔩θjX_{j}\in\mathfrak{l}_{\theta_{j}} on the Higgs field φj(zj)\varphi_{j}(z_{j}) is induced by the adjoint action:

δXφj(zj)=ddε|ε=0Ad(exp(εXj))φj(zj)=[Xj,φj(zj)].\delta_{X}\varphi_{j}(z_{j})=\left.\frac{d}{d\varepsilon}\right|_{\varepsilon=0}\mathrm{Ad}\left(\exp(\varepsilon X_{j})\right)\varphi_{j}(z_{j})=[X_{j},\varphi_{j}(z_{j})].

This commutator arises naturally from the linearization of the adjoint action at the identity. ∎

We may now determine the action of the level group on parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsors.

Theorem 5.14.

(Freeness) The level group GDG_{D} acts freely on the regularly stable moduli space 𝒰rs(X,𝒢𝛉)\mathcal{U}^{rs}(X,\mathcal{G}_{\boldsymbol{\theta}}) of parahoric 𝒢𝛉\mathcal{G}_{\boldsymbol{\theta}}-torsors over XX with DD-level structure.

Proof.

Let (,η)(\mathcal{E},\eta) be a point in 𝒰(X,𝒢𝜽)\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}}). For each i=1,,si=1,...,s, the level structure is a reduction of structure group from GθxiG_{\theta_{x_{i}}} to Gθxi+G_{\theta_{x_{i}}}^{+} over a formal disk 𝔻xi\mathbb{D}_{x_{i}}, and the Levi factor LθiL_{\theta_{i}} naturally changes this local reduction. Hence i=1sLθi\prod_{i=1}^{s}L_{\theta_{i}} acts on such data, and this action factors through GD=(i=1sLθi)/ZG_{D}=(\prod_{i=1}^{s}L_{\theta_{i}})/Z because the center ZZ acts trivially.

To show the action is free, let gGDg\in G_{D} fix a point (,η)𝒰rs(X,𝒢𝜽)(\mathcal{E},\eta)\in\mathcal{U}^{rs}(X,\mathcal{G}_{\boldsymbol{\theta}}). Lift gg to some element g~=(g1,,gs)i=1sLθi\tilde{g}=(g_{1},\dots,g_{s})\in\prod_{i=1}^{s}L_{\theta_{i}}. The action of g~\tilde{g} on (,η)(\mathcal{E},\eta) by multiplying gig_{i} on the left of each ηi\eta_{i} yields a new pair (,η)(\mathcal{E}^{\prime},\eta^{\prime}). The fact that gg fixes (,η)(\mathcal{E},\eta) means there exists a global isomorphism of parahoric torsors ϕ:\phi:\mathcal{E}\to\mathcal{E}^{\prime} that is compatible with the level structures, i.e., ϕη=η\phi_{*}\eta=\eta^{\prime}.

The action of g~\tilde{g} is defined locally. On each formal disk 𝔻xi\mathbb{D}_{x_{i}}, the element giLθig_{i}\in L_{\theta_{i}} modifies the local trivialization of the torsor, which in turn defines the new torsor \mathcal{E}^{\prime} and level structure η\eta^{\prime} via gluing. The condition that g~\tilde{g} preserves the isomorphism class of (,η)(\mathcal{E},\eta) means that the newly constructed torsor (,η)(\mathcal{E}^{\prime},\eta^{\prime}) is isomorphic to the original one. This isomorphism ϕ:\phi:\mathcal{E}\to\mathcal{E}^{\prime}\cong\mathcal{E} is an automorphism of the parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor \mathcal{E}.

Because \mathcal{E} is regularly stable, all of its global automorphisms lie in the center ZGZ\subset G by definition.

If ϕZ\phi\in Z, then on i=1sLθi\prod_{i=1}^{s}L_{\theta_{i}}, the element g~\tilde{g} is also in ZZ (viewed diagonally). Thus gGDg\in G_{D} is the identity element of GDG_{D}. Therefore any gGDg\in G_{D} fixing a point of 𝒰(X,𝒢𝜽)\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}}) is the identity. This proves the action of GDG_{D} on 𝒰rs(X,𝒢𝜽)\mathcal{U}^{rs}(X,\mathcal{G}_{\boldsymbol{\theta}}) is free. ∎

5.3. Poisson Action of GDG_{D} and the Moment Map

We now prove that the natural action of GDG_{D} on the moduli space 𝒰(X,𝒢𝜽)\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}}) extends to a Poisson action on its cotangent bundle T𝒰(X,𝒢𝜽)T^{*}\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}}). In particular, we identify a canonical moment map, showing explicitly that it arises from a (co)residue pairing when viewed through Serre duality.

We can perform a Hamiltonian lifting of the GDG_{D}-action from 𝒰(X,𝒢𝜽)\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}}) to T𝒰(X,𝒢𝜽)T^{*}\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}}) as in Example 5.8. Hence we will get a Poisson GDG_{D}-action on the cotangent immediately once we exhibit a moment map

μ:T𝒰(X,𝒢𝜽)𝔤D.\mu:T^{*}\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}})\;\longrightarrow\;\mathfrak{g}_{D}^{*}.

In this direction, let \mathcal{E} be a parahoric torsor and consider the short exact sequence

0(𝔤)η(𝔤)πxjD(𝔤θxj/𝔤θxj+)0.0\to\mathcal{E}(\mathfrak{g})_{\eta}\longrightarrow\mathcal{E}(\mathfrak{g})\stackrel{{\scriptstyle\pi}}{{\longrightarrow}}\bigoplus_{x_{j}\in D}\mathcal{E}(\mathfrak{g}_{\theta_{x_{j}}}/\mathfrak{g}_{\theta_{x_{j}}}^{+})\to 0.

Taking cohomology yields

0H0(X,(𝔤)η)H0(X,(𝔤))xjDH0(𝔻xj,(𝔤θxj/𝔤θxj+))0\to H^{0}\bigl{(}X,\mathcal{E}(\mathfrak{g})_{\eta}\bigr{)}\longrightarrow H^{0}\bigl{(}X,\mathcal{E}(\mathfrak{g})\bigr{)}\longrightarrow\bigoplus_{x_{j}\in D}H^{0}\Bigl{(}\mathbb{D}_{x_{j}},\mathcal{E}(\mathfrak{g}_{\theta_{x_{j}}}/\mathfrak{g}_{\theta_{x_{j}}}^{+})\Bigr{)}
H1(X,(𝔤)η)H1(π)H1(X,(𝔤))0.\longrightarrow H^{1}\bigl{(}X,\mathcal{E}(\mathfrak{g})_{\eta}\bigr{)}\stackrel{{\scriptstyle H^{1}(\pi)}}{{\longrightarrow}}H^{1}\bigl{(}X,\mathcal{E}(\mathfrak{g})\bigr{)}\to 0.

By Serre duality, we have the commutative diagram:

H1(X,(𝔤)η){H^{1}\bigl{(}X,\mathcal{E}(\mathfrak{g})_{\eta}\bigr{)}}H1(X,(𝔤)){H^{1}\bigl{(}X,\mathcal{E}(\mathfrak{g})\bigr{)}}H0(X,(𝔤)KX(D)){H^{0}\Bigl{(}X,\mathcal{E}(\mathfrak{g})^{*}\otimes K_{X}(D)\Bigr{)}^{*}}H0(X,(𝔤)KX).{H^{0}\Bigl{(}X,\mathcal{E}(\mathfrak{g})^{*}\otimes K_{X}\Bigr{)}^{*}.}H1(π)\scriptstyle{H^{1}(\pi)}\scriptstyle{\cong}\scriptstyle{\cong}(H0(πidKX))\scriptstyle{(H^{0}(\pi\otimes\mathrm{id}_{K_{X}}))^{*}}

Thus, one obtains canonical isomorphisms

[xjDH0(𝔻xj,(𝔤θxj/𝔤θxj+))H0(X,(𝔤))]Ker[H0(πidKX)].\Biggl{[}\frac{\bigoplus_{x_{j}\in D}H^{0}\bigl{(}\mathbb{D}_{x_{j}},\mathcal{E}(\mathfrak{g}_{\theta_{x_{j}}}/\mathfrak{g}_{\theta_{x_{j}}}^{+})\bigr{)}}{H^{0}\bigl{(}X,\mathcal{E}(\mathfrak{g})\bigr{)}}\Biggr{]}\cong\mathrm{Ker}\Bigl{[}H^{0}\bigl{(}\pi\otimes\operatorname{id}_{K_{X}}\bigr{)}^{*}\Bigr{]}.

Now, identifying the cokernel of H0(πidKX)H^{0}(\pi\otimes\operatorname{id}_{K_{X}}) with the dual of the kernel above, we define a homomorphism

μ:H0(X,(𝔤)KX(D))[xjDH0(𝔻xj,(𝔤θxj/𝔤θxj+))H0(X,(𝔤))].\mu_{\mathcal{E}}:H^{0}\Bigl{(}X,\mathcal{E}(\mathfrak{g})^{*}\otimes K_{X}(D)\Bigr{)}\longrightarrow\Biggl{[}\frac{\bigoplus_{x_{j}\in D}H^{0}\Bigl{(}\mathbb{D}_{x_{j}},\mathcal{E}(\mathfrak{g}_{\theta_{x_{j}}}/\mathfrak{g}_{\theta_{x_{j}}}^{+})\Bigr{)}}{H^{0}\bigl{(}X,\mathcal{E}(\mathfrak{g})\bigr{)}}\Biggr{]}^{*}.

More precisely, after composing with the natural projection and injection, one obtains

μ:H0(X,(𝔤)KX(D))[xDH0(x,(𝔤θx/𝔤θx+))].\mu_{\mathcal{E}}:\;H^{0}\Bigl{(}X,\mathcal{E}(\mathfrak{g})^{*}\otimes K_{X}(D)\Bigr{)}\longrightarrow\Biggl{[}\frac{\bigoplus_{x\in D}H^{0}\bigl{(}x,\mathcal{E}(\mathfrak{g}_{\theta_{x}}/\mathfrak{g}_{\theta_{x}}^{+})\bigr{)}}{\mathbb{C}}\Biggr{]}^{*}.

We thus introduce the following:

Definition 5.15.

Define the moment map

μ:T𝒰(X,𝒢𝜽)𝔤D\mu:T^{*}\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}})\longrightarrow\mathfrak{g}_{D}^{*}

by sending a point represented by (,φ,η)(\mathcal{E},\varphi,\eta) to

μ([(,η)],φ)=η(μ(φ))η1.\mu(\bigl{[}(\mathcal{E},\eta)\bigr{]},\varphi)=\eta\circ\bigl{(}\mu_{\mathcal{E}}(\varphi)\bigr{)}\circ\eta^{-1}.

Since η\eta is a DD-level structure, it identifies each fiber of \mathcal{E} at xjDx_{j}\in D with GθxjG_{\theta_{x_{j}}}. Hence we compose with the appropriate adjoint-conjugation η()η1\eta\circ(-)\circ\eta^{-1} so that everything is intrinsically defined in 𝔤D\mathfrak{g}_{D}^{*}, independent of the choice of local trivialization. Symbolically, we get ημ(φ)η1\eta\circ\mu_{\mathcal{E}}(\varphi)\circ\eta^{-1}. Notice also that although the level structure η\eta is defined only up to the natural Aut()\operatorname{Aut}(\mathcal{E})-action, the composition ημ(φ)η1\eta\circ\mu_{\mathcal{E}}(\varphi)\circ\eta^{-1} is well-defined.

The homomorphism μ\mu_{\mathcal{E}} above is induced by the natural pairing

H0(X,(𝔤)KX(D))xDH0(𝔻x,(𝔤θx/𝔤θx+))H0(X,KX(D))Res.H^{0}\Bigl{(}X,\mathcal{E}(\mathfrak{g})^{*}\otimes K_{X}(D)\Bigr{)}\otimes\bigoplus_{x\in D}H^{0}\Bigl{(}\mathbb{D}_{x},\mathcal{E}(\mathfrak{g}_{\theta_{x}}/\mathfrak{g}_{\theta_{x}}^{+})\Bigr{)}\longrightarrow H^{0}\bigl{(}X,K_{X}(D)\bigr{)}\stackrel{{\scriptstyle\operatorname{Res}}}{{\longrightarrow}}\mathbb{C}. (5.1)

This pairing provides the link to the local structure. At a point xjDx_{j}\in D, let ϕj\phi_{j} be a local representative of the Higgs field φ\varphi. The Lie algebra 𝔤θj\mathfrak{g}_{\theta_{j}} has the Levi decomposition 𝔤θj=𝔩^θj𝔤θj+\mathfrak{g}_{\theta_{j}}=\hat{\mathfrak{l}}_{\theta_{j}}\oplus\mathfrak{g}_{\theta_{j}}^{+}, which is an orthogonal direct sum with respect to the Killing form, which we denote by (,)(\cdot,\cdot). The term (𝔤θj)/(𝔤θj+)\mathcal{E}(\mathfrak{g}_{\theta_{j}})/\mathcal{E}(\mathfrak{g}_{\theta_{j}}^{+}) is the bundle associated to the Levi quotient 𝔩θj𝔩^θj\mathfrak{l}_{\theta_{j}}\cong\hat{\mathfrak{l}}_{\theta_{j}}. When pairing ϕj\phi_{j} with an element Yj𝔩^θjY_{j}\in\hat{\mathfrak{l}}_{\theta_{j}}, the component of ϕj\phi_{j} in 𝔤θj+\mathfrak{g}_{\theta_{j}}^{+} gives a zero contribution. Thus, the pairing only sees the projection of ϕj\phi_{j} onto 𝔩^θj\hat{\mathfrak{l}}_{\theta_{j}}. We call this projection the residue, Resxj(φ)\operatorname{Res}_{x_{j}}(\varphi). The pairing (ϕj,Yj)(\phi_{j},Y_{j}) becomes (Resxj(φ),Yj)(\operatorname{Res}_{x_{j}}(\varphi),Y_{j}).

The resulting functional on 𝔩^θj\hat{\mathfrak{l}}_{\theta_{j}} is the coresidue of φ\varphi at xjx_{j}, an element of 𝔩^θj\hat{\mathfrak{l}}_{\theta_{j}}^{*}. The moment map μ\mu is therefore the collection of these coresidues over all xjDx_{j}\in D.

Theorem 5.16.

The group GDG_{D} acts Poisson on T𝒰(X,𝒢𝛉)T^{*}\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}}). Moreover, the canonical moment map

μ:T𝒰(X,𝒢𝜽)𝔤D\mu\colon T^{*}\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}})\;\longrightarrow\;\mathfrak{g}_{D}^{*}

is given by dualizing the infinitesimal action and can be explicitly described via coresidues at the divisor DD. Its image is the element of 𝔤D=j=1s𝔩^θj\mathfrak{g}_{D}^{*}=\bigoplus_{j=1}^{s}\hat{\mathfrak{l}}_{\theta_{j}}^{*} given by the direct sum of the coresidues at each point xjDx_{j}\in D:

μ([(,η)],φ)=j=1sCoResxj(φ).\mu([(\mathcal{E},\eta)],\varphi)=\bigoplus_{j=1}^{s}\operatorname{CoRes}_{x_{j}}(\varphi).

Explicitly, for any element Y=(Y1,,Ys)𝔤D=j=1s𝔩^θjY=(Y_{1},\dots,Y_{s})\in\mathfrak{g}_{D}=\bigoplus_{j=1}^{s}\hat{\mathfrak{l}}_{\theta_{j}}, the pairing is given by the sum of Killing form pairings over the divisor DD:

μ([(,η)],φ),Y=j=1s(Resxj(φ),Yj).\langle\mu([(\mathcal{E},\eta)],\varphi),Y\rangle=\sum_{j=1}^{s}(\operatorname{Res}_{x_{j}}(\varphi),Y_{j}).
Proof.

Let us denote here for convenience U:=𝒰(X,𝒢𝜽)U:=\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}}). For a stable parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor \mathcal{E} with DD-level structure η\eta, we have an isomorphism T[(,η)]UH1(X,(𝔤)η)T_{[(\mathcal{E},\eta)]}U\cong H^{1}(X,\mathcal{E}(\mathfrak{g})_{\eta}), where (𝔤)η(𝔤)\mathcal{E}(\mathfrak{g})_{\eta}\subset\mathcal{E}(\mathfrak{g}) denotes the subsheaf consisting of local adjoint-valued sections that preserve the DD-level structure.

Dualizing and invoking Serre duality (together with the twist by DD at the points supporting the parahoric structure), one obtains T[(,η)]UH1(X,(𝔤)η)H0(X,(𝔤)K(D))T^{*}_{[(\mathcal{E},\eta)]}U\cong H^{1}(X,\mathcal{E}(\mathfrak{g})_{\eta})^{*}\simeq H^{0}(X,\mathcal{E}(\mathfrak{g})\otimes K(D)). Hence a point of TUT^{*}U can be represented by a logarithmic Higgs field φ\varphi.

To identify μ\mu explicitly, we must describe how 𝔤D\mathfrak{g}_{D} acts infinitesimally on UU. Let

σ:𝔤DT[(,η)]UH1(X,(𝔤)η)\sigma\colon\mathfrak{g}_{D}\;\longrightarrow\;T_{[(\mathcal{E},\eta)]}U\;\cong\;H^{1}(X,\mathcal{E}(\mathfrak{g})_{\eta})

be the differential of the map ρ[(,η)]:GDU\rho_{[(\mathcal{E},\eta)]}\colon G_{D}\to U sending gg[(,η)]g\mapsto g\cdot[(\mathcal{E},\eta)]. Concretely, each element of 𝔤D\mathfrak{g}_{D} corresponds to an infinitesimal transformation of the DD-level structure at each parahoric point, and these local transformations glue to give a global 1-cocycle in (𝔤)η\mathcal{E}(\mathfrak{g})_{\eta}.

By general principles of Hamiltonian actions on cotangent bundles, the moment map μ\mu of Definition 5.15 is precisely the dual of σ\sigma. For a cotangent vector ([(,η)],φ)([(\mathcal{E},\eta)],\varphi) and an element X𝔤DX\in\mathfrak{g}_{D}, we have μ([(,η)],φ)(X)=φ,σ(X)\mu([(\mathcal{E},\eta)],\varphi)(X)=\langle\varphi,\sigma(X)\rangle. As established in the discussion following (5.1), this pairing precisely computes the coresidue.

To show equivariance, we must prove that for gGDg\in G_{D} and φ=Ad(g)φ\varphi^{\prime}=\operatorname{Ad}(g)\varphi, the moment map transforms via the coadjoint action, i.e., μ(φ)=Ad(g)μ(φ)\mu(\varphi^{\prime})=\operatorname{Ad}^{*}(g)\mu(\varphi). It is sufficient to verify this at each point xjDx_{j}\in D. Let us denote the duality pairing between 𝔩^θj\hat{\mathfrak{l}}_{\theta_{j}}^{*} and 𝔩^θj\hat{\mathfrak{l}}_{\theta_{j}} by ,\langle\cdot,\cdot\rangle, and the Killing form on 𝔩^θj\hat{\mathfrak{l}}_{\theta_{j}} by (,)(\cdot,\cdot). We need to show that CoResxj(φ)=Ad(gj)CoResxj(φ)\operatorname{CoRes}_{x_{j}}(\varphi^{\prime})=\operatorname{Ad}^{*}(g_{j})\operatorname{CoRes}_{x_{j}}(\varphi), but this is immediate by the definition of coresidue and its action.

Since we have shown that CoResxj(φ),Yj=Ad(gj)CoResxj(φ),Yj\langle\operatorname{CoRes}_{x_{j}}(\varphi^{\prime}),Y_{j}\rangle=\langle\operatorname{Ad}^{*}(g_{j})\operatorname{CoRes}_{x_{j}}(\varphi),Y_{j}\rangle, for all Yj𝔩^θjY_{j}\in\hat{\mathfrak{l}}_{\theta_{j}}, we conclude that CoResxj(φ)=Ad(gj)CoResxj(φ)\operatorname{CoRes}_{x_{j}}(\varphi^{\prime})=\operatorname{Ad}^{*}(g_{j})\operatorname{CoRes}_{x_{j}}(\varphi). The total moment map transforms accordingly, proving the theorem. ∎

5.4. The Infinitesimal Deformation Complex

Let XX be a smooth complex projective curve. In [24] we have constructed the moduli space H(X,𝒢𝜽)\mathcal{M}_{H}(X,\mathcal{G}_{\boldsymbol{\theta}}) of RR-semistable logahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-Higgs torsors (,φ)(\mathcal{E},\varphi) over XX. We now consider isomorphism classes of triples (,φ,η)(\mathcal{E},\varphi,\eta), where

  • \mathcal{E} is a stable parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor over XX,

  • φH0(X,(𝔤)K(D))\varphi\in H^{0}\bigl{(}X,\mathcal{E}(\mathfrak{g})\otimes K(D)\bigr{)} is a logarithmic Higgs field, and

  • η\eta is a DD-level structure on \mathcal{E}.

One can construct a coarse moduli space MLH(X,𝒢𝜽)M_{LH}(X,\mathcal{G}_{\boldsymbol{\theta}}) of triples (,φ,η)(\mathcal{E},\varphi,\eta) as above using a similar approach as the one in Section 3.3 for constructing the moduli space 𝒰(X,𝒢𝜽)\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}}). In particular, this involves the correspondence between pairs (,φ)(\mathcal{E},\varphi) over XX and equivariant logarithmic GG-Higgs bundles on a Galois cover of XX; see [24, Section 3] for this correspondence. Then, equipping the pairs (,φ)(\mathcal{E},\varphi) with a DD-level structure not affecting the stability condition of the logahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-Higgs torsor allows us to construct a good quotient as in Section 3.3. We call the moduli space MLH(X,𝒢𝜽)M_{LH}(X,\mathcal{G}_{\boldsymbol{\theta}}), the moduli space of leveled logahoric 𝒢𝛉\mathcal{G}_{\boldsymbol{\theta}}-Higgs torsors over XX. Note that then the cotangent T𝒰X(G,D)T^{*}\mathcal{U}_{X}(G,D) of the moduli space 𝒰X(G,D)\mathcal{U}_{X}(G,D) is an open subset of this moduli space MLH(X,𝒢𝜽)M_{LH}(X,\mathcal{G}_{\boldsymbol{\theta}}).

We would like to consider the deformation theory for this moduli space. Given a logarithmic Higgs field

φH0(X,(𝔤)K(D)),\varphi\in H^{0}\bigl{(}X,\mathcal{E}(\mathfrak{g})\otimes K(D)\bigr{)},

we define the adjoint action

adφ:(𝔤)(𝔤)K(D), with adφ(ψ)=φψψφ.\operatorname{ad}\varphi:\mathcal{E}(\mathfrak{g})\to\mathcal{E}(\mathfrak{g})\otimes K(D),\quad\text{ with }\operatorname{ad}\varphi(\psi)=\varphi\circ\psi-\psi\circ\varphi.

Then, a natural two-term complex governing the deformations of the triple (,φ,η)(\mathcal{E},\varphi,\eta) is

𝒦,φ,η:(𝔤)φ,η((𝔤)K(D))xjD[(𝔤θxj/𝔤θxj+)],\mathcal{K}_{\mathcal{E},\varphi,\eta}\colon\quad\mathcal{E}(\mathfrak{g})\xrightarrow{\partial_{\varphi,\eta}}\bigl{(}\mathcal{E}(\mathfrak{g})\otimes K(D)\bigr{)}\oplus\bigoplus_{x_{j}\in D}\left[\mathcal{E}(\mathfrak{g}_{\theta_{x_{j}}}/\mathfrak{g}_{\theta_{x_{j}}}^{+})\right],

where the differential is given by

φ,η(ψ)=adφ(ψ)mη(ψ).\partial_{\varphi,\eta}(\psi)=\operatorname{ad}\varphi(\psi)\oplus m_{\eta}(\psi).

Here, the term mη(ψ)m_{\eta}(\psi) arises from the level structure. In practice, the level structure η\eta defines a reduction of the fibers of (𝔤)\mathcal{E}(\mathfrak{g}) over DD (via, say, a choice of splitting or a canonical quotient) and induces a homomorphism

mη:(𝔤)xjD[(𝔤θxj/𝔤θxj+)].m_{\eta}\colon\mathcal{E}(\mathfrak{g})\to\bigoplus_{x_{j}\in D}\left[\mathcal{E}(\mathfrak{g}_{\theta_{x_{j}}}/\mathfrak{g}_{\theta_{x_{j}}}^{+})\right]. (5.2)

We will prove in Lemma 5.19 later on that this complex is quasi-isomorphic to the complex

𝒦,φ:(𝔤+)(𝔤)K(D),\mathcal{K}_{\mathcal{E},\varphi}\colon\quad\mathcal{E}(\mathfrak{g}^{+})\xrightarrow{\partial}\mathcal{E}(\mathfrak{g})\otimes K(D),

by the mapping (φ)=ad(φ)\partial(\varphi)=\operatorname{ad}(\varphi), for the Lie algebra bundle (𝔤+)\mathcal{E}(\mathfrak{g}^{+}) defined by the gluing of the Lie algebra bundles (𝔤θxj+)\mathcal{E}(\mathfrak{g}^{+}_{\theta_{x_{j}}}) from Section 2.3.

To prove that this complex is indeed the deformation, we perform a local calculation. Choose a Čech covering {Uα}\{U_{\alpha}\} of XX and let 𝒲:={Wα}\mathcal{W}^{\prime}:=\{W_{\alpha}\} be the covering of S×XS\times X, for Wα=S×UαW_{\alpha}=S\times U_{\alpha}, where S=Spec([ϵ]/(ϵ2))S=\mathrm{Spec}(\mathbb{C}[\epsilon]/(\epsilon^{2})) is an infinitesimal family. We have the following:

Lemma 5.17.

Fix a logahoric 𝒢𝛉\mathcal{G}_{\boldsymbol{\theta}}-Higgs torsor (0,φ0)(\mathcal{E}_{0},\varphi_{0}) with a DD-level structure η0\eta_{0}. A cochain

(f^,φ^,η^):=((f^αβ),φ^α,η^α)(\hat{f},\hat{\varphi},\hat{\eta}):=((\hat{f}_{\alpha\beta}),\,\hat{\varphi}_{\alpha},\,\hat{\eta}_{\alpha})

in C0(𝒲,𝒦,φ,η)C^{0}(\mathcal{W}^{\prime},\mathcal{K}_{\mathcal{E},\varphi,\eta}) is a cocycle if and only if:

  1. (1)

    The cochain f^=(f^αβ)\hat{f}=(\hat{f}_{\alpha\beta}) is a cocycle in Z1(𝒰,0(𝔤))Z^{1}(\mathcal{U},\mathcal{E}_{0}(\mathfrak{g})), that is, it defines an infinitesimal deformation of the parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor 0\mathcal{E}_{0}.

  2. (2)

    The perturbed Higgs field

    φ0+ϵφ^C0(𝒲,p(0(𝔤)K(D)))\varphi_{0}+\epsilon\cdot\hat{\varphi}\in C^{0}\Bigl{(}\mathcal{W}^{\prime},\,p^{*}\bigl{(}\mathcal{E}_{0}(\mathfrak{g})\otimes K(D)\bigr{)}\Bigr{)}

    is a global section of (𝔤)pK(D)\mathcal{E}(\mathfrak{g})\otimes p^{*}{K(D)}, where \mathcal{E} is the infinitesimal family of parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsors over S×XS\times X defined by the new gluing transformations

    fαβ=id+ϵf^αβf_{\alpha\beta}=\operatorname{id}+\epsilon\cdot\hat{f}_{\alpha\beta}

    for p0p^{*}\mathcal{E}_{0}, where pp is the projection on XX.

  3. (3)

    The perturbed level structure

    η0+ϵη^C0(𝒲,p(xjD[(𝔤θxj/𝔤θxj+)]))\eta_{0}+\epsilon\cdot\hat{\eta}\in C^{0}\Bigl{(}\mathcal{W}^{\prime},\,p^{*}\Bigl{(}\bigoplus_{x_{j}\in D}\Bigl{[}\mathcal{E}(\mathfrak{g}_{\theta_{x_{j}}}/\mathfrak{g}_{\theta_{x_{j}}}^{+})\Bigr{]}\Bigr{)}\Bigr{)}

    is a global section η\eta of the sheaf defining the DD^{\prime}-level structure (with the fibers VxjV_{x_{j}} obtained from the natural quotients associated to the parahoric groups at xjDx_{j}\in D^{\prime}).

Moreover, two cocycles

((f^αβ),φ^α,η^α)and((f^αβ),φ^α,η^α)((\hat{f}_{\alpha\beta}),\,\hat{\varphi}_{\alpha},\,\hat{\eta}_{\alpha})\quad\text{and}\quad((\hat{f}^{\prime}_{\alpha\beta}),\,\hat{\varphi}^{\prime}_{\alpha},\,\hat{\eta}^{\prime}_{\alpha})

represent the same hypercohomology class if and only if the corresponding infinitesimal families

(,φ,η)and(,φ,η)(\mathcal{E},\varphi,\eta)\quad\text{and}\quad(\mathcal{E}^{\prime},\varphi^{\prime},\eta^{\prime})

are isomorphic.

Proof.

A cochain (f^,φ^,η^)(\hat{f},\hat{\varphi},\hat{\eta}) in the complex 𝒦,φ,η\mathcal{K}_{\mathcal{E},\varphi,\eta} is a cocycle if and only if the following conditions hold:

  1. (1)

    The Čech differential δ(f^)\delta(\hat{f}) vanishes, i.e.,

    δ(f^)=0,\delta(\hat{f})=0,

    so that f^\hat{f} defines an infinitesimal deformation of the parahoric torsor (more precisely, of its adjoint bundle) via the usual cocycle condition in Z1(𝒰,0(𝔤))Z^{1}(\mathcal{U},\mathcal{E}_{0}(\mathfrak{g})).

  2. (2)

    The compatibility between the deformed Higgs field and the new gluing is expressed by

    adφ0(f^)=δ(φ^),-\operatorname{ad}_{\varphi_{0}}(\hat{f})=\delta(\hat{\varphi}),

    which ensures that the perturbed Higgs field φ0+ϵφ^\varphi_{0}+\epsilon\hat{\varphi} is well–defined on the deformed torsor. To see this, note that on triple overlaps one must have

    fβγφ0fαβ1=φα,f_{\beta\gamma}\varphi_{0}f_{\alpha\beta}^{-1}=\varphi_{\alpha},

    and writing

    fαβ=id+ϵf^αβ,φα=φ0+ϵφ^α,f_{\alpha\beta}=\operatorname{id}+\epsilon\hat{f}_{\alpha\beta},\quad\varphi_{\alpha}=\varphi_{0}+\epsilon\hat{\varphi}_{\alpha},

    one checks (to first order in ϵ\epsilon) that the condition is equivalent to

    adφ0(f^αβ)=φ^αφ^β,-\operatorname{ad}_{\varphi_{0}}(\hat{f}_{\alpha\beta})=\hat{\varphi}_{\alpha}-\hat{\varphi}_{\beta},

    that is, δ(φ^)=adφ0(f^)\delta(\hat{\varphi})=-\operatorname{ad}_{\varphi_{0}}(\hat{f}).

  3. (3)

    Similarly, the compatibility of the deformed level structure is encoded by the equation

    η0f^=δ(η^).-\eta_{0}\circ\hat{f}=\delta(\hat{\eta}).

    Indeed, on overlaps the new level structure must satisfy

    ηz=(η0+ϵη^)(id+ϵf^)=η0+ϵ(η^+η0f^),\eta_{z}=(\eta_{0}+\epsilon\hat{\eta})\circ(\operatorname{id}+\epsilon\hat{f})=\eta_{0}+\epsilon\Bigl{(}\hat{\eta}+\eta_{0}\circ\hat{f}\Bigr{)},

    so that the cocycle condition for η^\hat{\eta} becomes δ(η^)=η0f^\delta(\hat{\eta})=-\eta_{0}\circ\hat{f}.

Finally, one verifies that a cocycle (f^,φ^,η^)(\hat{f},\hat{\varphi},\hat{\eta}) represents a trivial deformation (i.e. a coboundary) if and only if there exists a 0-cochain g^=(g^α)\hat{g}=(\hat{g}_{\alpha}) in C0(𝒰,0(𝔤))C^{0}(\mathcal{U},\mathcal{E}_{0}(\mathfrak{g})) such that:

  1. a)

    f^αβ=g^αg^β\hat{f}_{\alpha\beta}=\hat{g}_{\alpha}-\hat{g}_{\beta},

  2. b)

    φ^α=adφ0(g^α)\hat{\varphi}_{\alpha}=-\operatorname{ad}_{\varphi_{0}}(\hat{g}_{\alpha}),

  3. c)

    η^α=η0g^α\hat{\eta}_{\alpha}=-\eta_{0}\circ\hat{g}_{\alpha}.

This precisely corresponds to the statement that the infinitesimal families (,φ,η)(\mathcal{E},\varphi,\eta) and (,φ,η)(\mathcal{E}^{\prime},\varphi^{\prime},\eta^{\prime}) are isomorphic. ∎

This leads to the following characterization:

Proposition 5.18.

Let (,φ,η)(\mathcal{E},\varphi,\eta) be a point in LH(X,𝒢θ)\mathcal{M}_{LH}(X,\mathcal{G}_{\theta}). Then, (,φ,η)(\mathcal{E},\varphi,\eta) determines a canonical isomorphism

T(,φ,η)LH(X,𝒢θ)1(𝒦,φ,η)T_{(\mathcal{E},\varphi,\eta)}\mathcal{M}_{LH}(X,\mathcal{G}_{\theta})\cong\mathbb{H}^{1}\bigl{(}\mathcal{K}_{\mathcal{E},\varphi,\eta}\bigr{)}

between the Zariski tangent space at (,φ,η)(\mathcal{E},\varphi,\eta) and the hypercohomology of the complex 𝒦,φ,η\mathcal{K}_{\mathcal{E},\varphi,\eta}.

We finally get a relationship between the complexes 𝒦,φ\mathcal{K}_{\mathcal{E},\varphi} and 𝒦,φ,η\mathcal{K}_{\mathcal{E},\varphi,\eta}:

Lemma 5.19.

Let [(,φ,η)]S[(\mathcal{E},\varphi,\eta)]_{S} be a flat family of leveled parahoric 𝒢𝛉\mathcal{G}_{\boldsymbol{\theta}}-Higgs torsors parametrized by a scheme SS. There exists a canonical isomorphism

RPSi(𝒦,φ,η)RPSi(𝒦,φ)R^{i}_{{P_{S^{*}}}}\bigl{(}\mathcal{K}_{\mathcal{E},\varphi,\eta}\bigr{)}\cong R^{i}_{{P_{S^{*}}}}\bigl{(}\mathcal{K}_{\mathcal{E},\varphi}\bigr{)}

for i=0,1i=0,1, induced by a canonical quasi-isomorphism.

Proof.

We define a morphism of complexes

qi:𝒦,φ,η𝒦,φqi\colon\mathcal{K}_{\mathcal{E},\varphi,\eta}\longrightarrow\mathcal{K}_{\mathcal{E},\varphi}

by specifying its components in degrees 0 and 11.

Degree 0: The level structure η\eta forces the reduction (𝔤)\mathcal{E}(\mathfrak{g}) to lie in the subbundle (𝔤+)\mathcal{E}(\mathfrak{g}^{+}). Therefore, we define

qi0:(𝔤)(𝔤+)qi_{0}\colon\mathcal{E}(\mathfrak{g})\rightarrow\mathcal{E}(\mathfrak{g}^{+})

to be the natural inclusion.

Degree 1: In degree 11 we have

𝒦,φ,η1=(𝔤)K(D)(xjD[(𝔤θxj/𝔤θxj+)])\mathcal{K}_{\mathcal{E},\varphi,\eta}^{1}=\mathcal{E}(\mathfrak{g})\otimes K(D)\oplus\Biggl{(}\bigoplus_{x_{j}\in D}\Bigl{[}\mathcal{E}(\mathfrak{g}_{\theta_{x_{j}}}/\mathfrak{g}_{\theta_{x_{j}}}^{+})\Bigr{]}\Biggr{)}

and

𝒦,φ1=(𝔤)K(D).\mathcal{K}_{\mathcal{E},\varphi}^{1}=\mathcal{E}(\mathfrak{g})\otimes K(D).

We define

qi1:(𝔤)K(D)(xjD[(𝔤θxj/𝔤θxj+)])(𝔤)K(D)qi_{1}\colon\mathcal{E}(\mathfrak{g})\otimes K(D)\oplus\Biggl{(}\bigoplus_{x_{j}\in D}\Bigl{[}\mathcal{E}(\mathfrak{g}_{\theta_{x_{j}}}/\mathfrak{g}_{\theta_{x_{j}}}^{+})\Bigr{]}\Biggr{)}\longrightarrow\mathcal{E}(\mathfrak{g})\otimes K(D)

by projecting onto the first summand:

qi1(α,β)=α.qi_{1}(\alpha,\beta)=\alpha.

It is straightforward to check that qi=(qi0,qi1)qi=(qi_{0},qi_{1}) defines a morphism of complexes.

We now show that qiqi is a quasi-isomorphism, i.e., that it induces isomorphisms on the cohomology sheaves.

Degree 0: The degree 0 cohomology of 𝒦,φ\mathcal{K}_{\mathcal{E},\varphi} is

0(𝒦,φ)=Ker(adφ:(𝔤+)(𝔤)K(D)).\mathcal{H}^{0}\bigl{(}\mathcal{K}_{\mathcal{E},\varphi}\bigr{)}=\mathrm{Ker}\Bigl{(}\operatorname{ad}\varphi\colon\mathcal{E}(\mathfrak{g}^{+})\to\mathcal{E}(\mathfrak{g})\otimes K(D)\Bigr{)}.

For the complex 𝒦,φ,η\mathcal{K}_{\mathcal{E},\varphi,\eta} we have

0(𝒦,φ,η)=Ker(adφmη:(𝔤)((𝔤)K(D))xjD[(𝔤θxj/𝔤θxj+)]),\mathcal{H}^{0}\bigl{(}\mathcal{K}_{\mathcal{E},\varphi,\eta}\bigr{)}=\mathrm{Ker}\Bigl{(}\operatorname{ad}\varphi\oplus m_{\eta}\colon\mathcal{E}(\mathfrak{g})\to\bigl{(}\mathcal{E}(\mathfrak{g})\otimes K(D)\bigr{)}\oplus\bigoplus_{x_{j}\in D}\Bigl{[}\mathcal{E}(\mathfrak{g}_{\theta_{x_{j}}}/\mathfrak{g}_{\theta_{x_{j}}}^{+})\Bigr{]}\Bigr{)},

for mηm_{\eta} as defined in (5.2). Since the condition mη(ψ)=0m_{\eta}(\psi)=0 forces ψ\psi to lie in (𝔤+)\mathcal{E}(\mathfrak{g}^{+}), we deduce that

0(𝒦,φ,η)=Ker(adφ:(𝔤+)(𝔤)K(D)).\mathcal{H}^{0}\bigl{(}\mathcal{K}_{\mathcal{E},\varphi,\eta}\bigr{)}=\mathrm{Ker}\Bigl{(}\operatorname{ad}\varphi\colon\mathcal{E}(\mathfrak{g}^{+})\to\mathcal{E}(\mathfrak{g})\otimes K(D)\Bigr{)}.

Thus, qi0qi_{0} induces an isomorphism on degree 0 cohomology.

Degree 1: The degree 11 cohomology of 𝒦,φ\mathcal{K}_{\mathcal{E},\varphi} is

1(𝒦,φ)=coker(adφ:(𝔤+)(𝔤)K(D)).\mathcal{H}^{1}\bigl{(}\mathcal{K}_{\mathcal{E},\varphi}\bigr{)}=\operatorname{coker}\Bigl{(}\operatorname{ad}\varphi\colon\mathcal{E}(\mathfrak{g}^{+})\to\mathcal{E}(\mathfrak{g})\otimes K(D)\Bigr{)}.

Similarly, one computes

1(𝒦,φ,η)=(𝔤)K(D)(xjD[(𝔤θxj/𝔤θxj+)])Im(adφmη).\mathcal{H}^{1}\bigl{(}\mathcal{K}_{\mathcal{E},\varphi,\eta}\bigr{)}=\frac{\mathcal{E}(\mathfrak{g})\otimes K(D)\oplus\Biggl{(}\bigoplus_{x_{j}\in D}\Bigl{[}\mathcal{E}(\mathfrak{g}_{\theta_{x_{j}}}/\mathfrak{g}_{\theta_{x_{j}}}^{+})\Bigr{]}\Biggr{)}}{\operatorname{Im}\Bigl{(}\operatorname{ad}\varphi\oplus m_{\eta}\Bigr{)}}.

A local analysis (analogous to the vector bundle case) shows that the extra summand coming from the level structure exactly compensates for passing from (𝔤)\mathcal{E}(\mathfrak{g}) to (𝔤+)\mathcal{E}(\mathfrak{g}^{+}). Consequently, we obtain an isomorphism

1(𝒦,φ,η)coker(adφ:(𝔤+)(𝔤)K(D))=1(𝒦,φ).\mathcal{H}^{1}\bigl{(}\mathcal{K}_{\mathcal{E},\varphi,\eta}\bigr{)}\cong\operatorname{coker}\Bigl{(}\operatorname{ad}\varphi\colon\mathcal{E}(\mathfrak{g}^{+})\to\mathcal{E}(\mathfrak{g})\otimes K(D)\Bigr{)}=\mathcal{H}^{1}\bigl{(}\mathcal{K}_{\mathcal{E},\varphi}\bigr{)}.

Since qi1qi_{1} is the projection onto the first factor (acting as the identity on (𝔤)K(D)\mathcal{E}(\mathfrak{g})\otimes K(D)), it follows that qi1qi_{1} induces an isomorphism on degree 11 cohomology.

Thus, the morphism qiqi is a quasi-isomorphism. In particular, for each i=0,1i=0,1, we have the canonical isomorphism

RPSi(𝒦,φ,η)RPSi(𝒦,φ).R^{i}_{{P_{S^{*}}}}\bigl{(}\mathcal{K}_{\mathcal{E},\varphi,\eta}\bigr{)}\cong R^{i}_{{P_{S^{*}}}}\bigl{(}\mathcal{K}_{\mathcal{E},\varphi}\bigr{)}.

This completes the proof. ∎

5.5. Extension of Poisson Structure

Finally, we would like to consider the Poisson structure on the moduli space of logahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-Higgs torsors.

Let (E,φ)(E,\varphi) be a stable logahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-Higgs torsor in H(X,𝒢θ)\mathcal{M}_{H}(X,\mathcal{G}_{\theta}). Its tangent space is naturally identified with the first hypercohomology group of the two‐term complex

𝒦¯,φ:(𝔤)[,φ](𝔤)K(D),\bar{\mathcal{K}}_{\mathcal{E},\varphi}:\mathcal{E}(\mathfrak{g})\xrightarrow{\,[\,\cdot\,,\varphi]\,}\mathcal{E}(\mathfrak{g})\otimes K(D),

where (𝔤)\mathcal{E}(\mathfrak{g}) denotes the sheaf of endomorphisms of \mathcal{E} that preserve the parahoric structure, and K(D)K(D) is the canonical bundle twisted by the divisor DD. To describe a Poisson bracket we next consider the dual complex obtained by tensoring with K(D)K(D) and reversing the sign of the differential. This yields

𝒦¯,φ:(𝔤+)[,φ](𝔤+)K(D).\bar{\mathcal{K}}_{\mathcal{E},\varphi}^{\vee}:\mathcal{E}(\mathfrak{g}^{+})\xrightarrow{\,-[\cdot,\varphi]\,}\mathcal{E}(\mathfrak{g}^{+})\otimes K(D).

There is then a natural injection from this dual complex into the original one; that is, we have a commutative diagram

(𝔤+)\textstyle{\mathcal{E}(\mathfrak{g}^{+})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}[,Φ]\scriptstyle{-[\cdot,\Phi]}id\scriptstyle{\mathrm{id}}(𝔤)\textstyle{\mathcal{E}(\mathfrak{g})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}[,Φ]\scriptstyle{[\cdot,\Phi]}(𝔤+)K(D)\textstyle{\mathcal{E}(\mathfrak{g}^{+})\otimes K(D)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ididK(D)\scriptstyle{-\,\mathrm{id}\otimes\mathrm{id}_{K(D)}}(𝔤)K(D).\textstyle{\mathcal{E}(\mathfrak{g})\otimes K(D).}

Using Serre duality for hypercohomology, this inclusion produces an antisymmetric linear map

:T(,φ)H(X,𝒢θ)1(𝒦¯,φ)1(𝒦¯,φ)T(,φ)H(X,𝒢θ).\sharp\;:\;T^{*}_{(\mathcal{E},\varphi)}\mathcal{M}_{H}(X,\mathcal{G}_{\theta})\cong\mathbb{H}^{1}\Bigl{(}\bar{\mathcal{K}}_{\mathcal{E},\varphi}^{\vee}\Bigr{)}\longrightarrow\mathbb{H}^{1}\Bigl{(}\bar{\mathcal{K}}_{\mathcal{E},\varphi}\Bigr{)}\cong T_{(\mathcal{E},\varphi)}\mathcal{M}_{H}(X,\mathcal{G}_{\theta}). (5.3)

The specific choices of signs guarantee that \sharp is antisymmetric, and we will show that it indeed defines a Poisson structure on the moduli space H(X,𝒢θ)\mathcal{M}_{H}(X,\mathcal{G}_{\theta}) inheriting the Poisson structure on LH(X,𝒢θ)\mathcal{M}_{LH}(X,\mathcal{G}_{\theta}).

Recall that we have the following:

T(,φ,η)LH(X,𝒢θ)\displaystyle T_{(\mathcal{E},\varphi,\eta)}\mathcal{M}_{LH}(X,\mathcal{G}_{\theta}) 𝒦,φ:(𝔤+)[,φ](𝔤)K(D),\displaystyle\cong\mathcal{K}_{\mathcal{E},\varphi}\colon\quad\mathcal{E}(\mathfrak{g}^{+})\xrightarrow{\,[\,\cdot\,,\varphi]\,}\mathcal{E}(\mathfrak{g})\otimes K(D),
T(,φ,η)LH(X,𝒢θ)\displaystyle T_{(\mathcal{E},\varphi,\eta)}^{*}\mathcal{M}_{LH}(X,\mathcal{G}_{\theta}) 𝒦,φ:(𝔤+)[,φ](𝔤)K(D),\displaystyle\cong\mathcal{K}^{\vee}_{\mathcal{E},\varphi}\colon\quad\mathcal{E}(\mathfrak{g}^{+})\xrightarrow{\,-[\,\cdot\,,\varphi]\,}\mathcal{E}(\mathfrak{g})\otimes K(D),
T(,φ)H(X,𝒢θ)\displaystyle T_{(\mathcal{E},\varphi)}\mathcal{M}_{H}(X,\mathcal{G}_{\theta}) 𝒦¯,φ:(𝔤)[,φ](𝔤)K(D),\displaystyle\cong\bar{\mathcal{K}}_{\mathcal{E},\varphi}\colon\quad\mathcal{E}(\mathfrak{g})\xrightarrow{\,[\,\cdot\,,\varphi]\,}\mathcal{E}(\mathfrak{g})\otimes K(D),
T(,φ)H(X,𝒢θ)\displaystyle T_{(\mathcal{E},\varphi)}^{*}\mathcal{M}_{H}(X,\mathcal{G}_{\theta}) 𝒦¯,φ:(𝔤+)[,φ](𝔤+)K(D).\displaystyle\cong\bar{\mathcal{K}}^{\vee}_{\mathcal{E},\varphi}\colon\quad\mathcal{E}(\mathfrak{g^{+}})\xrightarrow{\,[\,\cdot\,,\varphi]\,}\mathcal{E}(\mathfrak{g^{+}})\otimes K(D).

Then, the next theorem provides the Poisson structure on H(X,𝒢𝜽)\mathcal{M}_{H}(X,\mathcal{G}_{\boldsymbol{\theta}}):

Theorem 5.20.

The forgetful map

:LH(X,𝒢θ)H(X,𝒢θ)\ell:\mathcal{M}_{LH}(X,\mathcal{G}_{\theta})\longrightarrow\mathcal{M}_{H}(X,\mathcal{G}_{\theta})

defined by forgetting the DD-level structure is a Poisson map. The induced Poisson structure on H(X,𝒢𝛉)\mathcal{M}_{H}(X,\mathcal{G}_{\boldsymbol{\theta}}) is given by the map \sharp defined in (5.3).

Proof.

Let

𝐳:=(,φ,η)LH(X,𝒢θ)\mathbf{z}:=(\mathcal{E},\varphi,\eta)\in\mathcal{M}_{LH}(X,\mathcal{G}_{\theta})

be a leveled logahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-Higgs torsor over XX and set

𝐲:=(𝐳)=(,φ)H(X,𝒢θ).\mathbf{y}:=\ell(\mathbf{z})=(\mathcal{E},\varphi)\in\mathcal{M}_{H}(X,\mathcal{G}_{\theta}).

The tangent spaces are given by

T𝐳LH(X,𝒢θ)1(𝒦¯E,ϕ,φ)T_{\mathbf{z}}\mathcal{M}_{LH}(X,\mathcal{G}_{\theta})\cong\mathbb{H}^{1}\Bigl{(}\bar{\mathcal{K}}_{E,\phi,\varphi}\Bigr{)}

and

T𝐲H(X,𝒢θ)1(𝒦¯,φ).T_{\mathbf{y}}\mathcal{M}_{H}(X,\mathcal{G}_{\theta})\cong\mathbb{H}^{1}\Bigl{(}\bar{\mathcal{K}}_{\mathcal{E},\varphi}\Bigr{)}.

Since the DD-level structure is forgotten by \ell, the differential

d(𝐳):T𝐳LH(X,𝒢θ)T𝐲H(X,𝒢θ)d\ell(\mathbf{z}):T_{\mathbf{z}}\mathcal{M}_{LH}(X,\mathcal{G}_{\theta})\longrightarrow T_{\mathbf{y}}\mathcal{M}_{H}(X,\mathcal{G}_{\theta})

is induced by the natural inclusion of complexes

𝒦,φ𝒦¯,φ.\mathcal{K}_{\mathcal{E},\varphi}\ \hookrightarrow\ \bar{\mathcal{K}}_{\mathcal{E},\varphi}.

By Serre duality in hypercohomology, the cotangent spaces are identified:

T𝐳LH(X,𝒢θ)1(𝒦,φ)andT𝐲H(X,𝒢θ)1(𝒦¯,φ).T^{*}_{\mathbf{z}}\mathcal{M}_{LH}(X,\mathcal{G}_{\theta})\cong\mathbb{H}^{1}\Bigl{(}\mathcal{K}_{\mathcal{E},\varphi}^{\vee}\Bigr{)}\quad\text{and}\quad T^{*}_{\mathbf{y}}\mathcal{M}_{H}(X,\mathcal{G}_{\theta})\cong\mathbb{H}^{1}\Bigl{(}\bar{\mathcal{K}}_{\mathcal{E},\varphi}^{\vee}\Bigr{)}.

The dual differential

d(𝐳):T𝐲H(X,𝒢θ)T𝐳LH(X,𝒢θ)d\ell(\mathbf{z})^{*}:T^{*}_{\mathbf{y}}\mathcal{M}_{H}(X,\mathcal{G}_{\theta})\longrightarrow T^{*}_{\mathbf{z}}\mathcal{M}_{LH}(X,\mathcal{G}_{\theta})

arises from the inclusion

𝒦¯,φ𝒦¯,φ.\bar{\mathcal{K}}_{\mathcal{E},\varphi}^{\vee}\ \hookrightarrow\ \bar{\mathcal{K}}_{\mathcal{E},\varphi}^{\vee}.

On LH(X,𝒢θ)\mathcal{M}_{LH}(X,\mathcal{G}_{\theta}) the symplectic form ω\omega induces the isomorphism

ω(,φ,η)1:T𝐳LH(X,𝒢θ)T𝐳LH(X,𝒢θ).\omega^{-1}_{(\mathcal{E},\varphi,\eta)}:T^{*}_{\mathbf{z}}\mathcal{M}_{LH}(X,\mathcal{G}_{\theta})\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}T_{\mathbf{z}}\mathcal{M}_{LH}(X,\mathcal{G}_{\theta}).

Meanwhile, the map \sharp on H(X,𝒢θ)\mathcal{M}_{H}(X,\mathcal{G}_{\theta}) is

:T𝐲H(X,𝒢θ)T𝐲H(X,𝒢θ)\sharp:T^{*}_{\mathbf{y}}\mathcal{M}_{H}(X,\mathcal{G}_{\theta})\longrightarrow T_{\mathbf{y}}\mathcal{M}_{H}(X,\mathcal{G}_{\theta})

which is induced by the inclusion 𝒦¯E,φ𝒦¯E,φ\bar{\mathcal{K}}_{E,\varphi}^{\vee}\ \hookrightarrow\ \bar{\mathcal{K}}_{E,\varphi}.

In order to prove that \ell is Poisson, we must show that for every wT𝐲H(X,𝒢θ)w\in T^{*}_{\mathbf{y}}\mathcal{M}_{H}(X,\mathcal{G}_{\theta}), the identity

d(𝐳)ω(,φ,η)1d(𝐳)(w)=(w)d\ell(\mathbf{z})\circ\omega^{-1}_{(\mathcal{E},\varphi,\eta)}\circ d\ell(\mathbf{z})^{*}(w)=\sharp(w) (5.4)

holds or, equivalently, that the diagram

T𝐲H(X,𝒢θ){T^{*}_{\mathbf{y}}\mathcal{M}_{H}(X,\mathcal{G}_{\theta})}T𝐲H(X,𝒢θ){T_{\mathbf{y}}\mathcal{M}_{H}(X,\mathcal{G}_{\theta})}T𝐳LH(X,𝒢θ){T^{*}_{\mathbf{z}}\mathcal{M}_{LH}(X,\mathcal{G}_{\theta})}T𝐳LH(X,𝒢θ){T_{\mathbf{z}}\mathcal{M}_{LH}(X,\mathcal{G}_{\theta})}\scriptstyle{\sharp}d(𝐳)\scriptstyle{d\ell(\mathbf{z})^{*}}ω(,φ,η)1\scriptstyle{\omega^{-1}_{(\mathcal{E},\varphi,\eta)}}d(𝐳)\scriptstyle{d\ell(\mathbf{z})}

commutes.

By the functoriality of hypercohomology, the maps in the above diagram are induced by the natural inclusions of the deformation complexes (and their duals). In particular, the map \sharp is induced by the inclusion

𝒦¯,φ𝒦¯,φ,\bar{\mathcal{K}}_{\mathcal{E},\varphi}^{\vee}\hookrightarrow\bar{\mathcal{K}}_{\mathcal{E},\varphi},

while the maps d(𝐳)d\ell(\mathbf{z}) and d(𝐳)d\ell(\mathbf{z})^{*} come from the inclusions

𝒦,φ𝒦¯,φand𝒦¯,φ𝒦,φ,\mathcal{K}_{\mathcal{E},\varphi}\hookrightarrow\bar{\mathcal{K}}_{\mathcal{E},\varphi}\quad\text{and}\quad\bar{\mathcal{K}}_{\mathcal{E},\varphi}^{\vee}\hookrightarrow\mathcal{K}_{\mathcal{E},\varphi}^{\vee},

respectively. Consequently, the composite map

d(𝐳)ω(,φ,η)1d(𝐳)d\ell(\mathbf{z})\circ\omega^{-1}_{(\mathcal{E},\varphi,\eta)}\circ d\ell(\mathbf{z})^{*}

is precisely the hypercohomology map induced by the inclusion 𝒦¯,φ𝒦¯,φ\bar{\mathcal{K}}_{\mathcal{E},\varphi}^{\vee}\hookrightarrow\bar{\mathcal{K}}_{\mathcal{E},\varphi}, which is exactly the map \sharp. This establishes the identity (5.4), and shows that \ell is a Poisson map and that \sharp is Poisson on the open dense subset of H(X,𝒢θ)\mathcal{M}_{H}(X,\mathcal{G}_{\theta}). As so, the map \sharp extends to a Poisson structure on all H(X,𝒢θ)\mathcal{M}_{H}(X,\mathcal{G}_{\theta}). ∎

Remark 5.21.

Note that in [23], a Poisson structure on moduli spaces of twisted stable GG-Higgs bundles over stacky curves was obtained using an Atiyah sequence. However, that construction was not determining a canonical moment map on the cotangent bundle that could be used in order to describe the symplectic leaves of the Poisson manifold.

6. Parahoric Hitchin Fibration and Abelianization

We next study the Hitchin fibration on the space of logahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-Higgs torsors. The main result here is that the generic fibers of this fibration are Lagrangian with respect to the symplectic leaves of the Poisson space H(X,𝒢𝜽)\mathcal{M}_{H}(X,\mathcal{G}_{\boldsymbol{\theta}}).

6.1. Hitchin Fibration

Any invariant homogeneous degree ii-polynomial naturally defines a map

ai:H0((𝔤)KX(D))H0(K(D)i).a_{i}:H^{0}(\mathcal{E}(\mathfrak{g})\otimes K_{X}(D))\to H^{0}(K(D)^{i}).

Equivalence classes of Higgs bundles on the moduli space H(X,𝒢𝜽)\mathcal{M}_{H}(X,\mathcal{G}_{\boldsymbol{\theta}}) of logahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-Higgs torsors are defined using the adjoint action of G(K)G(K) on 𝔤(K)\mathfrak{g}(K), the loop Lie algebra of G(K)G(K), which is induced from the action of the complex reductive group GG on 𝔤\mathfrak{g}. If the Lie algebra 𝔤(K)\mathfrak{g}(K) has rank ll and pip_{i} are polynomials of degree degpi=mi+1\text{deg}p_{i}=m_{i}+1, for i=1,,li=1,...,l, forming a basis of the algebra of invariant polynomials on the Lie algebra 𝔤(K)\mathfrak{g}(K), then the corresponding maps aia_{i} combine to give a Hitchin fibration

h𝜽:H(X,𝒢𝜽)𝜽,h_{\boldsymbol{\theta}}:\mathcal{M}_{H}(X,\mathcal{G}_{\boldsymbol{\theta}})\to\mathcal{B}_{\boldsymbol{\theta}},

defined by h𝜽(,φ)=(p1(φ),,pl(φ))h_{\boldsymbol{\theta}}(\mathcal{E},\varphi)=(p_{1}(\varphi),...,p_{l}(\varphi)). Note that as in the parabolic case, the map h𝜽h_{\boldsymbol{\theta}} is blind to the parahoric structure at each parahoric point in the divisor DD, as it only depends on the Higgs field φH0((𝔤)KX(D))\varphi\in H^{0}(\mathcal{E}(\mathfrak{g})\otimes K_{X}(D)) and the line bundle KX(D)K_{X}(D).

We set the following definition (cf. [28], [35]):

Definition 6.1 (Hitchin Base and its Image).

The (ambient) parahoric Hitchin base is the affine space

𝜽=j=1lH0(X,(KX(D))mj+1).\mathcal{B}_{\boldsymbol{\theta}}=\bigoplus_{j=1}^{l}H^{0}(X,(K_{X}(D))^{m_{j}+1}).

The image of the Hitchin fibration, denoted by 𝒜𝜽\mathcal{A}_{\boldsymbol{\theta}}, is the subvariety

𝒜𝜽=h𝜽(H(X,𝒢𝜽))𝜽.\mathcal{A}_{\boldsymbol{\theta}}=h_{\boldsymbol{\theta}}(\mathcal{M}_{H}(X,\mathcal{G}_{\boldsymbol{\theta}}))\subset\mathcal{B}_{\boldsymbol{\theta}}.

6.2. Cameral covers

We now consider the construction of cameral covers of XX and the generalized Prym varieties adapting the original construction of Donagi for principal GG-Higgs bundles to the case of logahoric 𝒢𝜽\mathcal{G}_{{\boldsymbol{\theta}}}-Higgs torsors; see [14, Section 2] for a survey in the principal bundle case.

Let TGT\subset G be a maximal torus of GG. Chevalley’s theorem provides that the restriction map

[𝔤]G[𝔱]W\mathbb{C}[\mathfrak{g}]^{G}\to\mathbb{C}[\mathfrak{t}]^{W}

is an isomorphism from Ad-invariant polynomial functions on the Lie algebra 𝔤\mathfrak{g} to WW-invariant polynomial functions on the Cartan subalgebra 𝔱𝔤\mathfrak{t}\subset\mathfrak{g}. We then consider the injective ring homomorphism

[𝔱]W[𝔤]G[𝔤]\mathbb{C}[\mathfrak{t}]^{W}\cong\mathbb{C}[\mathfrak{g}]^{G}\hookrightarrow\mathbb{C}[\mathfrak{g}]

and take the prime spectrum of the rings to define a surjective GG-invariant morphism of affine varieties

𝔤𝔱/W.\mathfrak{g}\twoheadrightarrow\mathfrak{t}/{W}. (6.1)

Thus, taking fiber product with the quotient map 𝔱𝔱/W\mathfrak{t}\to\mathfrak{t}/W, we get

𝔤~:=𝔤×𝔱/W𝔱,\tilde{\mathfrak{g}}:=\mathfrak{g}\times_{\mathfrak{t}/W}\mathfrak{t}, (6.2)

and the projection π:𝔤~𝔤\pi:\tilde{\mathfrak{g}}\to\mathfrak{g} is a finite WW-Galois morphism. This is called the cameral cover of the Lie algebra 𝔤\mathfrak{g}. The fiber π1(g)\pi^{-1}(g) of a regular semisimple element g𝔤g\in\mathfrak{g} is identified with the set of chambers in 𝔱\mathfrak{t}^{*}.

Now let (,φ)(\mathcal{E},\varphi) be a logahoric 𝒢𝜽\mathcal{G}_{{\boldsymbol{\theta}}}-Higgs torsor over XX. The Higgs field φ\varphi is a holomorphic section of (𝔤)KX(D)\mathcal{E}(\mathfrak{g})\otimes{{K}_{X}}(D). Since the morphism (6.1) is GG-invariant and \mathbb{C}^{*}-equivariant, it can be extended to a morphism |(𝔤)KX(D)||𝔱KX(D)|/W\left|\mathcal{E}(\mathfrak{g})\otimes{{K}_{X}}(D)\right|\to\left|\mathfrak{t}\otimes{{K}_{X}}(D)\right|/W, where ||\left|\,\cdot\,\right| denotes here the total space. Forming the fiber product with 𝔱KX(D)\mathfrak{t}\otimes{{K}_{X}}(D) as in (6.2) and pulling-back to XX by the Higgs field φ\varphi, we get the following definition analogously to [14, Definition 2.6]:

Definition 6.2.

The cameral cover of XX determined by the logahoric 𝒢𝜽\mathcal{G}_{{\boldsymbol{\theta}}}-Higgs torsor (,φ)(\mathcal{E},\varphi) is defined by the projection π:X~X\pi:\tilde{X}\to X, where

X~=φ(|(𝔤)KX(D)|×|𝔱KX(D)|/W|𝔱KX(D)|)\tilde{X}={{\varphi}^{*}}\left(\left|\mathcal{E}(\mathfrak{g})\otimes{{K}_{X}}(D)\right|{{\times}_{\left|\mathfrak{t}\otimes{{K}_{X}}(D)\right|/W}}\left|\mathfrak{t}\otimes{{K}_{X}}(D)\right|\right)

and π\pi is the projection onto the first factor.

The cameral cover X~X\tilde{X}\to X is a WW-Galois cover which generically parameterizes the chambers determined by the Higgs field φ\varphi; we are pulling back by φ\varphi the covers 𝔤~𝔤\tilde{\mathfrak{g}}\to\mathfrak{g} to take covers of open subsets in XX over which the bundle is trivialized, and then we glue these covers together. The ramification of X~\tilde{X} is determined by the order of the zeroes of φ\varphi and X~\tilde{X} is a closed subscheme of |(𝔤)KX(D)|\left|\mathcal{E}(\mathfrak{g})\otimes{{K}_{X}}(D)\right| that can be singular or non-reduced. The cameral cover inherits from |(𝔤)KX(D)|\left|\mathcal{E}(\mathfrak{g})\otimes{{K}_{X}}(D)\right| a WW-action thus has lots of automorphisms. Generically, X~\tilde{X} is a non-singular Galois WW-cover with simple ramification.

Definition 6.3.

We shall denote by 𝒜𝒜𝜽\mathcal{A}^{\prime}\subset\mathcal{A}_{\boldsymbol{\theta}} the open and dense subspace of 𝒜𝜽\mathcal{A}_{\boldsymbol{\theta}} such that X~s\tilde{X}_{s} is smooth whenever s𝒜s\in\mathcal{A}^{\prime}. We call the fibers h𝜽1(s)h_{\boldsymbol{\theta}}^{-1}(s) for s𝒜s\in\mathcal{A}^{\prime}, the generic fibers of the parahoric Hitchin fibration h𝜽h_{\boldsymbol{\theta}}.

6.3. Generalized Prym varieties

We have constructed the cameral cover X~\tilde{X} as a WW-Galois cover of XX. Thus, there is an induced action of [W]\mathbb{Z}[W] on X~\tilde{X}, hence on H(X~,)H_{*}(\tilde{X},\mathbb{Z}) and on the Picard group Pic(X~)\text{Pic}(\tilde{X}).

Definition 6.4.

For an irreducible [W]\mathbb{Z}[W]-module Λ\Lambda, the generalized Prym variety of X~\tilde{X} is defined as the set of equivariant maps of the [W]\mathbb{Z}[W]-module Λ\Lambda to PicX~\text{Pic}\tilde{X}

PrymΛ(X~):=HomW(Λ,PicX~).Prym_{\Lambda}(\tilde{X}):=\text{Hom}_{W}(\Lambda,\text{Pic}\tilde{X}).

The generalized Prym variety PrymΛ(X~)Prym_{\Lambda}(\tilde{X}) is an algebraic group. For a generic point s𝒜s\in\mathcal{A}^{\prime}, the cameral cover X~\tilde{X} is a smooth projective curve, in which case then PicX~\text{Pic}\tilde{X} is an abelian variety, therefore PrymΛ(X~)Prym_{\Lambda}(\tilde{X}) is also abelian; we refer to [14, Section 5] for further information. We thus have:

Proposition 6.5.

The generalized Prym varieties associated to the cameral cover XsX_{s}, for generic s𝒜s\in\mathcal{A}^{\prime}, are abelian varieties.

6.4. The Hitchin fibration for Γ\Gamma-equivariant GG-Higgs bundles

Let XX be a smooth complex projective curve. Suppose that a second curve YY is equipped with an effective action of the finite abelian group

Γ=i=1r/ni,\Gamma=\prod_{i=1}^{r}\mathbb{Z}/n_{i}\mathbb{Z},

so that there is a corresponding Γ\Gamma-Galois cover

π:YX,X=Y/Γ.\pi\colon Y\to X,\qquad X=Y/\Gamma.

Equivalently, one may work on the root stack over XX. In what follows, we study the moduli space HΓ(Y,G)\mathcal{M}^{\Gamma}_{H}(Y,G) of semistable Γ\Gamma-equivariant GG-Higgs bundles on YY, where GG is a reductive algebraic group over \mathbb{C} with Lie algebra 𝔤\mathfrak{g}. In particular,

HΓ(Y,G)={(P,s):P is a Γ-equivariant G-bundle on YsH0(Y,adPKY(D~))};\mathcal{M}^{\Gamma}_{H}(Y,G)=\Bigl{\{}(P,s):\,P\mbox{ is a $\Gamma$-equivariant $G$-bundle on $Y$, }\,s\in H^{0}\bigl{(}Y,\operatorname{ad}P\otimes K_{Y}(\tilde{D})\bigr{)}\Bigr{\}};

we refer to [24] for the construction of this moduli space.

Let KY(D~)K_{Y}(\tilde{D}) denote the canonical bundle on YY twisted by the parahoric divisor D~\tilde{D}. By deformation theory and Serre duality, a point in the cotangent bundle THΓ(Y,G)T^{*}\mathcal{M}^{\Gamma}_{H}(Y,G) is a pair (P,s)(P,s) with PP a stable Γ\Gamma-equivariant principal GG-bundle on YY and

sH0(Y,adPKY(D~))s\in H^{0}\Bigl{(}Y,\,\operatorname{ad}P\otimes K_{Y}(\tilde{D})\Bigr{)}

a (necessarily Γ\Gamma-invariant) section.

Let h1,,hkh_{1},\dots,h_{k} be homogeneous generators of the algebra of invariant polynomials on 𝔤\mathfrak{g} with deghi=:di\deg h_{i}=:d_{i}. Each hih_{i} induces a map

i:adPKY(D~)(KY(D~))di,\mathcal{H}_{i}\colon\operatorname{ad}P\otimes K_{Y}(\tilde{D})\to\Bigl{(}K_{Y}(\tilde{D})\Bigr{)}^{d_{i}},

and consequently one defines the Hitchin fibration

:THΓ(Y,G)𝒦:=i=1kH0(Y,(KY(D~))di)\mathcal{H}\colon T^{*}\mathcal{M}^{\Gamma}_{H}(Y,G)\to\mathcal{K}:=\bigoplus_{i=1}^{k}H^{0}\Bigl{(}Y,\Bigl{(}K_{Y}(\tilde{D})\Bigr{)}^{d_{i}}\Bigr{)} (6.3)

by

(P,s)ϕ=(ϕ1,,ϕk),ϕi=i(s).(P,s)\mapsto\phi=(\phi_{1},\dots,\phi_{k}),\quad\phi_{i}=\mathcal{H}_{i}(s).

Fix a maximal torus TGT\subset G with associated root system =(G,T)\mathcal{R}=\mathcal{R}(G,T) and Weyl group W=NG(T)/T.W=N_{G}(T)/T. Choose a Borel subgroup BTB\supset T; this determines a set of positive roots +\mathcal{R}^{+}\subset\mathcal{R}. Denote by 𝔱\mathfrak{t} the Lie algebra of TT. Then the differential of each root α\alpha gives a map

dα:𝔱KY(D~)KY(D~).d\alpha\colon\mathfrak{t}\otimes K_{Y}(\tilde{D})\to K_{Y}(\tilde{D}).

Moreover, restricting the invariant polynomials hih_{i} to 𝔱\mathfrak{t} yields homogeneous, WW-invariant polynomials σ1,,σk\sigma_{1},\dots,\sigma_{k}, which define the Galois covering

σ¯=(σ1,,σk):𝔱KY(D~)i=1k(KY(D~))di.\underline{\sigma}=(\sigma_{1},\dots,\sigma_{k})\colon\mathfrak{t}\otimes K_{Y}(\tilde{D})\to\bigoplus_{i=1}^{k}\Bigl{(}K_{Y}(\tilde{D})\Bigr{)}^{d_{i}}.

Its discriminant Ξ\Xi is the zero locus of the WW-invariant function αRdα.\prod_{\alpha\in R}d\alpha.

For generic ϕ𝒦\phi\in\mathcal{K}, we define the cameral cover of YY by

Y~:=ϕ(𝔱KY(D~)).\widetilde{Y}:=\phi^{*}\Bigl{(}\mathfrak{t}\otimes K_{Y}(\tilde{D})\Bigr{)}.

Then Y~\widetilde{Y} is a ramified covering of YY with m=|W|m=|W| sheets. Denote by RamYRam\subset Y the branch locus; by construction one has

𝒪(Ram)(KY(D~))|R|(KY(D~))dimGrankG.\mathcal{O}(Ram)\cong\Bigl{(}K_{Y}(\tilde{D})\Bigr{)}^{|R|}\equiv\Bigl{(}K_{Y}(\tilde{D})\Bigr{)}^{\dim G-\operatorname{rank}G}. (6.4)

Let

ι:Y~𝔱KY(D~)\iota\colon\widetilde{Y}\to\mathfrak{t}\otimes K_{Y}(\tilde{D})

be the natural inclusion. By definition, for each wWw\in W

ι(wη)=Ad(nw)ι(η),\iota(w\,\eta)=\operatorname{Ad}(n_{w})\,\iota(\eta),

where nwNG(T)n_{w}\in N_{G}(T) is any representative of ww. Moreover, if π:Y~Y\pi\colon\widetilde{Y}\to Y is the projection, then for each αR\alpha\in R the composition dαιd\alpha\circ\iota is a holomorphic section of πKY(D~)\pi^{*}K_{Y}(\tilde{D}).

These relations are summarized in the commutative diagram

Y~ι𝔱KY(D~)πYϕi=1k(KY(D~))di.\begin{array}[]{rccc}&\widetilde{Y}&\stackrel{{\scriptstyle\iota}}{{\longrightarrow}}&\mathfrak{t}\otimes K_{Y}(\tilde{D})\\ &\scriptstyle{\pi}\ \downarrow&&\downarrow\\ &Y&\stackrel{{\scriptstyle\phi}}{{\longrightarrow}}&\displaystyle\bigoplus_{i=1}^{k}\Bigl{(}K_{Y}(\tilde{D})\Bigr{)}^{d_{i}}\,.\end{array}

Under our genericity assumptions the cover Y~\widetilde{Y} is smooth and irreducible, and each ramification point pπ1(Ram)p\in\pi^{-1}(Ram) is simple (i.e. the section

αR+(dαι):Y~πKY(D~)|R|/2\prod_{\alpha\in R^{+}}\Bigl{(}d\alpha\circ\iota\Bigr{)}\colon\widetilde{Y}\to\pi^{*}K_{Y}(\tilde{D})^{|R|/2}

has a simple zero at pp).

Let X(T)X(T) denote the group of characters of TT. The group of isomorphism classes of holomorphic Γ\Gamma-equivariant principal TT-bundles on Y~\widetilde{Y} is identified with the group

PicΓ(Y~)X(T),\operatorname{Pic}^{\Gamma}(\widetilde{Y})\otimes X(T)^{*},

where

X(T)=Hom(X(T),)X(T)^{*}=\operatorname{Hom}(X(T),\mathbb{Z})

is the dual group. Similarly, the group of topologically trivial Γ\Gamma-equivariant TT-bundles is

J(Y~)X(T),J(\widetilde{Y})\otimes X(T)^{*},

with JΓ(Y~)J^{\Gamma}(\widetilde{Y}) the Jacobian of Y~\widetilde{Y}. The Weyl group WW acts naturally on both JΓ(Y~)J^{\Gamma}(\widetilde{Y}) and X(T)X(T)^{*} (on the latter by conjugation). A TT-bundle is written as

τ=D1χ1++Dlχl,\tau=D_{1}\otimes\chi_{1}+\cdots+D_{l}\otimes\chi_{l},

where each DiD_{i}, i=1,,li=1,...,l is a divisor on Y~\tilde{Y} representing a point in the Jacobian, and the χi\chi_{i}’s are cocharacters. Writing each divisor DiD_{i} as

Di=pY~npp,D_{i}=\sum_{p\in\tilde{Y}}n_{p}\,p,

the action of an element wWw\in W on DiD_{i} is defined by

wDi:=pY~np[w(p)].w\cdot D_{i}:=\sum_{p\in\tilde{Y}}n_{p}\,\bigl{[}w(p)\bigr{]}.

In other words, ww acts by sending each point pp in the support of DiD_{i} to w(p)w(p), preserving the multiplicities. This operation induces a natural WW-action on the divisor group Div(Y~)\operatorname{Div}(\tilde{Y}) and, hence, on the Picard group Pic(Y~)\operatorname{Pic}(\tilde{Y}). On the other hand, the twisted character χiw{}^{w}\chi_{i} is given by

χiw(t)=χi(w1tw).{}^{w}\chi_{i}(t)=\chi_{i}\bigl{(}w^{-1}t\,w\bigr{)}. (6.5)

We have for any wWw\in W an action

τw=wD1χ1w++wDlχlw.{}^{w}\tau=wD_{1}\otimes{}^{w}\chi_{1}+\cdots+wD_{l}\otimes{}^{w}\chi_{l}.

A critical aspect in order to introduce the definition of the generalized Prym variety PrymΓ(Y)\operatorname{Prym}^{\Gamma}(Y) on YY (Definition 6.8 below) and the subsequent abelianization result (Theorem 6.9) is the compatibility between the Γ\Gamma-equivariant structure on TT-bundles over Y~\widetilde{Y} and the natural action of the Weyl group WW on these bundles. We need to ensure that if a TT-bundle is Γ\Gamma-equivariant, its WW-transforms are also Γ\Gamma-equivariant in a consistent manner.

Lemma 6.6.

The Γ\Gamma-action on YY lifts to an action on the cameral cover ϖ:Y~Y\varpi:\widetilde{Y}\to Y, denoted by γ¯:Y~Y~\bar{\gamma}:\widetilde{Y}\to\widetilde{Y} for γΓ\gamma\in\Gamma. This lifted Γ\Gamma-action commutes with the sheet-permuting action of the Weyl group WW on Y~\widetilde{Y}, denoted by w¯:Y~Y~\bar{w}:\widetilde{Y}\to\widetilde{Y} for wWw\in W. That is, for all pY~p\in\widetilde{Y}, γ¯(w¯(p))=w¯(γ¯(p))\bar{\gamma}(\bar{w}(p))=\bar{w}(\bar{\gamma}(p)).

Proof.

The Γ\Gamma-equivariant GG-Higgs bundle (P,s)(P,s) on YY features a Γ\Gamma-invariant Higgs field sH0(Y,adPKY(D~))s\in H^{0}(Y,\mathrm{ad}P\otimes K_{Y}(\tilde{D})). Consequently, the associated characteristic polynomial sections ϕ=(ϕ1,,ϕk)\phi=(\phi_{1},\dots,\phi_{k}), which define the Hitchin map (P,s)=ϕ\mathcal{H}(P,s)=\phi, are Γ\Gamma-invariant: δϕ=ϕ\delta^{*}\phi=\phi for all δΓ\delta\in\Gamma, where δ\delta^{*} is the pullback on sections over YY. The cameral cover is Y~={(y,ψy)yY,ψy(𝔱KY(D~))y,σ¯(ψy)=ϕ(y)}\widetilde{Y}=\{(y,\psi_{y})\mid y\in Y,\psi_{y}\in(\mathfrak{t}\otimes K_{Y}(\tilde{D}))_{y},\underline{\sigma}(\psi_{y})=\phi(y)\}. For δΓ\delta\in\Gamma, define its action on p=(y,ψy)Y~p=(y,\psi_{y})\in\widetilde{Y} by δ¯(p)=(δy,δψy)\bar{\delta}(p)=(\delta\cdot y,\delta_{*}\psi_{y}), where δψy\delta_{*}\psi_{y} represents the natural transformation of the fiber element ψy\psi_{y} under the action of δ\delta (which acts on YY and the bundle KY(D~)K_{Y}(\tilde{D})). Since σ¯\underline{\sigma} is GG-invariant (acting on the 𝔱\mathfrak{t}-component) and ϕ\phi is Γ\Gamma-invariant, if (y,ψy)Y~(y,\psi_{y})\in\widetilde{Y}, then

σ¯(δψδ1y)=δ(σ¯(ψδ1y))=δ(ϕ(δ1y))=(δϕ)(y)=ϕ(y).\underline{\sigma}(\delta_{*}\psi_{\delta^{-1}y})=\delta_{*}(\underline{\sigma}(\psi_{\delta^{-1}y}))=\delta_{*}(\phi(\delta^{-1}y))=(\delta_{*}\phi)(y)=\phi(y).

Hence, δ¯(p)Y~\bar{\delta}(p)\in\widetilde{Y}, so Y~\widetilde{Y} admits a Γ\Gamma-action covering the Γ\Gamma-action on YY.

The Weyl group WW acts on p=(y,ψy)Y~p=(y,\psi_{y})\in\widetilde{Y} by w¯(p)=(y,wψy)\bar{w}(p)=(y,w\cdot\psi_{y}), where wψyw\cdot\psi_{y} is the standard WW-action on the 𝔱\mathfrak{t}-component of ψy𝔱(KY(D~))y\psi_{y}\in\mathfrak{t}\otimes(K_{Y}(\tilde{D}))_{y}. This action fixes yYy\in Y.

To show commutativity, consider a point p=(y,ψy)p=(y,\psi_{y}). Then,

(δ¯w¯)(p)\displaystyle(\bar{\delta}\circ\bar{w})(p) =δ¯(y,wψy)=(δy,δ(wψy)), and\displaystyle=\bar{\delta}(y,w\cdot\psi_{y})=(\delta\cdot y,\delta_{*}(w\cdot\psi_{y})),\text{ and}
(w¯δ¯)(p)\displaystyle(\bar{w}\circ\bar{\delta})(p) =w¯(δy,δψy)=(δy,w(δψy)).\displaystyle=\bar{w}(\delta\cdot y,\delta_{*}\psi_{y})=(\delta\cdot y,w\cdot(\delta_{*}\psi_{y})).

Equality holds if δ(wψy)=w(δψy)\delta_{*}(w\cdot\psi_{y})=w\cdot(\delta_{*}\psi_{y}). The term ψy\psi_{y} can be locally written as jcjtjαj(y)\sum_{j}c_{j}t_{j}\otimes\alpha_{j}(y), where tj𝔱t_{j}\in\mathfrak{t} and αj(y)(KY(D~))y\alpha_{j}(y)\in(K_{Y}(\tilde{D}))_{y}. The action wψy=jcj(wtj)αj(y)w\cdot\psi_{y}=\sum_{j}c_{j}(w\cdot t_{j})\otimes\alpha_{j}(y). Then δ(wψy)=jcj(wtj)(δαj)(δy)\delta_{*}(w\cdot\psi_{y})=\sum_{j}c_{j}(w\cdot t_{j})\otimes(\delta_{*}\alpha_{j})(\delta\cdot y). Conversely, δψy=jcjtj(δαj)(δy)\delta_{*}\psi_{y}=\sum_{j}c_{j}t_{j}\otimes(\delta_{*}\alpha_{j})(\delta\cdot y). Then w(δψy)=jcj(wtj)(δαj)(δy)w\cdot(\delta_{*}\psi_{y})=\sum_{j}c_{j}(w\cdot t_{j})\otimes(\delta_{*}\alpha_{j})(\delta\cdot y). The two expressions are identical because the WW-action only affects the 𝔱\mathfrak{t}-coefficients and the δ\delta_{*}-action affects the KY(D~)K_{Y}(\tilde{D}) coefficients and the base point. Thus, δ¯w¯=w¯δ¯\bar{\delta}\circ\bar{w}=\bar{w}\circ\bar{\delta} as automorphisms of Y~\widetilde{Y}. ∎

Proposition 6.7.

Let \mathcal{L} be a Γ\Gamma-equivariant TT-bundle on Y~\widetilde{Y}. For any wWw\in W, the transformed TT-bundle w{}^{w}\mathcal{L} carries a natural Γ\Gamma-equivariant structure inherited from \mathcal{L}. Consequently, the WW-action is well-defined on the set of isomorphism classes of Γ\Gamma-equivariant TT-bundles on Y~\widetilde{Y}.

Proof.

A TT-bundle \mathcal{L} on Y~\widetilde{Y} is Γ\Gamma-equivariant if for each δΓ\delta\in\Gamma, there is an isomorphism uδ:(δ¯1)u_{\delta}:(\bar{\delta}^{-1})^{*}\mathcal{L}\stackrel{{\scriptstyle\sim}}{{\to}}\mathcal{L} such that uδ1δ2=uδ1(δ1¯1)uδ2u_{\delta_{1}\delta_{2}}=u_{\delta_{1}}\circ(\bar{\delta_{1}}^{-1})^{*}u_{\delta_{2}}, for all δ1,δ2Γ\delta_{1},\delta_{2}\in\Gamma. The WW-transform is w=(w¯1)χw{}^{w}\mathcal{L}=(\bar{w}^{-1})^{*}\mathcal{L}\otimes\chi_{w}, where χw\chi_{w} represents the action of ww on the characters defining the TT-structure (denoted by χiw{}^{w}\chi_{i} in (6.5)). For simplicity of notation for the geometric part, let w:=(w¯1)\mathcal{L}_{w}:=(\bar{w}^{-1})^{*}\mathcal{L}.

We define a Γ\Gamma-equivariant structure vδ:(δ¯1)wwv_{\delta}:(\bar{\delta}^{-1})^{*}\mathcal{L}_{w}\stackrel{{\scriptstyle\sim}}{{\to}}\mathcal{L}_{w} for a given w\mathcal{L}_{w}. Using Lemma 6.6, we have δ¯1w¯1=w¯1δ¯1\bar{\delta}^{-1}\circ\bar{w}^{-1}=\bar{w}^{-1}\circ\bar{\delta}^{-1}. Thus,

(δ¯1)w=(δ¯1)(w¯1)=((w¯δ¯)1)=((δ¯w¯)1)=(w¯1)(δ¯1).(\bar{\delta}^{-1})^{*}\mathcal{L}_{w}=(\bar{\delta}^{-1})^{*}(\bar{w}^{-1})^{*}\mathcal{L}=((\bar{w}\circ\bar{\delta})^{-1})^{*}\mathcal{L}=((\bar{\delta}\circ\bar{w})^{-1})^{*}\mathcal{L}=(\bar{w}^{-1})^{*}(\bar{\delta}^{-1})^{*}\mathcal{L}.

Define vδ:=(w¯1)uδ:(w¯1)(δ¯1)(w¯1)=wv_{\delta}:=(\bar{w}^{-1})^{*}u_{\delta}:(\bar{w}^{-1})^{*}(\bar{\delta}^{-1})^{*}\mathcal{L}\stackrel{{\scriptstyle\sim}}{{\to}}(\bar{w}^{-1})^{*}\mathcal{L}=\mathcal{L}_{w}. This map vδv_{\delta} is an isomorphism (δ¯1)ww(\bar{\delta}^{-1})^{*}\mathcal{L}_{w}\stackrel{{\scriptstyle\sim}}{{\to}}\mathcal{L}_{w}. We verify the cocycle condition for vδv_{\delta}:

vδ1δ2\displaystyle v_{\delta_{1}\delta_{2}} =(w¯1)uδ1δ2\displaystyle=(\bar{w}^{-1})^{*}u_{\delta_{1}\delta_{2}}
=(w¯1)(uδ1(δ1¯1)uδ2)(by cocycle condition for u)\displaystyle=(\bar{w}^{-1})^{*}(u_{\delta_{1}}\circ(\bar{\delta_{1}}^{-1})^{*}u_{\delta_{2}})\quad(\text{by cocycle condition for }u)
=((w¯1)uδ1)((w¯1)(δ1¯1)uδ2)\displaystyle=((\bar{w}^{-1})^{*}u_{\delta_{1}})\circ((\bar{w}^{-1})^{*}(\bar{\delta_{1}}^{-1})^{*}u_{\delta_{2}})
=vδ1((δ1¯1)(w¯1)uδ2)(using w¯1δ1¯1=δ1¯1w¯1)\displaystyle=v_{\delta_{1}}\circ((\bar{\delta_{1}}^{-1})^{*}(\bar{w}^{-1})^{*}u_{\delta_{2}})\quad(\text{using }\bar{w}^{-1}\bar{\delta_{1}}^{-1}=\bar{\delta_{1}}^{-1}\bar{w}^{-1})
=vδ1(δ1¯1)vδ2.\displaystyle=v_{\delta_{1}}\circ(\bar{\delta_{1}}^{-1})^{*}v_{\delta_{2}}.

Thus, {vδ}δΓ\{v_{\delta}\}_{\delta\in\Gamma} defines a Γ\Gamma-equivariant structure on w=(w¯1)\mathcal{L}_{w}=(\bar{w}^{-1})^{*}\mathcal{L}. The full bundle w=wχw{}^{w}\mathcal{L}=\mathcal{L}_{w}\otimes\chi_{w} is then also Γ\Gamma-equivariant, assuming Γ\Gamma acts trivially on the abstract character group X(T)X(T) (which is standard, as TT is a fixed group). Therefore, the condition w{}^{w}\mathcal{L}\cong\mathcal{L} in the definition of PrymΓ(Y)\operatorname{Prym}^{\Gamma}(Y) can be understood as an isomorphism of Γ\Gamma-equivariant TT-bundles. ∎

We can now introduce the following:

Definition 6.8.

The generalized Prym variety with respect to the Γ\Gamma-action is defined as

PrymΓ(Y):=[JΓ(Y~)X(T)]W,\operatorname{Prym}^{\Gamma}(Y):=\Bigl{[}J^{\Gamma}(\widetilde{Y})\otimes X(T)^{*}\Bigr{]}^{W},

i.e. the subgroup of those topologically trivial TT-bundles τ\tau on Y~\widetilde{Y} satisfying

τwτ,for all wW.{}^{w}\tau\cong\tau,\quad\text{for all }w\in W.

Note that PrymΓ(Y)\operatorname{Prym}^{\Gamma}(Y) is an algebraic group whose null connected component PrymΓ(Y)0\operatorname{Prym}^{\Gamma}(Y)_{0} is an abelian variety.

The generic fibers of the Hitchin fibration in the case of stable principal GG-bundles on a compact Riemann surface were studied by Faltings in [16]. Since we have seen that the Γ\Gamma-equivariant structure on TT-bundles and the natural Weyl group action on these bundles are compatible, the constructions of the cameral cover and the generalized Prym variety proceed as in the non-equivariant principal GG-bundle case. Moreover, the proof that the generic Hitchin fibers are isomorphic to generalized Prym varieties by Scognamillo [31] (simplifying the original proof of Faltings) repeats word by word in the presence of a Γ\Gamma-equivariant action on the principal TT-bundle as above. We thus conclude to the following:

Theorem 6.9 (Abelianization of Γ\Gamma-Equivariant GG-Higgs Bundles).

Let (P,s)(P,s) be a stable Γ\Gamma-equivariant GG-Higgs bundle on YY. Assume that the Hitchin base is generic so that the associated cameral cover π:Y~Y\pi\colon\widetilde{Y}\to Y is smooth. Then there exists a canonical construction of a Γ\Gamma-equivariant TT-bundle

𝒯(P,s)PicΓ(Y~)X(T)\mathcal{T}(P,s)\in\operatorname{Pic}^{\Gamma}(\widetilde{Y})\otimes X^{*}(T)

satisfying

𝒯w(P,s)𝒯(P,s),for all wW.{}^{w}\mathcal{T}(P,s)\cong\mathcal{T}(P,s),\quad\text{for all }w\in W.

In particular, the assignment (P,s)𝒯(P,s)(P,s)\longmapsto\mathcal{T}(P,s) defines an injective morphism from (each connected component of) the Hitchin fiber

1(ϕ)HΓ(Y,G)\mathcal{H}^{-1}(\phi)\subset\mathcal{M}_{H}^{\Gamma}(Y,G)

to the generalized Prym variety

PrymΓ(Y)={QPicΓ(Y~)X(T)|QwQ for all wW}.\operatorname{Prym}^{\Gamma}(Y)=\Bigl{\{}Q\in\operatorname{Pic}^{\Gamma}(\widetilde{Y})\otimes X^{*}(T)\;\Bigm{|}\;{}^{w}Q\cong Q\text{ for all }w\in W\Bigr{\}}.

Thus, the generic Hitchin fiber is an abelian torsor.

In view of the correspondence between the moduli space of logahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-Higgs torsors on XX and the Γ\Gamma-equivariant GG-Higgs bundles on YY from [24, Theorem 3.7], we now have the following:

Corollary 6.10.

The generic Hitchin fibers of the parahoric Hitchin fibration

h𝜽:H(X,𝒢𝜽)𝒜𝜽h_{\boldsymbol{\theta}}:\mathcal{M}_{H}(X,\mathcal{G}_{\boldsymbol{\theta}})\to\mathcal{A}_{\boldsymbol{\theta}}

are abelian torsors.

6.5. Regular Centralizer

This section establishes the duality that underlies the complete integrability of the parahoric Hitchin system. We begin with a local analysis of the regular centralizer in the parahoric context, which we then globalize to the moduli space of logahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-Higgs torsors. The section culminates in showing that the generic fibers of the parahoric Hitchin fibration are Lagrangian.

In the parahoric setting, for a given weight θ\theta, we work with the parahoric group scheme 𝒢θ\mathcal{G}_{\theta} whose Lie algebra is 𝔤θ\mathfrak{g}_{\theta}. For an element φ𝔤θ\varphi\in\mathfrak{g}_{\theta}^{*}, the parahoric centralizer is the subgroup

C𝒢θ(φ):={g𝒢θAd(g)(φ)=φ},C_{\mathcal{G}_{\theta}}(\varphi):=\{g\in\mathcal{G}_{\theta}\mid\operatorname{Ad}^{*}(g)(\varphi)=\varphi\},

with Lie algebra

𝔤θφ:={X𝔤θad(X)(φ)=0}.\mathfrak{g}_{\theta}^{\varphi}:=\{X\in\mathfrak{g}_{\theta}\mid\operatorname{ad}^{*}(X)(\varphi)=0\}.

These assemble into a group scheme C𝔤θ𝔤θC_{\mathfrak{g}_{\theta}}\to\mathfrak{g}_{\theta}^{*}. The regular locus is the intersection (𝔤θ)reg:=𝔤θ(𝔤)reg(\mathfrak{g}_{\theta}^{*})^{\mathrm{reg}}:=\mathfrak{g}_{\theta}^{*}\cap(\mathfrak{g}^{*})^{\mathrm{reg}}. The universal centralizer JJ over the Chevalley basis 𝔠:=𝔱/W\mathfrak{c}^{*}:=\mathfrak{t}^{*}/W pulls back via the Chevalley map χ:𝔤𝔠\chi\colon\mathfrak{g}^{*}\to\mathfrak{c}^{*} to give an isomorphism over the regular locus:

χJ|(𝔤θ)regC𝒢θ|(𝔤θ)reg.\chi^{*}J\big{|}_{(\mathfrak{g}_{\theta}^{*})^{\mathrm{reg}}}\cong C_{\mathcal{G}_{\theta}}\big{|}_{(\mathfrak{g}_{\theta}^{*})^{\mathrm{reg}}}.
Remark 6.11 (𝔾m\mathbb{G}_{m}-action and the moment map derivative).

The multiplicative group 𝔾m\mathbb{G}_{m} acts on 𝔤\mathfrak{g}^{*} by scalar multiplication, inducing an action on 𝔠\mathfrak{c}^{*}. This action preserves centralizers, so the action on C𝔤C_{\mathfrak{g}^{*}} is t(g,φ):=(g,tφ)t\cdot(g,\varphi):=(g,t\varphi). This action preserves the regular locus. The derivative of the moment map for the GG-action on TGT^{*}G, restricted to the centralizer, gives a morphism

dm:χLie(J𝔠)𝔤×𝔤.dm\colon\chi^{*}\operatorname{Lie}(J_{\mathfrak{c}^{*}})\longrightarrow\mathfrak{g}\times\mathfrak{g}^{*}.

This map is equivariant for the 𝔾m\mathbb{G}_{m}-action where 𝔾m\mathbb{G}_{m} acts trivially on 𝔤\mathfrak{g} and by scaling on 𝔤\mathfrak{g}^{*}. Identifying the cotangent bundle T𝔤𝔤(1)×𝔤T^{*}\mathfrak{g}^{*}\cong\mathfrak{g}(-1)\times\mathfrak{g}^{*}, we can view dmdm as a morphism

dm:χLie(J𝔠)(1)T𝔤.dm\colon\chi^{*}\operatorname{Lie}(J_{\mathfrak{c}^{*}})(-1)\longrightarrow T^{*}\mathfrak{g}^{*}. (6.6)

Its restriction to the regular locus (𝔤)reg(\mathfrak{g}^{*})^{\mathrm{reg}} (and hence to (𝔤θ)reg(\mathfrak{g}_{\theta}^{*})^{\mathrm{reg}}) is injective.

Remark 6.12 (Derivative of the Chevalley map).

The Chevalley map χ:𝔤𝔠\chi\colon\mathfrak{g}^{*}\to\mathfrak{c}^{*} is GG-invariant and 𝔾m\mathbb{G}_{m}-equivariant. Its derivative

dχ:T𝔤χT𝔠d\chi\colon T\mathfrak{g}^{*}\longrightarrow\chi^{*}T\mathfrak{c}^{*} (6.7)

is therefore also 𝔾m\mathbb{G}_{m}-equivariant. The restriction of dχd\chi to the regular locus is surjective, a fact that remains true upon further restriction to (𝔤θ)reg(\mathfrak{g}_{\theta}^{*})^{\mathrm{reg}}.

Lemma 6.13.

The canonical pairing on T𝔤×𝔤T𝔤T\mathfrak{g}^{*}\times_{\mathfrak{g}^{*}}T^{*}\mathfrak{g}^{*} induces a G×𝔾mG\times\mathbb{G}_{m}-equivariant perfect pairing on the regular locus

χLie(J)|(𝔤)reg(1)×(𝔤)regχT𝔠|(𝔤)reg(0),\chi^{*}\operatorname{Lie}(J)\big{|}_{(\mathfrak{g}^{*})^{\mathrm{reg}}}(-1)\times_{(\mathfrak{g}^{*})^{\mathrm{reg}}}\chi^{*}T\mathfrak{c}^{*}\big{|}_{(\mathfrak{g}^{*})^{\mathrm{reg}}}\longrightarrow\mathbb{C}(0),

which yields an isomorphism of vector bundles over 𝔠reg\mathfrak{c}^{*}_{\mathrm{reg}}:

Lie(J)(1)T𝔠.\operatorname{Lie}(J)^{*}(1)\cong T\mathfrak{c}^{*}.
Proof.

From Remarks 6.11 and 6.12, χLie(J)|(𝔤)reg(1)\chi^{*}\operatorname{Lie}(J)|_{(\mathfrak{g}^{*})^{\mathrm{reg}}}(-1) is a subbundle of T𝔤|(𝔤)regT^{*}\mathfrak{g}^{*}|_{(\mathfrak{g}^{*})^{\mathrm{reg}}} and χT𝔠|(𝔤)reg\chi^{*}T\mathfrak{c}^{*}|_{(\mathfrak{g}^{*})^{\mathrm{reg}}} is a quotient bundle of T𝔤|(𝔤)regT\mathfrak{g}^{*}|_{(\mathfrak{g}^{*})^{\mathrm{reg}}}. Both bundles have the same rank. Since χ\chi is constant on GG-orbits, the tangent space to the GG-orbit at a regular element φ\varphi, Vφ:=Im(𝔤ad()(φ)Tφ𝔤)V_{\varphi}:=\operatorname{Im}(\mathfrak{g}\xrightarrow{\mathrm{ad}^{*}(\cdot)(\varphi)}T_{\varphi}\mathfrak{g}^{*}), is contained in Ker(dχφ)\mathrm{Ker}(d\chi_{\varphi}). A dimension count shows that for regular φ\varphi, dimVφ=dim𝔤rank(𝔤)\dim V_{\varphi}=\dim\mathfrak{g}-\mathrm{rank}(\mathfrak{g}), which equals the dimension of Ker(dχφ)\mathrm{Ker}(d\chi_{\varphi}) since dχφd\chi_{\varphi} is surjective. Thus, Ker(dχφ)=Vφ\mathrm{Ker}(d\chi_{\varphi})=V_{\varphi}. By GG-invariance, the annihilator of the orbit tangent space under the canonical pairing is the centralizer Lie algebra, Vφ=𝔤φLie(Jφ)V_{\varphi}^{\perp}=\mathfrak{g}^{\varphi}\cong\mathrm{Lie}(J_{\varphi}). This establishes the perfect pairing. ∎

6.6. Globalization and Duality

We now globalize this local construction. Let LH(X,𝒢𝜽)\mathcal{M}_{LH}(X,\mathcal{G}_{\boldsymbol{\theta}}) be the moduli space of leveled logahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-Higgs torsors (,φ,η)(\mathcal{E},\varphi,\eta). This space is an open subset of the cotangent bundle T𝒰(X,𝒢𝜽)T^{*}\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}}) and is therefore a symplectic manifold with a canonical symplectic form, which we denote by ωLH\omega_{LH}.

Let :LH(X,𝒢𝜽)H(X,𝒢𝜽)\ell:\mathcal{M}_{LH}(X,\mathcal{G}_{\boldsymbol{\theta}})\to\mathcal{M}_{H}(X,\mathcal{G}_{\boldsymbol{\theta}}) be the forgetful map, which is a Poisson map as we have seen. The Hitchin fibration on the leveled space is the composition

H~𝜽:=h𝜽:LH(X,𝒢𝜽)𝒜𝜽.{\widetilde{H}_{\boldsymbol{\theta}}}:=h_{\boldsymbol{\theta}}\circ\ell:\mathcal{M}_{LH}(X,\mathcal{G}_{\boldsymbol{\theta}})\to\mathcal{A}_{\boldsymbol{\theta}}.

The group scheme JJ on 𝔠\mathfrak{c}^{*} pulls back along the Hitchin fibration h𝜽h_{\boldsymbol{\theta}} to a group scheme J𝒜𝜽J_{\mathcal{A}_{\boldsymbol{\theta}}} over the Hitchin base 𝒜𝜽\mathcal{A}_{\boldsymbol{\theta}}. The moduli space of J𝒜𝜽J_{\mathcal{A}_{\boldsymbol{\theta}}}-torsors on XX forms a group scheme P𝒜𝜽P_{\mathcal{A}_{\boldsymbol{\theta}}} over 𝒜𝜽\mathcal{A}_{\boldsymbol{\theta}}, whose fiber PaP_{a} over a point a𝒜𝜽a\in\mathcal{A}_{\boldsymbol{\theta}} is the generalized Prym variety associated to the cameral cover determined by aa. This induces an action on the Hitchin fibers, which lifts to the leveled space:

act𝜽:P𝒜𝜽×𝒜𝜽LH(X,𝒢𝜽)LH(X,𝒢𝜽).\mathrm{act}_{\boldsymbol{\theta}}\colon P_{\mathcal{A}_{\boldsymbol{\theta}}}\times_{\mathcal{A}_{\boldsymbol{\theta}}}\mathcal{M}_{LH}(X,{\mathcal{G}_{\boldsymbol{\theta}}})\longrightarrow\mathcal{M}_{LH}(X,{\mathcal{G}_{\boldsymbol{\theta}}}).

The following statement can be now obtained analogously to [13, Proposition A.12]:

Proposition 6.14.

There exists a canonical isomorphism of vector bundles over 𝒜𝛉\mathcal{A}_{\boldsymbol{\theta}},

Lie(P𝒜𝜽/𝒜𝜽)T𝒜𝜽,\mathrm{Lie}\Bigl{(}P_{\mathcal{A}_{\boldsymbol{\theta}}}/\mathcal{A}_{\boldsymbol{\theta}}\Bigr{)}\;\cong\;T^{*}\mathcal{A}_{\boldsymbol{\theta}},

such that the differential of the lifted action,

dact𝜽:H~𝜽Lie(P𝒜𝜽/𝒜𝜽)TLH(X,𝒢𝜽),d\mathrm{act}_{\boldsymbol{\theta}}\colon{\widetilde{H}_{\boldsymbol{\theta}}}^{*}\mathrm{Lie}\Bigl{(}P_{\mathcal{A}_{\boldsymbol{\theta}}}/\mathcal{A}_{\boldsymbol{\theta}}\Bigr{)}\longrightarrow T\mathcal{M}_{LH}(X,{\mathcal{G}_{\boldsymbol{\theta}}}),

and the differential of the lifted Hitchin fibration,

dH~𝜽:TLH(X,𝒢𝜽)H~𝜽T𝒜𝜽d{\widetilde{H}_{\boldsymbol{\theta}}}\colon T\mathcal{M}_{LH}(X,{\mathcal{G}_{\boldsymbol{\theta}}})\longrightarrow{\widetilde{H}_{\boldsymbol{\theta}}}^{*}T\mathcal{A}_{\boldsymbol{\theta}}

are dual to each other with respect to the canonical symplectic form ωLH\omega_{LH} on the smooth locus of LH(X,𝒢𝛉)\mathcal{M}_{LH}(X,{\mathcal{G}_{\boldsymbol{\theta}}}).

Proof.

Let (,φ,η)(\mathcal{E},\varphi,\eta) be a leveled logahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-Higgs torsor mapping to a𝒜𝜽a\in\mathcal{A}_{\boldsymbol{\theta}}. The tangent space to the moduli space at this point, T(,φ,η)LHT_{(\mathcal{E},\varphi,\eta)}\mathcal{M}_{LH}, is given by the hypercohomology group 1(X,𝒦,φ)\mathbb{H}^{1}(X,\mathcal{K}_{\mathcal{E},\varphi}), where 𝒦,φ\mathcal{K}_{\mathcal{E},\varphi} is the two-term complex [(𝔤+)ad(φ)(𝔤)KX(D)][\mathcal{E}(\mathfrak{g}^{+})\xrightarrow{\mathrm{ad}(\varphi)}\mathcal{E}(\mathfrak{g})\otimes K_{X}(D)] that governs deformations preserving the level structure.

The differential of the Hitchin map, dH~𝜽d{\widetilde{H}_{\boldsymbol{\theta}}}, is induced on hypercohomology by the morphism of complexes from 𝒦,φ\mathcal{K}_{\mathcal{E},\varphi} to χ(φ)T𝔠KX(D)\chi(\varphi)^{*}T\mathfrak{c}_{K_{X}(D)} given by the derivative of the Chevalley map, dχd\chi. The differential of the action, dact𝜽d\mathrm{act}_{\boldsymbol{\theta}}, is induced by a map from the complex aLie(J𝒜𝜽)[1]a^{*}\mathrm{Lie}(J_{\mathcal{A}_{\boldsymbol{\theta}}})[-1] into the complex 𝒦,φ\mathcal{K}_{\mathcal{E},\varphi}.

The local duality from Lemma 6.13 globalizes. The perfect pairing between χLie(J)\chi^{*}\operatorname{Lie}(J) and χT𝔠\chi^{*}T\mathfrak{c}^{*} induces, via Serre duality on hypercohomology, a perfect pairing:

1(X,aLie(J𝒜𝜽))×1(X,χ(φ)T𝔠KX(D)).\mathbb{H}^{1}\Bigl{(}X,a^{*}\mathrm{Lie}(J_{\mathcal{A}_{\boldsymbol{\theta}}})\Bigr{)}\times\mathbb{H}^{1}\Bigl{(}X,\chi(\varphi)^{*}T\mathfrak{c}_{K_{X}(D)}\Bigr{)}\longrightarrow\mathbb{C}.

Standard deformation theory identifies Lie(Pa)1(X,aLie(J𝒜𝜽))\mathrm{Lie}(P_{a})\cong\mathbb{H}^{1}(X,a^{*}\mathrm{Lie}(J_{\mathcal{A}_{\boldsymbol{\theta}}})) and

Ta𝒜𝜽1(X,χ(φ)T𝔠KX(D)).T_{a}^{*}\mathcal{A}_{\boldsymbol{\theta}}\cong\mathbb{H}^{1}(X,\chi(\varphi)^{*}T\mathfrak{c}_{K_{X}(D)}).

The duality of the hypercohomology groups translates directly to the asserted duality between the morphisms dact𝜽d\mathrm{act}_{\boldsymbol{\theta}} and dH~𝜽d{\widetilde{H}_{\boldsymbol{\theta}}} with respect to the symplectic form ωLH\omega_{LH}. ∎

We now use the duality principle established above in order to prove the complete integrability of the parahoric Hitchin system. The core of the argument is to first show that the generic fibers of the Hitchin fibration on the symplectic manifold LH(X,𝒢𝜽)\mathcal{M}_{LH}(X,\mathcal{G}_{\boldsymbol{\theta}}) are Lagrangian, and then to descend this property to the Poisson manifold H(X,𝒢𝜽)\mathcal{M}_{H}(X,\mathcal{G}_{\boldsymbol{\theta}}) via the forgetful map.

Theorem 6.15.

The generic fibers of the Hitchin fibration H~𝛉:LH(X,𝒢𝛉)𝒜𝛉{\widetilde{H}_{\boldsymbol{\theta}}}\colon\mathcal{M}_{LH}(X,\mathcal{G}_{\boldsymbol{\theta}})\to\mathcal{A}_{\boldsymbol{\theta}} are Lagrangian subvarieties of the symplectic manifold (LH(X,𝒢𝛉),ωLH)(\mathcal{M}_{LH}(X,\mathcal{G}_{\boldsymbol{\theta}}),\omega_{LH}).

Proof.

Let a𝒜a\in\mathcal{A}^{\prime} be a generic point in the Hitchin base. The fiber of the Hitchin fibration over this point is Fa:=H~𝜽1(a)F^{\prime}_{a}:={\widetilde{H}_{\boldsymbol{\theta}}}^{-1}(a). For any point pFap\in F^{\prime}_{a}, the tangent space to the fiber is given by the kernel of the differential of the Hitchin map, TpFa=Ker(dH~𝜽)pT_{p}F^{\prime}_{a}=\text{Ker}(d{\widetilde{H}_{\boldsymbol{\theta}}})_{p}.

The fiber FaF^{\prime}_{a} is acted upon by the generalized Prym variety PaP_{a}. The tangent space to the orbit of this action at pp is given by the image of the differential of the action map, Im(dact𝜽)p\mathrm{Im}(d\mathrm{act}_{\boldsymbol{\theta}})_{p}. As established in Proposition 6.14, these two subspaces of the tangent space TpLH(X,𝒢𝜽)T_{p}\mathcal{M}_{LH}(X,{\mathcal{G}_{\boldsymbol{\theta}}}) are symplectic orthogonals with respect to the form ωLH\omega_{LH}:

Ker(dH~𝜽)=(Im(dact𝜽)).\text{Ker}(d{\widetilde{H}_{\boldsymbol{\theta}}})=(\mathrm{Im}(d\mathrm{act}_{\boldsymbol{\theta}}))^{\perp}.

The action of the abelian group scheme P𝒜𝜽P_{\mathcal{A}_{\boldsymbol{\theta}}} on the fibers generates isotropic submanifolds. This is a standard result stemming from the fact that the action linearizes on the Jacobian of the cameral cover, and the symplectic form, when pulled back to the abelian Prym variety, is translation-invariant and must therefore be zero. Consequently, the tangent space to the Prym orbits, Im(dact𝜽)\mathrm{Im}(d\mathrm{act}_{\boldsymbol{\theta}}), is an isotropic subspace of TLH(X,𝒢𝜽)T\mathcal{M}_{LH}(X,\mathcal{G}_{\boldsymbol{\theta}}).

A fundamental result in symplectic linear algebra states that the symplectic orthogonal of an isotropic subspace is a coisotropic subspace. Therefore, the tangent space to the Hitchin fiber, Ker(dH~𝜽)\text{Ker}(d{\widetilde{H}_{\boldsymbol{\theta}}}), is coisotropic.

In conclusion, the subspaces FaF_{a}^{\prime} are coisotropic and isotropic subspaces of TLH(X,𝒢𝜽)T\mathcal{M}_{LH}(X,\mathcal{G}_{\boldsymbol{\theta}}), and therefore Lagrangian. ∎

Theorem 6.16.

The generic fibers of the parahoric Hitchin fibration h𝛉:H(X,𝒢𝛉)𝒜𝛉h_{\boldsymbol{\theta}}\colon\mathcal{M}_{H}(X,\mathcal{G}_{\boldsymbol{\theta}})\to\mathcal{A}_{\boldsymbol{\theta}} are Lagrangian subvarieties with respect to the symplectic leaves of the Poisson moduli space H(X,𝒢𝛉)\mathcal{M}_{H}(X,\mathcal{G}_{\boldsymbol{\theta}}).

Proof.

This theorem is a direct consequence of Theorem 6.15 and the fact that the forgetful map :LH(X,𝒢𝜽)H(X,𝒢𝜽)\ell:\mathcal{M}_{LH}(X,\mathcal{G}_{\boldsymbol{\theta}})\to\mathcal{M}_{H}(X,\mathcal{G}_{\boldsymbol{\theta}}) is a Poisson map, a result established in Theorem 5.20. ∎

Collecting Theorems 5.20, 6.10 and 6.16, we now conclude to the following:

Theorem 6.17.

Let XX be a smooth complex algebraic curve and DD be a reduced effective divisor on XX. Let GG be a connected complex reductive group. The moduli space H(X,𝒢𝛉)\mathcal{M}_{H}(X,\mathcal{G}_{\boldsymbol{\theta}}) of logahoric 𝒢𝛉\mathcal{G}_{\boldsymbol{\theta}}-Higgs torsors over XX is Poisson and is fibered via a map h𝛉:H(X,𝒢𝛉)𝒜𝛉h_{\boldsymbol{\theta}}:\mathcal{M}_{H}(X,\mathcal{G}_{\boldsymbol{\theta}})\to\mathcal{A}_{\boldsymbol{\theta}} by abelian torsors. Moreover, h𝛉:H(X,𝒢𝛉)𝒜𝛉h_{\boldsymbol{\theta}}:\mathcal{M}_{H}(X,\mathcal{G}_{\boldsymbol{\theta}})\to\mathcal{A}_{\boldsymbol{\theta}} is an algebraically completely integrable Hamiltonian system in the sense of Definition 5.10.

We will call the algebraically completely integrable Hamiltonian system of Theorem 6.17, the logahoric Hitchin integrable system.

7. Symplectic leaf foliation

We now study the symplectic leaves in the foliation of the Poisson moduli space H(X,𝒢𝜽)\mathcal{M}_{H}(X,\mathcal{G}_{\boldsymbol{\theta}}).

Lemma 7.1.

Let p:𝔤p\colon\mathfrak{g}\to\mathbb{C} be a homogeneous GG-invariant polynomial of positive degree. Then p(x)=0p(x)=0, for every nilpotent element x𝔤x\in\mathfrak{g}.

Proof.

Fix a nilpotent element x𝔤x\in\mathfrak{g}. By the Jacobson–Morozov Theorem, there exists a 1-parameter cocharacter λ:×G\lambda:\mathbb{C}^{\times}\!\to G such that Ad(λ(t))x=t 2x\operatorname{Ad}(\lambda(t))\,x\;=\;t^{\,2}\,x, for all t×t\in\mathbb{C}^{\times}. Because pp is GG-invariant, p(Ad(λ(t))x)=p(x)p\bigl{(}\operatorname{Ad}(\lambda(t))x\bigr{)}=p(x). Thus, p(x)=p(t 2x)=t2degpp(x)p(x)\;=\;p\bigl{(}t^{\,2}x\bigr{)}\;=\;t^{2\deg p}\,p(x) (homogeneity).

Now, pick t0t\neq 0. If p(x)0p(x)\neq 0 we could divide by it to get 1=t2degp1=t^{2\deg p}, which is impossible for arbitrary tt. Hence p(x)=0p(x)=0. ∎

From Lemma 7.1 we get for the Chevalley morphism that χ(X)=0\chi(X)=0, for all X𝔤θ+X\in\mathfrak{g}_{\theta}^{+}. We set the following:

Definition 7.2 (Strongly logahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-Higgs torsors and 𝒜𝜽+\mathcal{A}_{\boldsymbol{\theta}}^{+}).

A logarithmic Higgs field φH0(X,(𝔤)K(D))\varphi\in H^{0}(X,\mathcal{E}(\mathfrak{g})\otimes K(D)) is called strongly logarithmic if, for each xiDx_{i}\in D, its residue Resxiφ=0\mathrm{Res}_{x_{i}}\varphi=0. A strongly logahoric 𝒢𝛉\mathcal{G}_{\boldsymbol{\theta}}-Higgs torsor over XX is a parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-Higgs torsor over XX with a strongly logarithmic Higgs field. Since non-constant invariant polynomials pjp_{j} vanish on nilpotent elements (Lemma 7.1), then pj(Resxiφ)=0p_{j}(\mathrm{Res}_{x_{i}}\varphi)=0, implying that pj(φ)p_{j}(\varphi) vanishes along DD. Now, set

𝜽+:=j=1lH0(X,(KX(D))mj+1𝒪X(D)),\mathcal{B}_{\boldsymbol{\theta}}^{+}:=\bigoplus_{j=1}^{l}H^{0}(X,(K_{X}(D))^{m_{j}+1}\otimes\mathcal{O}_{X}(-D)),

which is a vector subspace of 𝜽\mathcal{B}_{\boldsymbol{\theta}}. The subvariety 𝒜𝜽+\mathcal{A}_{\boldsymbol{\theta}}^{+} is then the image under h𝜽h_{\boldsymbol{\theta}} of the strongly logahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-Higgs torsors, consisting of tuples (s1,,sl)𝒜𝜽(s_{1},\dots,s_{l})\in\mathcal{A}_{\boldsymbol{\theta}} such that each sjs_{j} vanishes along DD. In other words, 𝒜𝜽+=𝒜𝜽𝜽+\mathcal{A}_{\boldsymbol{\theta}}^{+}=\mathcal{A}_{\boldsymbol{\theta}}\cap\mathcal{B}_{\boldsymbol{\theta}}^{+}.

Definition 7.3 (Local Residue Data).

For each xiDx_{i}\in D, let GθiG_{\theta_{i}} be the parahoric subgroup of GG at xix_{i}, with Levi factor LθiL_{\theta_{i}} and Lie algebra 𝔩θi\mathfrak{l}_{\theta_{i}}. The residue Resxiφ\mathrm{Res}_{x_{i}}\varphi lies in 𝔩^θi𝔩θi\hat{\mathfrak{l}}_{\theta_{i}}\cong\mathfrak{l}_{\theta_{i}}. The effective structure group for residues is GD=(i=1sLθi)/Z(G)G_{D}=\left(\prod_{i=1}^{s}L_{\theta_{i}}\right)/Z(G), with Lie algebra 𝔤D=i=1s𝔩θi\mathfrak{g}_{D}=\bigoplus_{i=1}^{s}\mathfrak{l}_{\theta_{i}}. The space of local invariant data is the categorical quotient 𝔤D//GD=Spec([𝔤D]GD)\mathfrak{g}_{D}^{*}//G_{D}=\mathrm{Spec}(\mathbb{C}[\mathfrak{g}_{D}^{*}]^{G_{D}}), specified by LθiL_{\theta_{i}}-invariant polynomial values on Resxiφ𝔩θi\mathrm{Res}_{x_{i}}\varphi\in\mathfrak{l}_{\theta_{i}}, for each xiDx_{i}\in D.

For each basic Chevalley invariant pjp_{j}, the short exact sequence

0(KX(D))mj+1(D)(KX(D))mj+1resD𝒪D00\to(K_{X}(D))^{m_{j}+1}(-D)\to(K_{X}(D))^{m_{j}+1}\xrightarrow{\operatorname{res}_{D}}\mathcal{O}_{D}\to 0

gives a surjective linear map

resD:B𝜽IH0(D,𝒪D), with Ker(resD)=B𝜽+.\operatorname{res}_{D}\colon B_{\boldsymbol{\theta}}\;\twoheadrightarrow\;I\otimes H^{0}(D,\mathcal{O}_{D}),\quad\text{ with }\quad\text{Ker}(\operatorname{res}_{D})=B_{\boldsymbol{\theta}}^{+}. (7.1)

Here I=j=1Imj+1I=\bigoplus_{j=1}^{\ell}I_{m_{j}+1} is the span of the basic invariants. For each marked point xiDx_{i}\in D, let LθiL_{\theta_{i}} be the Levi factor of the parahoric subgroup GθiG_{\theta_{i}} with Lie algebra 𝔩θi\mathfrak{l}_{\theta_{i}}. The Jacobson–Morozov Theorem plus Lemma 7.1 imply that [𝔤θi]Gθi=[𝔩θi]Lθi\mathbb{C}[\mathfrak{g}_{\theta_{i}}]^{G_{\theta_{i}}}=\mathbb{C}[\mathfrak{l}_{\theta_{i}}]^{L_{\theta_{i}}}, hence the canonical isomorphism

𝔤θi//Gθi𝔩θi//Lθi.\mathfrak{g}_{\theta_{i}}//G_{\theta_{i}}\;\cong\;\mathfrak{l}_{\theta_{i}}//L_{\theta_{i}}.

Set GD:=(i=1sLθi)/Z(G),𝔤D:=i=1s𝔩θi.G_{D}:=\bigl{(}\prod_{i=1}^{s}L_{\theta_{i}}\bigr{)}/Z(G),\;\mathfrak{g}_{D}:=\bigoplus_{i=1}^{s}\mathfrak{l}_{\theta_{i}}. Then, Chevalley’s theorem gives

𝔤D//GDIH0(D,𝒪D)\mathfrak{g}_{D}^{*}//G_{D}\;\cong\;I\otimes H^{0}(D,\mathcal{O}_{D}) (7.2)

as an affine space; now define

𝒜𝜽/𝒜𝜽+:=the image of 𝒜𝜽 inside 𝜽/𝜽+.\mathcal{A}_{\boldsymbol{\theta}}/\mathcal{A}_{\boldsymbol{\theta}}^{+}\;:=\;\mbox{the image of }\mathcal{A}_{\boldsymbol{\theta}}\mbox{ inside }\mathcal{B}_{\boldsymbol{\theta}}/\mathcal{B}_{\boldsymbol{\theta}}^{+}.
Remark 7.4.

Note that 𝒜𝜽/𝒜𝜽+\mathcal{A}_{\boldsymbol{\theta}}/\mathcal{A}_{\boldsymbol{\theta}}^{+} above is not a well-defined quotient as 𝒜𝜽+\mathcal{A}_{\boldsymbol{\theta}}^{+} is not a vector space. However, we adopt this notation as to align with the vector space BL/B0B_{L}/B_{0} appearing in [26, Section 8.3].

Theorem 7.5.

There exists an algebraic morphism r¯:𝒜𝛉/𝒜𝛉+𝔤D//GD\overline{r}:\mathcal{A}_{\boldsymbol{\theta}}/\mathcal{A}_{\boldsymbol{\theta}}^{+}\to\mathfrak{g}_{D}^{*}//G_{D} which is an isomorphism of affine varieties:

𝒜𝜽/𝒜𝜽+𝔤D//GD.\mathcal{A}_{\boldsymbol{\theta}}/\mathcal{A}_{\boldsymbol{\theta}}^{+}\cong\mathfrak{g}_{D}^{*}//G_{D}.

As so this fits into a diagram

𝒜𝜽{\mathcal{A}_{\boldsymbol{\theta}}}𝒜𝜽/𝒜𝜽+{\mathcal{A}_{\boldsymbol{\theta}}/\mathcal{A}_{\boldsymbol{\theta}}^{+}}T𝒰(X,𝒢𝜽){T^{*}\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}})}𝔤D{\mathfrak{g}_{D}^{*}}𝔤D//GD{\mathfrak{g}_{D}^{*}//G_{D}}q\scriptstyle{q}\scriptstyle{\cong}μ\scriptstyle{\mu}H~𝜽\scriptstyle{\widetilde{H}_{\boldsymbol{\theta}}}c.q.\scriptstyle{c.q.} (7.3)

for the canonical quotient 𝔤D𝔤D//GD\mathfrak{g}_{D}^{*}\to\mathfrak{g}_{D}^{*}//G_{D} and the quotient map 𝒜𝛉𝒜𝛉/𝒜𝛉+\mathcal{A}_{\boldsymbol{\theta}}\to\mathcal{A}_{\boldsymbol{\theta}}/\mathcal{A}_{\boldsymbol{\theta}}^{+}.

Proof.

For a pair (,φ)(\mathcal{E},\varphi), let ξi=Resxiφ𝔤θi\xi_{i}=\operatorname{Res}_{x_{i}}\varphi\in\mathfrak{g}_{\theta_{i}}. By the paragraph above, the tuple of invariant values (p1(ξi),,p(ξi))(p_{1}(\xi_{i}),\dots,p_{\ell}(\xi_{i})) depends only on the Levi component ξi,s𝔩θi\xi_{i,s}\in\mathfrak{l}_{\theta_{i}} and defines a point of 𝔩θi//Lθi𝔤θi//Gθi\mathfrak{l}_{\theta_{i}}//L_{\theta_{i}}\cong\mathfrak{g}_{\theta_{i}}//G_{\theta_{i}}. Collecting over ii yields a map r:𝒜𝜽𝔤D//GD.r:\mathcal{A}_{\boldsymbol{\theta}}\to\mathfrak{g}_{D}^{*}//G_{D}. If φ\varphi is strongly logarithmic then each residue is nilpotent, so r(φ)=0r(\varphi)=0; hence rr factors through r¯:𝒜𝜽/𝒜𝜽+𝔤D//GD\bar{r}:\mathcal{A}_{\boldsymbol{\theta}}/\mathcal{A}_{\boldsymbol{\theta}}^{+}\to\mathfrak{g}_{D}^{*}//G_{D}.

Given any (ηi)𝔤D(\eta_{i})\in\mathfrak{g}_{D} we may choose local representatives with semisimple part ηi\eta_{i} and glue a Higgs field on XX whose residues are these ηi\eta_{i}, this is achievable since we assume our torsor to be generically split. Evaluating invariants shows every point of 𝔤D//GD\mathfrak{g}_{D}^{*}//G_{D} is in the image of r¯\bar{r}.

Now note that the induced map resD¯:𝜽/𝜽+𝔤D//GD\overline{res_{D}}:\mathcal{B}_{\boldsymbol{\theta}}/\mathcal{B}_{\boldsymbol{\theta}}^{+}\to\mathfrak{g}_{D}^{*}//G_{D} is injective because the kernel of 𝜽\mathcal{B}_{\boldsymbol{\theta}} is inside 𝜽+\mathcal{B}_{\boldsymbol{\theta}}^{+}. As so the image of 𝒜𝜽\mathcal{A}_{\boldsymbol{\theta}} inside 𝜽/𝜽+\mathcal{B}_{\boldsymbol{\theta}}/\mathcal{B}_{\boldsymbol{\theta}}^{+} maps to 𝔤D//GD\mathfrak{g}_{D}^{*}//G_{D} injectively from the diagram below equation (7.1)

𝒜𝜽𝜽𝜽/𝜽+resD¯𝔤D//GD.\mathcal{A}_{\boldsymbol{\theta}}\to\mathcal{B}_{\boldsymbol{\theta}}\to\mathcal{B}_{\boldsymbol{\theta}}/\mathcal{B}^{+}_{\boldsymbol{\theta}}\overset{\overline{res_{D}}}{\to}\mathfrak{g}_{D}^{*}//G_{D}.

In the diagram (7.3), note that the moment map μ\mu followed by the Chevalley map equals the evaluation of pj(φ)p_{j}(\varphi) at xix_{i}. Therefore, the map r¯\bar{r} is an algebraic isomorphism and the diagram (7.3) commutes. ∎

Corollary 7.6.

The foliation of the smooth locus H(X,𝒢𝛉)sm\mathcal{M}_{H}(X,\mathcal{G}_{\boldsymbol{\theta}})^{\mathrm{sm}} by its symplectic leaves refines the foliation by fibers of the map

qH~𝜽:H(X,𝒢𝜽)sm𝒜𝜽/𝒜𝜽+.q\circ\widetilde{H}_{\boldsymbol{\theta}}\;:\;\mathcal{M}_{H}(X,\mathcal{G}_{\boldsymbol{\theta}})^{\mathrm{sm}}\;\longrightarrow\;\mathcal{A}_{\boldsymbol{\theta}}/\mathcal{A}_{\boldsymbol{\theta}}^{+}.

Moreover, each fiber of qH~𝛉q\circ\widetilde{H}_{\boldsymbol{\theta}} contains a unique symplectic leaf of maximal dimension.

Proof.

On the open locus R|T𝒰(X,𝒢𝜽)|R\subset|T^{*}\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}})| we have

qH~𝜽=c.q.μ.q\circ\widetilde{H}_{\boldsymbol{\theta}}\;=\;\mathrm{c.q.}\,\circ\mu.

Since symplectic leaves in T𝒰(X,𝒢𝜽)T^{*}\mathcal{U}(X,\mathcal{G}_{\boldsymbol{\theta}}) are exactly the connected components of the pre-images under μ\mu of GDG_{D}-coadjoint orbits in 𝔤D\mathfrak{g}_{D}^{*}, Theorem 7.5 implies that each such orbit sits inside exactly one leaf in a given fiber of qH~𝜽q\circ\widetilde{H}_{\boldsymbol{\theta}}. Hence the fibers decompose into symplectic leaves, and the unique open (co)adjoint orbit in each fiber yields the unique leaf of maximal dimension. ∎

8. Examples of Logahoric Hitchin Integrable Systems

Many classical, as well as recently discovered, integrable systems can be realized as symplectic leaves of the integrable systems introduced in this paper, for suitable choices of the group GG, the curve XX, the divisor DD, and the parahoric data 𝜽\boldsymbol{\theta}. This section is devoted to demonstrating certain examples in the case when the base curve is the projective line, X=1X=\mathbb{P}^{1} or an elliptic curve X=ΣX=\Sigma. We will show that in the first case, our general framework generalizes the integrable system of Beauville [4], and recovers in its simplest form, the classical Gaudin model [17]. Moreover, for base curve an elliptic curve X=ΣX=\Sigma, we recover the space of Λ\Lambda-periodic KP elliptic solitons and the elliptic Calogero–Moser system.

8.1. Logarithmic Hitchin System on 1\mathbb{P}^{1}

For X=1X=\mathbb{P}^{1}, the geometry simplifies considerably. A principal GG-bundle on 1\mathbb{P}^{1} is semistable if and only if it is trivial. Since the stability of a parahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-torsor requires the underlying principal GG-bundle to be semistable, we can fix the underlying torsor \mathcal{E} to be the trivial one, i.e., =1×G\mathcal{E}=\mathbb{P}^{1}\times G.

Let D={x1,,xs}1D=\{x_{1},\dots,x_{s}\}\subset\mathbb{P}^{1} be a reduced effective divisor of degree s1s\geq 1. The cotangent space to the moduli space of leveled torsors 𝒰(1,𝒢𝜽)\mathcal{U}(\mathbb{P}^{1},\mathcal{G}_{\boldsymbol{\theta}}) is the space of logarithmic Higgs fields:

T[(,η)]𝒰(1,𝒢𝜽)H0(1,𝔤K1(D)).T^{*}_{[(\mathcal{E},\eta)]}\mathcal{U}(\mathbb{P}^{1},\mathcal{G}_{\boldsymbol{\theta}})\cong H^{0}(\mathbb{P}^{1},\mathfrak{g}\otimes K_{\mathbb{P}^{1}}(D)).

Let us choose a coordinate zz on 1\mathbb{P}^{1} such that no points in DD are at z=z=\infty. The canonical bundle is K1=𝒪1(2)K_{\mathbb{P}^{1}}=\mathcal{O}_{\mathbb{P}^{1}}(-2), so K1(D)=𝒪1(s2)K_{\mathbb{P}^{1}}(D)=\mathcal{O}_{\mathbb{P}^{1}}(s-2). A section ΦH0(1,𝔤𝒪1(s2))\Phi\in H^{0}(\mathbb{P}^{1},\mathfrak{g}\otimes\mathcal{O}_{\mathbb{P}^{1}}(s-2)) is a 𝔤\mathfrak{g}-valued polynomial in zz of degree at most s2s-2. Let 𝒫s2(𝔤)\mathcal{P}_{s-2}(\mathfrak{g}) denote the vector space of such polynomials. The phase space of the integrable system is a Poisson quotient of this space:

MD𝒫s2(𝔤)//G.M_{D}\cong\mathcal{P}_{s-2}(\mathfrak{g})//G.

The Hitchin fibration h𝜽:MD𝒜𝜽h_{\boldsymbol{\theta}}:M_{D}\to\mathcal{A}_{\boldsymbol{\theta}} is given by taking the characteristic coefficients of a polynomial matrix A(z)𝒫s2(𝔤)A(z)\in\mathcal{P}_{s-2}(\mathfrak{g}).

8.2. Gaudin model

We now demonstrate explicitly how this framework recovers the well-known classical Gaudin model also known as the Garnier integrable system [17]. The key is to make a specific choice for the parahoric data 𝜽\boldsymbol{\theta} that corresponds to the simplest pole structure. We choose the parahoric subgroup 𝒢θj\mathcal{G}_{\theta_{j}} at each point xjDx_{j}\in D to be an Iwahori subgroup.

With this choice, a logarithmic Higgs field ΦH0(1,𝔤K1(D))\Phi\in H^{0}(\mathbb{P}^{1},\mathfrak{g}\otimes K_{\mathbb{P}^{1}}(D)) is realized as a 𝔤\mathfrak{g}-valued meromorphic 1-form on 1\mathbb{P}^{1} that is holomorphic away from DD and has at worst simple (first-order) poles at each xjx_{j}. Such a 1-form can be written in terms of its residues. Let Xj=Resz=xjΦ𝔤X_{j}=\operatorname{Res}_{z=x_{j}}\Phi\in\mathfrak{g}. The 1-form can then be written in the coordinate zz as:

Φ(z)=(j=1sXjzxj)dz.\Phi(z)=\left(\sum_{j=1}^{s}\frac{X_{j}}{z-x_{j}}\right)dz. (8.1)

The coefficient of dzdz is immediately recognizable as the Gaudin Lax operator:

L(z)=j=1sXjzxj.L(z)=\sum_{j=1}^{s}\frac{X_{j}}{z-x_{j}}. (8.2)

The residues XjX_{j} of the Higgs field are identified with the dynamical spin variables of the Gaudin model. The condition that Φ(z)\Phi(z) is a regular 1-form at z=z=\infty implies the residue sum rule jXj=0\sum_{j}X_{j}=0, which defines the total momentum constraint of the Gaudin model.

The commuting Hamiltonians arise from the Hitchin map. Taking the simplest quadratic invariant (using a non-degenerate bilinear form ,\langle\cdot,\cdot\rangle on 𝔤\mathfrak{g}), we construct the generating function from the Lax operator:

H(z):=12L(z),L(z)=12j,k=1sXj,Xk(zxj)(zxk).H(z):=\frac{1}{2}\langle L(z),L(z)\rangle=\frac{1}{2}\sum_{j,k=1}^{s}\frac{\langle X_{j},X_{k}\rangle}{(z-x_{j})(z-x_{k})}. (8.3)

This is precisely the generating function for the Gaudin Hamiltonians. The general theorem on complete integrability for the logahoric Hitchin integrable system thus provides a deep geometric proof for the integrability of the classical Gaudin model.

The two descriptions of the phase space—as polynomials A(z)𝒫s2(𝔤)A(z)\in\mathcal{P}_{s-2}(\mathfrak{g}) and as rational functions L(z)L(z)—are algebraically equivalent. The isomorphism is given by clearing denominators.

Let P(z)=k=1s(zxk)P(z)=\prod_{k=1}^{s}(z-x_{k}) be the scalar polynomial whose roots are the poles of the Lax operator. Define the 𝔤\mathfrak{g}-valued polynomial A(z)A(z) by:

A(z):=P(z)L(z)=(k=1s(zxk))(j=1sXjzxj)=j=1sXjkj(zxk).A(z):=P(z)L(z)=\left(\prod_{k=1}^{s}(z-x_{k})\right)\left(\sum_{j=1}^{s}\frac{X_{j}}{z-x_{j}}\right)=\sum_{j=1}^{s}X_{j}\prod_{k\neq j}(z-x_{k}). (8.4)

Each term in the sum is a polynomial of degree s1s-1. However, due to the residue sum rule j=1sXj=0\sum_{j=1}^{s}X_{j}=0, the coefficient of the zs1z^{s-1} term vanishes:

coeff(zs1) in A(z)=j=1sXj=0.\text{coeff}(z^{s-1})\text{ in }A(z)=\sum_{j=1}^{s}X_{j}=0.

Therefore, the degree of the polynomial A(z)A(z) is at most s2s-2, and it lies in 𝒫s2(𝔤)\mathcal{P}_{s-2}(\mathfrak{g}). This establishes that the space of polynomials A(z)A(z) in the explicit realization is canonically isomorphic to the space of Gaudin Lax operators L(z)L(z) with the total momentum constraint. The Hitchin Hamiltonians can be generated from the invariants of either A(z)A(z) or L(z)L(z), as they are algebraically equivalent.

8.3. KP elliptic solitons

We now specialize to the case when the base curve XX is an elliptic curve Σ\Sigma. The triviality of the canonical bundle, KΣ𝒪ΣK_{\Sigma}\cong\mathcal{O}_{\Sigma}, simplifies the setup and reveals deep connections to well-known algebraically completely integrable Hamiltonian systems. A logarithmic Higgs field φH0(X,(𝔤)KX(D))\varphi\in H^{0}(X,\mathcal{E}(\mathfrak{g})\otimes K_{X}(D)) becomes a section of (𝔤)𝒪Σ(D)\mathcal{E}(\mathfrak{g})\otimes\mathcal{O}_{\Sigma}(D), i.e., a meromorphic section of the adjoint bundle with poles prescribed by the divisor DD. We will now show how the framework of the logahoric Hitchin integrable system recovers both the KP hierarchy and the appropriate symplectic leaf.

We first demonstrate that the space of elliptic solutions to the Kadomtsev–Petviashvili (KP) equation arises as a specific symplectic leaf of the moduli space of logahoric 𝒢𝜽\mathcal{G}_{\boldsymbol{\theta}}-Higgs torsors, following the treatment of Markman [26] and Treibich–Verdier [33].

Let the structure group be G=SL(r,)G=\mathrm{SL}(r,\mathbb{C}). We consider logahoric Higgs torsors of rank rr and degree 0. The underlying vector bundles are semistable, and the moduli space of the torsors themselves, 𝒰Σ(r,0)\mathcal{U}_{\Sigma}(r,0), is isomorphic to the symmetric product SymrΣ\mathrm{Sym}^{r}\Sigma. Let the divisor be a single point D={q}D=\{q\} and choose the parahoric and level structure at qq such that the Levi factor of the level group is LqSL(r,)L_{q}\cong\mathrm{SL}(r,\mathbb{C}). The Lie algebra of the level group is thus 𝔤D𝔰𝔩(r)\mathfrak{g}_{D}\cong\mathfrak{sl}(r), and its dual is 𝔤D𝔰𝔩(r)\mathfrak{g}_{D}^{*}\cong\mathfrak{sl}(r)^{*}.

The space of KP elliptic solitons is known to correspond to a specific coadjoint orbit in 𝔰𝔩(r)\mathfrak{sl}(r)^{*}. Let Orb(KP)\mathrm{Orb}(KP) be the coadjoint orbit of the element in 𝔰𝔩(r)\mathfrak{sl}(r)^{*} corresponding (via the Killing form) to the matrix diag(1,1,,1,r1)\mathrm{diag}(-1,-1,\dots,-1,r-1).

We define the KP symplectic leaf, M(KP,r)M(KP,r), to be the subvariety of our moduli space LH(Σ,𝒢𝜽)\mathcal{M}_{LH}(\Sigma,\mathcal{G}_{\boldsymbol{\theta}}) consisting of triples [(,φ,η)][(\mathcal{E},\varphi,\eta)] where the coresidue of the Higgs field lies in this orbit.

M(KP,r):={[(,φ,η)]LH(Σ,𝒢𝜽)μ([(,η)],φ)=CoResq(φ)Orb(KP)}.M(KP,r):=\left\{[(\mathcal{E},\varphi,\eta)]\in\mathcal{M}_{LH}(\Sigma,\mathcal{G}_{\boldsymbol{\theta}})\mid\mu([(\mathcal{E},\eta)],\varphi)=\operatorname{CoRes}_{q}(\varphi)\in\mathrm{Orb}(KP)\right\}.

This is a symplectic leaf of the Poisson manifold LH(Σ,𝒢𝜽)\mathcal{M}_{LH}(\Sigma,\mathcal{G}_{\boldsymbol{\theta}}). The results of Treibich and Verdier [33] provide a bijection between the space of solutions Sol(Λ,r)\mathrm{Sol}(\Lambda,r) and an open subset of this leaf. As a direct consequence of Theorem 6.15, we have:

Corollary 8.1.

The space Sol(Λ,r)\mathrm{Sol}(\Lambda,r) of Λ\Lambda-periodic KP elliptic solitons of order rr embeds as a Zariski open subset of the symplectic leaf M(KP,r)M(KP,r). It is thereby endowed with a canonical algebraically completely integrable Hamiltonian system structure inherited from the parahoric Hitchin fibration on LH(Σ,𝒢𝛉)\mathcal{M}_{LH}(\Sigma,\mathcal{G}_{\boldsymbol{\theta}}).

8.4. Elliptic Calogero–Moser system

The elliptic Calogero–Moser system can be realized intrinsically as a GG-Hitchin system, as demonstrated by Hurtubise and Markman in [22]. Their construction bypasses the difficulties of using semi-simple groups by defining a bespoke, non-semisimple structure group that naturally accommodates the geometry of the Calogero–Moser phase space.

Theorem 8.2 ([22], Sections 4 & 5).

Let Σ\Sigma be an elliptic curve with a marked point p0p_{0}, and let RR be a root system with torus HH and Weyl group WW. Let GHMG_{HM} be the non-semisimple group whose connected component is (αRα)H(\bigoplus_{\alpha\in R}\mathbb{C}_{\alpha})\rtimes H. The elliptic Calogero–Moser system for the root system RR is realized as a Hitchin system whose phase space is the symplectic reduction of the cotangent bundle TGHMT^{*}\mathcal{M}_{G_{HM}} of the moduli space of GHMG_{HM}-bundles on Σ\Sigma framed at p0p_{0}. The reduction is taken with respect to the action of GHMG_{HM}, where the moment map is the residue of the Higgs field at p0p_{0}, and is performed at a generic, WW-invariant coadjoint orbit element C(αRα)C\in(\bigoplus_{\alpha\in R}\mathbb{C}_{\alpha})^{*}. The resulting reduced phase space is symplectically isomorphic to (𝔥×𝔥)/Waff(\mathfrak{h}\times\mathfrak{h}^{*})/W_{\text{aff}}, where the coordinates x𝔥x\in\mathfrak{h} parameterize the underlying HH-bundle, and the points p𝔥p\in\mathfrak{h}^{*} correspond to the constant part of the Higgs field in the 𝔥\mathfrak{h}^{*} direction. The Hamiltonian of the system, HCM(x,p)=pp+αRm|α|(α(x))H_{CM}(x,p)=p\cdot p+\sum_{\alpha\in R}m_{|\alpha|}\wp(\alpha(x)), arises from the residue pairing of the canonical quadratic invariant on the Lie algebra 𝔤HM\mathfrak{g}_{HM} with the Weierstrass zeta function.

The Hurtubise–Markman construction can be precisely situated within the general framework of logahoric Higgs torsors. The elliptic Calogero–Moser system corresponds to a single symplectic leaf within the moduli space LH(Σ,𝒢𝜽)\mathcal{M}_{LH}(\Sigma,\mathcal{G}_{\boldsymbol{\theta}}) for a specific choice of data. We set the base curve to be the elliptic curve X=ΣX=\Sigma with divisor D={p0}D=\{p_{0}\}, and the structure group to be G=GHMG=G_{HM}, as in the statement of Theorem 8.2. The parahoric data at p0p_{0} is taken to be trivial (θp0=0\theta_{p_{0}}=0), corresponding to the standard parahoric group GHM([[z]])G_{HM}(\mathbb{C}[[z]]). This choice makes the level group GDGHM/Z(GHM)G_{D}\cong G_{HM}/Z(G_{HM}). The framing of the bundle at p0p_{0} used in [22] is a specific realization of a DD-level structure η\eta. The crucial identification is that the symplectic reduction in [22] at a fixed coadjoint orbit element C(αRα)C\in(\bigoplus_{\alpha\in R}\mathbb{C}_{\alpha})^{*} is equivalent to selecting the symplectic leaf μ1(AdGHM(C))\mu^{-1}(\mathrm{Ad}^{*}_{G_{HM}}(C)) in our framework, where μ\mu is the moment map. Consider the logahoric Hitchin integrable system for the group G=SL(r,)G=\mathrm{SL}(r,\mathbb{C}) over the elliptic curve X=ΣX=\Sigma with divisor D={q}D=\{q\}. We choose the trivial parahoric data at qq, so the level group is GD=SL(r,)/Z(SL(r,))G_{D}=\mathrm{SL}(r,\mathbb{C})/Z(\mathrm{SL}(r,\mathbb{C})). The moment map μSL(r)\mu_{\mathrm{SL}(r)} takes values in 𝔤D𝔰𝔩(r)\mathfrak{g}_{D}^{*}\cong\mathfrak{sl}(r)^{*}.

As described in [26] and [33], the space of elliptic solutions to the KP equation, Sol(Λ,r)\mathrm{Sol}(\Lambda,r), corresponds to the symplectic leaf defined by the coadjoint orbit of the element in 𝔰𝔩(r)\mathfrak{sl}(r)^{*} corresponding to the matrix diag(1,1,,1,r1)\mathrm{diag}(-1,-1,\dots,-1,r-1). We denote this orbit by Orb(KP)\mathrm{Orb}(KP). The KP symplectic leaf is thus:

M(KP,r):=μSL(r)1(Orb(KP))LH(Σ,𝒢SL(r),𝜽).M(KP,r):=\mu_{\mathrm{SL}(r)}^{-1}(\mathrm{Orb}(KP))\subset\mathcal{M}_{LH}(\Sigma,\mathcal{G}_{\mathrm{SL}(r),\boldsymbol{\theta}}).

The connection is then established by the embedding theorem from Hurtubise and Markman.

Theorem 8.3 ([22], Section 6).

For the root system R=Ar1R=A_{r-1}, there exists an equivariant embedding of the Hurtubise–Markman Hitchin system for GHMG_{HM} into the standard Hitchin system for GL(r,)\mathrm{GL}(r,\mathbb{C}). This embedding maps the phase space of the Ar1A_{r-1} Calogero–Moser system precisely onto the symplectic leaf that describes the elliptic solutions of the KP hierarchy.

Corollary 8.4.

The symplectic leaf M(KP,r)LH(Σ,𝒢SL(r),𝛉)M(KP,r)\subset\mathcal{M}_{LH}(\Sigma,\mathcal{G}_{\mathrm{SL}(r),\boldsymbol{\theta}}) corresponding to the KP hierarchy is symplectically isomorphic to the symplectic leaf

μGHM1(AdGHM(C))LH(Σ,𝒢GHM,𝜽)\mu_{G_{HM}}^{-1}(\mathrm{Ad}^{*}_{G_{HM}}(C))\subset\mathcal{M}_{LH}(\Sigma,\mathcal{G}_{G_{HM}},\boldsymbol{\theta^{\prime}})

corresponding to the Ar1A_{r-1} Calogero–Moser system.

Therefore, the general framework of logahoric Higgs torsors provides a unified geometric perspective. It demonstrates that the phase space for elliptic KP solutions is a specific symplectic leaf of the SL(r,)\mathrm{SL}(r,\mathbb{C})-Hitchin system, and that this very same leaf can be described intrinsically for the type Ar1A_{r-1} root system via the Hurtubise–Markman construction, yielding the corresponding Calogero–Moser system.

Acknowledgements. We warmly thank Pengfei Huang and Hao Sun for many fruitful discussions and remarks. We are also very grateful to the following Research Institutes for their support and hospitality, where part of this work was completed: The Kavli Institute for the Physics and Mathematics of the Universe at the University of Tokyo, the Brin Mathematics Research Center at the University of Maryland and the Tianyuan Mathematics Research Center. Georgios Kydonakis was supported by the Scientific Committee of the University of Patras through the program “Medicos”. Lutian Zhao was supported by JSPS KAKENHI Grant Number JP25K17226.

References

  • [1] Arnol’d, V. I., and Givental, A. B., Symplectic geometry. in Dynamical Systems IV, Encyclopedia of Mathematical Sciences, vol. 4 (Springer, 1990).
  • [2] Balaji, V., and Seshadri, C. S., “Moduli of parahoric 𝒢\mathcal{G}-torsors on a compact Riemann surface”. J. Algebraic Geom. 24 (2015), no. 1, 1–49.
  • [3] Baraglia, D., Kamgarpour, M., and Varma, R., “Complete integrability of the parahoric Hitchin system”. Int. Math. Res. Not. 2019 (2019), no. 21, 6499–6528.
  • [4] Beauville, A., “Jacobiennes des courbes spectrales et systèmes hamiltoniens complètement intégrables”. Acta Math. 164 (1990), no. 3–4, 211–235.
  • [5] Beauville, A., and Laszlo, Y., “Un lemme de descente”. C. R. Acad. Sci. Paris 320 (1995).
  • [6] Biswas, I., and Hoffmann, N., “A Torelli theorem for moduli spaces of principal bundles over a curve”. Ann. Inst. Fourier 62 (2012), no. 1, 87–106.
  • [7] Biswas, I., Logares, M., and Peón-Nieto, A., “Moduli spaces of framed GG-Higgs bundles and symplectic geometry”. Commun. Math. Phys. 376 (2020), no. 3, 1875–1908.
  • [8] Boalch, P., “Riemann–Hilbert for tame complex parahoric connections”. Transform. Groups 16 (2011), 27–50.
  • [9] Bottacin, F., “Symplectic geometry on moduli spaces of stable pairs”. Ann. Sci. Éc. Norm. Supér. (4) 28 (1995), no. 4, 391–433.
  • [10] Chernousov, V., Gille, P., and Pianzola, A., “Torsors over the punctured affine line”. Amer. J. Math. 134 (2012), no. 6, 1541–1583.
  • [11] Damiolini, C., and Hong, J., “Local types of (Γ,G)(\Gamma,G)-bundles and parahoric group schemes”. Proc. LMS 27 (2023).
  • [12] Damiolini, C., and Hong, J., “Line bundles on the moduli stack of parahoric bundles”. arXiv:2412.08826 (2024).
  • [13] de Cataldo, M. A. A., Heinloth, J., Migliorini, L., “A support theorem for the Hitchin fibration: the case of GLn\mathrm{GL}_{n} and KCK_{C}”. J. Reine Angew. Math. 780 (2021), 41–77.
  • [14] Donagi, R., Decomposition of spectral covers. Journées de géométrie algébrique d’Orsay, France, juillet 20-26, 1992. Paris: Société Mathématique de France, Astérisque 218, 145-175 (1993)
  • [15] Donagi, R., and Markman, E., Spectral covers, algebraically completely integrable, Hamiltonian systems, and moduli of bundles. in Francaviglia, M. (ed.) et al., Integrable systems and quantum groups. Lectures given at the 1st session of the Centro Internazionale Matematico Estivo (CIME) held in Montecatini Terme, Italy, June 14-22, 1993. Berlin: Springer-Verlag. Lect. Notes Math. 1620, 1–119 (1996).
  • [16] Faltings, G.: “Stable GG-bundles and projective connections”. J. Algebr. Geom. 2 (1993), No. 3, 507–568.
  • [17] Feigin, B., Frenkel, E., and Reshetikhin, N. “Gaudin Model, Bethe Ansatz and Critical Level”. Commun. Math. Phys. 166 (1994), No. 1, 27–62.
  • [18] Haines, T., and Rapoport, M., “On parahoric subgroups”. Adv. Math. 219 (2008), no. 1, 188–198.
  • [19] Heinloth, J., “Uniformization of 𝒢\mathcal{G}-bundles”. Math. Ann. 347 (2010), no. 3, 499–528.
  • [20] Hitchin, N., “Stable bundles and integrable systems”. Duke Math. J. 54 (1987), 91–114.
  • [21] Huang, P., Kydonakis, G., Sun, H., and Zhao, L., “Tame parahoric nonabelian Hodge correspondence on curves”. arXiv:2205.15475 (2022).
  • [22] Hurtubise, J. C., and Markman, E., “Calogero–Moser Systems and Hitchin Systems”. Commun. Math. Phys. 223 (2001), 533–582.
  • [23] Kydonakis, G., Sun, H., and Zhao, L., “Poisson structures on moduli spaces of Higgs bundles over stacky curves”. Adv. Geom. 24 (2024), no. 2, 163–182.
  • [24] Kydonakis, G., Sun, H., and Zhao, L., “Logahoric Higgs torsors for a complex reductive group”. Math. Ann. 388(3) (2024), 3183-3228.
  • [25] Logares, M., and Martens, J., “Moduli of parabolic Higgs bundles and Atiyah algebroids”. J. Reine Angew. Math. 649 (2010), 89–116.
  • [26] Markman, E., “Spectral curves and integrable systems”. Compositio Math. 93 (1994), no. 3, 255–290.
  • [27] Moy, A., and Prasad, G., “Unrefined minimal K-types for pp-adic groups”. Invent. Math. 116 (1994), 393–408.
  • [28] Nguyen, T. Q. T., Parahoric Hitchin fibration. Ph.D. thesis, University of Edinburgh (2021).
  • [29] Ramanathan, A., “Stable principal bundles on a compact Riemann surface”. Math. Ann. 213 (1975), no. 2, 129–152.
  • [30] Reyman, A. G., and Semenov-Tyan-Shansky, M. A., Group theoretical methods in the theory of finite dimensional integrable systems. in: Dynamical Systems VII, Arnol’d, V. I. and Novikov, S. P. (eds.), Encyclopedia of Mathematical Sciences, vol. 16, pp. 119–193 (1987) (Russian).
  • [31] Scognamillo, R., “An elementary approach to the abelianization of the Hitchin system for arbitrary reductive groups”. Compos. Math. 110 (1998), No. 1, 17–37
  • [32] Seshadri, C. S., Fibrés vectoriels sur les courbes algébriques. Astérisque, vol. 96 (1982).
  • [33] Treibich,A., and Verdier, J.-L., Solitons elliptiques. in The Grothendieck Festschrift, Vol. III, 437–480, Progress in Mathematics vol. 88, Birkhäuser Boston, Boston, MA (1990).
  • [34] Vanhaecke, P., Integrable systems in the realm of algebraic geometry. 2nd ed. Lecture Notes in Mathematics 1638. Berlin: Springer. x, 256 p. (2001).
  • [35] Wang, B., “On parahoric Hitchin systems over curves”. Int. J. Math. 34 (2023), no. 13, 2350081.
  • [36] Yun, Z., “Global Springer theory”. Adv. Math. 228 (2011), 266–328.
  • [37] Zelaci, H., Moduli spaces of anti-invariant vector bundles over curves and conformal blocks, Ph.D. thesis, COMUE Université Côte d’Azur (2017).

Department of Mathematics, University of Patras
University Campus, Patras 26504, Greece
E-mail address: gkydonakis@math.upatras.gr


Kavli Institute for the Physics and Mathematics of the Universe, The University of Tokyo
5-1-5 Kashiwanoha, Kashiwa, Chiba, 277-8583, Japan
E-mail address: lutian.zhao@ipmu.jp