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

\customizeamsrefs

Lectures on Springer theories and orbital integrals

Zhiwei Yun Department of Mathematics, Stanford University, 450 Serra Mall, Building 380, Stanford, CA 94305 zwyun@stanford.edu
Abstract.

These are the expanded lecture notes from the author’s mini-course during the graduate summer school of the Park City Math Institute in 2015. The main topics covered are: geometry of Springer fibers, affine Springer fibers and Hitchin fibers; representations of (affine) Weyl groups arising from these objects; relation between affine Springer fibers and orbital integrals.

Research supported by the NSF grant DMS-1302071, the Packard Foundation and the PCMI.

0. Introduction

0.1. Topics of these notes

These are the lectures notes from a mini-course that I gave at the PCMI graduate summer school in 2015. The goal is twofold. First I would like to introduce to the audience some interesting geometric objects that have representation-theoretic applications, and Springer fibers and their generalizations are nice examples of such. Second is to introduce orbital integrals with emphasis on its relationship with affine Springer fibers, and thereby supplying background material for B-C. Ngô’s mini-course on the Fundamental Lemmas.

The geometric part of these lectures (everything except §3) consists of the study of three types of “fibers”: Springer fibers, affine Springer fibers and Hitchin fibers, with increasing complexity. We will study their geometric properties such as connectivity and irreducible components. We will construct certain group actions on these varieties, and use the action to study several nontrivial examples. Most importantly we will study certain Weyl group actions on the cohomology of these fibers which do not come from actions on the varieties themselves. The representation-theoretic significance of these three types of fibers and the analogy between them can be summarized in the following table.

Springer fibers affine Springer fibers Hitchin fibers
field kk local field F=k((t))F=k(\!(t)\!) global field k(X)k(X)
symmetry WW W~\widetilde{W} W~\widetilde{W}
extended sym graded AHA graded DAHA graded DAHA
rep theory characters of orbital integrals trace formula
when k=𝔽qk=\mathbb{F}_{q} G(k)G(k) for G(F)G(F) for GG over k(X)k(X)

Here AHAAHA stands for the Affine Hecke Algebra, while DAHA stands for the Double Affine Hecke Algebra; XX denotes an algebraic curve over kk; WW and W~\widetilde{W} are the Weyl group and extended affine Weyl group.

In these lecture notes we do not try to give complete proofs to all statements but instead to point out interesting phenomena and examples. We do, however, give more or less complete proofs of several key results, such as

  • Theorem 3 (the Springer correspondence);

  • Theorem 17 (finiteness properties of affine Springer fibers);

  • Theorem 39 (cohomological interpretation of stable orbital integrals).

0.2. What we assume from the readers

The target readers for these lectures are beginning graduate students interested in geometric representation theory. We assume some basic algebraic geometry (scheme theory, moduli problems, point counting over a finite field, etc) though occasionally we will use the language of algebraic stacks. We also assume some Lie theory (reductive groups over an algebraic closed field and over a local field), but knowing GLn\textup{GL}_{n} and SLn\textup{SL}_{n} should be enough to understand most of these notes.

The next remark is about the cohomology theory we use in these notes. Since we work with algebraic varieties over a general field kk instead of \mathbb{C}, we will be using the étale cohomology with coefficients in \ell-adic sheaves (usually the constant sheaf \mathbb{Q}_{\ell}) on these varieties. We denote the étale cohomology of a scheme XX over kk with \mathbb{Q}_{\ell}-coefficients simply by H(X)\textup{H}^{*}({X}). Readers not familiar with étale cohomology are encouraged to specialize to the case k=k=\mathbb{C} and understand H(X)\textup{H}^{*}({X}) as the singular cohomology of X()X(\mathbb{C}) with \mathbb{Q}-coefficients. Perverse sheaves will be used only in §1.5.

Acknowledgement

I would like to thank the co-organizers, lecturers and the staff of the PCMI summer program in 2015. I would also like to thank the audience of my lectures for their feedback. I am especially grateful to Jingren Chi who carefully read through the first draft of these notes and provided helpful suggestions.

1. Lecture I: Springer fibers

Springer fibers are classical and fundamental objects in geometric representation theory. Springer [Springer] first discovered that their cohomology groups realized representations of Weyl groups, a phenomenon known as the Springer correspondence. As singular algebraic varieties, Springer fibers are interesting geometric objects by themselves. They are also connected to the representation theory of finite groups of Lie type via character sheaves.

1.1. The setup

In this section, let kk be an algebraically closed field. Let GG be a connected reductive group over kk whose adjoint group is simple (so the adjoint group GadG^{\textup{ad}} is determined by one of the seven series of Dynkin diagrams). Assume that char(k)\textup{char}(k) is large compared to GG. Let rr be the rank of GG.

Sometimes it will be convenient to fix a Cartan subalgebra 𝔱\mathfrak{t} of 𝔤\mathfrak{g}, or equivalently a maximal torus TGT\subset G. Once we have done this, we may talk about the roots of the TT-action on 𝔤\mathfrak{g}.

Let \mathcal{B} be the flag variety of GG: this is the GG-homogeneous projective variety parametrizing Borel subgroups of GG. Choosing a Borel subgroup BGB\subset G, we may identify \mathcal{B} with G/BG/B.

Let 𝔤\mathfrak{g} be the Lie algebra of GG. For X𝔤X\in\mathfrak{g}, let CG(X)C_{G}(X) denote the centralizer of XX in GG, i.e., the stabilizer at XX of the adjoint action of GG on 𝔤\mathfrak{g}.

Let 𝒩𝔤\mathcal{N}\subset\mathfrak{g} be the subvariety of nilpotent elements. This is a cone: it is stable under the action of 𝔾m\mathbb{G}_{m} on 𝔤\mathfrak{g} by scaling. It is known that there are finitely many GG-orbits on 𝒩\mathcal{N} under the adjoint action.

1.2. Springer fibers

1.2.1. The Springer resolution

The cotangent bundle 𝒩~:=T\widetilde{\mathcal{N}}:=T^{*}\mathcal{B} classifies pairs (e,B)(e,B) where e𝒩e\in\mathcal{N} and BB is a Borel subgroup of GG such that e𝔫e\in\mathfrak{n}, where 𝔫\mathfrak{n} is the nilpotent radical of LieB\textup{Lie}\ B. The Springer resolution is the forgetful map

π:𝒩~𝒩\pi:\widetilde{\mathcal{N}}\to\mathcal{N}

sending (e,B)(e,B) to ee. This map is projective.

For e𝒩e\in\mathcal{N}, the fiber e:=π1(e)\mathcal{B}_{e}:=\pi^{-1}(e) is called the Springer fiber of ee. By definition, e\mathcal{B}_{e} is the closed subscheme of \mathcal{B} consisting of those Borel subgroups BGB\subset G such that ee is contained in the nilpotent radical of LieB\textup{Lie}\ B.

1.2.2. The Grothendieck alteration

Consider the variety 𝔤~\widetilde{\mathfrak{g}} of pairs (X,B)(X,B) where X𝔤X\in\mathfrak{g} and BB\in\mathcal{B} such that XLieBX\in\textup{Lie}\ B. The forgetful map (X,B)X(X,B)\mapsto X

π𝔤:𝔤~𝔤\pi_{\mathfrak{g}}:\widetilde{\mathfrak{g}}\to\mathfrak{g}

is called the Grothendieck alteration 111The term “alteration” refers to a proper, generically finite map whose source is smooth over kk., also known as the Grothendieck simultaneous resolution.

Let 𝔠=𝔤G𝔱W\mathfrak{c}=\mathfrak{g}\sslash G\cong\mathfrak{t}\sslash W be the categorical quotient of 𝔤\mathfrak{g} by the adjoint action of GG. Then we have a commutative diagram

(1) 𝔤~\textstyle{\widetilde{\mathfrak{g}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π𝔤\scriptstyle{\pi_{\mathfrak{g}}}χ~\scriptstyle{\widetilde{\chi}}𝔱\textstyle{\mathfrak{t}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔤\textstyle{\mathfrak{g}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}χ\scriptstyle{\chi}𝔠\textstyle{\mathfrak{c}}

Here χ:𝔤𝔠\chi:\mathfrak{g}\to\mathfrak{c} is the natural quotient map and χ~:𝔤~𝔱\widetilde{\chi}:\widetilde{\mathfrak{g}}\to\mathfrak{t} sends (X,B)𝔤~(X,B)\in\widetilde{\mathfrak{g}} to the image of XX in 𝔟/𝔫\mathfrak{b}/\mathfrak{n} (where 𝔟=LieB\mathfrak{b}=\textup{Lie}\ B with nilpotent radical 𝔫\mathfrak{n}), which can be canonically identified with 𝔱\mathfrak{t} (upon choosing a Borel containing 𝔱\mathfrak{t}). The diagram (1) is Cartesian when restricting the left column to regular elements 222An element X𝔤X\in\mathfrak{g} is called regular if its centralizer in GG has dimension rr, the rank of GG. in 𝔤\mathfrak{g}. In particular, if we restrict the diagram (1) to the regular semisimple locus 𝔠rs𝔠\mathfrak{c}^{\textup{rs}}\subset\mathfrak{c} 333An element X𝔱X\in\mathfrak{t} is regular semisimple if α(X)0\alpha(X)\neq 0 for any root α\alpha. Denote by 𝔱rs𝔱\mathfrak{t}^{\textup{rs}}\subset\mathfrak{t} the open subscheme of regular semisimple elements in 𝔱\mathfrak{t}. Since 𝔱rs\mathfrak{t}^{\textup{rs}} is stable under WW, it is the full preimage of an open subset 𝔠rs𝔠\mathfrak{c}^{\textup{rs}}\subset\mathfrak{c}. The open subset 𝔤rs=χ1(𝔠rs)\mathfrak{g}^{\textup{rs}}=\chi^{-1}(\mathfrak{c}^{\textup{rs}}) is by definition the locus of regular semisimple elements of 𝔤\mathfrak{g}., we see that π𝔤|𝔤~rs:𝔤~rs𝔤rs\pi_{\mathfrak{g}}|_{\widetilde{\mathfrak{g}}^{\textup{rs}}}:\widetilde{\mathfrak{g}}^{\textup{rs}}\to\mathfrak{g}^{\textup{rs}} is a WW-torsor; i.e., there is an action of WW on π𝔤1(𝔤rs)\pi_{\mathfrak{g}}^{-1}(\mathfrak{g}^{\textup{rs}}) preserving the projection to 𝔤rs\mathfrak{g}^{\textup{rs}} making the fibers of π𝔤\pi_{\mathfrak{g}} into principal homogeneous spaces for WW. The map π𝔤\pi_{\mathfrak{g}} becomes branched but still finite over the regular locus 𝔤reg𝔤\mathfrak{g}^{\textup{reg}}\subset\mathfrak{g}.

For X𝔤X\in\mathfrak{g}, we denote the fiber π𝔤1(X)\pi_{\mathfrak{g}}^{-1}(X) simply by ~X\widetilde{\mathcal{B}}_{X}. Restricting π𝔤\pi_{\mathfrak{g}} to 𝒩\mathcal{N}, the fibers ~e\widetilde{\mathcal{B}}_{e} for a nilpotent element e𝒩e\in\mathcal{N} is the closed subvariety of \mathcal{B} consisting of those BB such that eLieBe\in\textup{Lie}\ B. Clearly e\mathcal{B}_{e} is a subscheme of ~e\widetilde{\mathcal{B}}_{e}, and the two schemes have the same reduced structure. However, as schemes, ~e\widetilde{\mathcal{B}}_{e} and e\mathcal{B}_{e} are different in general. See §1.3.2 below and Exercise 1.7.2.

1.3. Examples of Springer fibers

1.3.1.

When e=0e=0, e=\mathcal{B}_{e}=\mathcal{B}.

1.3.2. Regular nilpotent elements

The unique dense open GG-orbit consists of regular nilpotent elements, i.e., those ee such that dimCG(e)=r\dim C_{G}(e)=r. When ee is a regular nilpotent element, e\mathcal{B}_{e} is a single point: there is a unique Borel subgroup of GG whose Lie algebra contains ee. What is this Borel subgroup?

By the Jacobson-Morosov theorem, we may extend ee to an 𝔰𝔩2\mathfrak{sl}_{2}-triple (e,h,f)(e,h,f) in 𝔤\mathfrak{g}. The adjoint action of hh on 𝔤\mathfrak{g} has integer weights, and it decomposes 𝔤\mathfrak{g} into weight spaces 𝔤(n)\mathfrak{g}(n), nn\in\mathbb{Z}. Let 𝔟=n0𝔤(n)\mathfrak{b}=\oplus_{n\geq 0}\mathfrak{g}(n). This is a Borel subalgebra of 𝔤\mathfrak{g}, and the corresponding Borel subgroup is the unique point in e\mathcal{B}_{e}.

When ee is regular, the fiber ~e\widetilde{\mathcal{B}}_{e} of the Grothendieck alteration is a non-reduced scheme whose underlying reduced scheme is a point. The coordinate ring of ~e\widetilde{\mathcal{B}}_{e} is isomorphic to the coinvariant algebra Sym(𝔱)/(Sym(𝔱)+W)\textup{Sym}(\mathfrak{t}^{*})/(\textup{Sym}(\mathfrak{t}^{*})^{W}_{+}), which is, interestingly, also isomorphic to the cohomology ring with kk-coefficients of the complex flag variety \mathcal{B}_{\mathbb{C}}.

1.3.3.

When G=SL(V)G=\textup{SL}(V) for some vector space VV of dimension nn, \mathcal{B} is the moduli space of full flags 0=V0V1V2Vn1Vn=V0=V_{0}\subset V_{1}\subset V_{2}\subset\cdots\subset V_{n-1}\subset V_{n}=V. The Springer fiber e\mathcal{B}_{e} consists of those flags such that eViVi1eV_{i}\subset V_{i-1}.

1.3.4.

Consider the case G=SL3G=\textup{SL}_{3} and e=(001000000)e=\left(\begin{array}[]{ccc}0&0&1\\ 0&0&0\\ 0&0&0\end{array}\right) under the standard basis {v1,v2,v3}\{v_{1},v_{2},v_{3}\} of VV. Then e\mathcal{B}_{e} is the union of two 1\mathbb{P}^{1}s: the first 1\mathbb{P}^{1} consisting of flags 0V1v1,v2V0\subset V_{1}\subset\langle{v_{1},v_{2}}\rangle\subset V with varying V1V_{1} inside the fixed plane ker(e)=v1,v2\ker(e)=\langle{v_{1},v_{2}}\rangle; the second 1\mathbb{P}^{1} consisting of flags 0v1V2V0\subset\langle{v_{1}}\rangle\subset V_{2}\subset V with a varying V2V_{2} containing Im(e)=v1\textup{Im}(e)=\langle{v_{1}}\rangle.

1.3.5.

Consider the case G=SL4G=\textup{SL}_{4} and the nilpotent element e:v3v10,v4v20e:v_{3}\mapsto v_{1}\mapsto 0,v_{4}\mapsto v_{2}\mapsto 0 under the standard basis {v1,v2,v3,v4}\{v_{1},v_{2},v_{3},v_{4}\}. If a flag 0V1V2V3V0\subset V_{1}\subset V_{2}\subset V_{3}\subset V is in e\mathcal{B}_{e}, then V1ker(e)=v1,v2V_{1}\subset\ker(e)=\langle{v_{1},v_{2}}\rangle, and V3Im(e)=v1,v2V_{3}\supset\textup{Im}(e)=\langle{v_{1},v_{2}}\rangle. We denote H:=v1,v2H:=\langle{v_{1},v_{2}}\rangle. There are two cases:

  1. (1)

    V2=HV_{2}=H. Then we may choose V1V_{1} to be any line in HH and V3V_{3} to be any hyperplane containing V2V_{2}. We get a closed subvariety of e\mathcal{B}_{e} isomorphic to (H)×(H)1×1\mathbb{P}(H)\times\mathbb{P}(H)\cong\mathbb{P}^{1}\times\mathbb{P}^{1}. We denote this closed subvariety of e\mathcal{B}_{e} by C1C_{1}.

  2. (2)

    V2HV_{2}\neq H. This defines an open subscheme UU of e\mathcal{B}_{e}. Suppose the line V1HV_{1}\subset H is spanned by av1+bv2av_{1}+bv_{2} for some [a:b](H)[a:b]\in\mathbb{P}(H), then the image of V2V_{2} in V/H=v3,v4V/H=\langle{v_{3},v_{4}}\rangle is spanned by av3+bv4av_{3}+bv_{4}. Fixing V1V_{1}, the choices of V2V_{2} are given by Hom(av3+bv4,H/V1)Hom(V1,H/V1)\textup{Hom}(\langle{av_{3}+bv_{4}}\rangle,H/V_{1})\cong\textup{Hom}(V_{1},H/V_{1}). Once V2V_{2} is fixed, V3=V2+HV_{3}=V_{2}+H is also fixed. Therefore UU is isomorphic to the total space of the line bundle 𝒪(2)\mathcal{O}(2) over (H)1\mathbb{P}(H)\cong\mathbb{P}^{1}.

From the above discussion we see that e\mathcal{B}_{e} has dimension 22, C1C_{1} is an irreducible component of e\mathcal{B}_{e} and so is the closure of UU, which we denote by C2C_{2}. We have C1C2=eC_{1}\cup C_{2}=\mathcal{B}_{e} and C1C2C_{1}\cap C_{2} is the diagonal inside C11×1C_{1}\cong\mathbb{P}^{1}\times\mathbb{P}^{1}.

1.3.6. Components of type AA Springer fibers

When G=SLn=SL(V)G=\textup{SL}_{n}=\textup{SL}(V), Spaltenstein [Spa76] and Steinberg [St] gave a description of the irreducible components of e\mathcal{B}_{e} using standard Young tableaux of size nn. This will be relevant to the Springer correspondence that we will discuss later, see §1.5.8. Below we follow the presentation of [Spaltenstein, Ch II, §5].

Fix a nilpotent element e𝒩e\in\mathcal{N} whose Jordan type is a partition λ\lambda of nn. This means, if the partition λ\lambda is n=λ1+λ2+n=\lambda_{1}+\lambda_{2}+\cdots, ee has Jordan blocks of sizes λ1,λ2,\lambda_{1},\lambda_{2},\cdots. We shall construct a (non-algebraic) map eST(λ)\mathcal{B}_{e}\to ST(\lambda), where ST(λ)ST(\lambda) is the discrete set of standard Young tableau for the partition λ\lambda. For each full flag 0=V0V1Vn1Vn=V0=V_{0}\subset V_{1}\subset\cdots\subset V_{n-1}\subset V_{n}=V such that eViVi1eV_{i}\subset V_{i-1}, ee induces a nilpotent endomorphism of V/VniV/V_{n-i}. Let μi\mu_{i} be the Jordan type of the ee on V/VniV/V_{n-i}, then μi\mu_{i} is a partition of ii. The increasing sequence of partitions μ1,μ2,,μn=λ\mu_{1},\mu_{2},\cdots,\mu_{n}=\lambda satisfies that μi\mu_{i} is obtained from μi1\mu_{i-1} by increasing one part of μi1\mu_{i-1} by 1 (including creating a part of size 11). This gives an increasing sequence Y1,,Yn=Y(λ)Y_{1},\cdots,Y_{n}=Y(\lambda) of subdiagrams of the Young diagram Y(λ)Y(\lambda) of λ\lambda. We label the unique box in YiYi1Y_{i}-Y_{i-1} by ii to get a standard Young tableau.

Spaltenstein [Spaltenstein, Ch II, Prop 5.5] showed that the closure of the preimage of each standard Young tableaux in e\mathcal{B}_{e} is an irreducible component. Moreover, all irreducible components of e\mathcal{B}_{e} arise in this way and they all have the same dimension de=12iλi(λi1)d_{e}=\frac{1}{2}\sum_{i}\lambda^{*}_{i}(\lambda^{*}_{i}-1), where λ\lambda^{*} is the conjugate partition of λ\lambda. In particular, the top dimensional cohomology H2de(e)\textup{H}^{2d_{e}}({\mathcal{B}_{e}}) has dimension equal to #ST(λ)\#ST(\lambda), which is also the dimension of an irreducible representation of the symmetric group SnS_{n}. This statement is a numerical shadow of the Springer correspondence, which says that H2de(e)\textup{H}^{2d_{e}}({\mathcal{B}_{e}}) is naturally an irreducible representation of SnS_{n}.

Spaltenstein [Spaltenstein, Ch II, Prop 5.9] also showed that there exists a stratification of \mathcal{B} into affine spaces such that e\mathcal{B}_{e} is a union of strata. This implies that the restriction map on cohomology H()H(e)\textup{H}^{*}({\mathcal{B}})\to\textup{H}^{*}({\mathcal{B}_{e}}) is surjective.

1.3.7.

Consider the case G=Sp(V)G=\textup{Sp}(V) for some symplectic vector space VV of dimension 2n2n, then \mathcal{B} is the moduli space of full flags

0=V0V1V2VnV2n1V2n=V0=V_{0}\subset V_{1}\subset V_{2}\subset\cdots\subset V_{n}\subset\cdots\subset V_{2n-1}\subset V_{2n}=V

such that Vi=V2niV^{\bot}_{i}=V_{2n-i} for i=1,,ni=1,\cdots,n. The Springer fiber e\mathcal{B}_{e} consists of those flags such that eViVi1eV_{i}\subset V_{i-1} for all ii.

Consider the case where dimV=4\dim V=4. We choose a basis {v1,v2,v3,v4}\{v_{1},v_{2},v_{3},v_{4}\} for VV such that the symplectic form ω\omega on VV satisfies ω(vi,v5i)=1\omega(v_{i},v_{5-i})=1 if i=1i=1 and 22, and ω(vi,vj)=0\omega(v_{i},v_{j})=0 for i+j5i+j\neq 5. Let ee be the nilpotent element in 𝔤=𝔰𝔭(V)\mathfrak{g}=\mathfrak{sp}(V) given by e:v4v10,v30,v20e:v_{4}\mapsto v_{1}\mapsto 0,v_{3}\mapsto 0,v_{2}\mapsto 0. Then a flag 0V1V2V1V0\subset V_{1}\subset V_{2}\subset V_{1}^{\bot}\subset V in e\mathcal{B}_{e} must satisfy v1V2v1,v2,v3\langle{v_{1}}\rangle\subset V_{2}\subset\langle{v_{1},v_{2},v_{3}}\rangle, and this is the only condition for it to lie in e\mathcal{B}_{e} (Exercise 1.7.3). Such a V2V_{2} corresponds to a line V2/v1v2,v3V_{2}/\langle{v_{1}}\rangle\subset\langle{v_{2},v_{3}}\rangle, hence a point in 1=(v2,v3)\mathbb{P}^{1}=\mathbb{P}(\langle{v_{2},v_{3}}\rangle). Over this 1\mathbb{P}^{1} we have a tautological rank two bundle 𝒱2\mathcal{V}_{2} whose fiber at V2/v1V_{2}/\langle{v_{1}}\rangle is the two-dimensional vector space V2V_{2}. The further choice of V1V_{1} gives a point in the projectivization of 𝒱2\mathcal{V}_{2}. The exact sequence 0v1V2V2/v100\to\langle{v_{1}}\rangle\to V_{2}\to V_{2}/\langle{v_{1}}\rangle\to 0 gives an exact sequence of vector bundles 0𝒪𝒱2𝒪(1)00\to\mathcal{O}\to\mathcal{V}_{2}\to\mathcal{O}(-1)\to 0 over 1\mathbb{P}^{1}. Therefore 𝒱2\mathcal{V}_{2} is isomorphic to 𝒪(1)𝒪\mathcal{O}(-1)\oplus\mathcal{O}, and e(𝒪(1)𝒪)\mathcal{B}_{e}\cong\mathbb{P}(\mathcal{O}(-1)\oplus\mathcal{O}) is a Hirzebruch surface.

1.3.8. Subregular Springer fibers

The example considered in §1.3.4 is a simplest case of a subregular Springer fiber. There is a unique nilpotent orbit 𝒪subreg\mathcal{O}_{\operatorname{subreg}} of codimension 2 in 𝒩\mathcal{N}, which is called the subregular nilpotent orbit. For e𝒪subrege\in\mathcal{O}_{\operatorname{subreg}}, it is known that e\mathcal{B}_{e} is a union of 1\mathbb{P}^{1}’s whose configuration we now describe. We may form the dual graph to e\mathcal{B}_{e} whose vertices are the irreducible components of e\mathcal{B}_{e} and two vertices are joined by an edge if the two corresponding components intersect (it turns out that they intersection at a single point).

For simplicity assume GG is of adjoint type. Let GG^{\prime} be another adjoint simple group whose type is defined as follows. When GG is simply-laced, take G=GG^{\prime}=G. When GG is of type BnB_{n}, CnC_{n}, F4F_{4} and G2G_{2}, take GG^{\prime} to be of type A2n1A_{2n-1}, Dn+1D_{n+1}, E6E_{6} and D4D_{4} respectively. One can show that e\mathcal{B}_{e} is always a union of 1\mathbb{P}^{1} whose dual graph is the Dynkin diagram of GG^{\prime}. The rule in the non-simply-laced case is that each long simple root corresponds to 2 or 3 1\mathbb{P}^{1}’s while each short simple root corresponds to a unique 1\mathbb{P}^{1}. Such a configuration of 1\mathbb{P}^{1}’s is called a Dynkin curve, see [StConj, §3.10, Definition and Prop 2] and [Slodowy, §6.3].

For example, when GG is of type AnA_{n}, then e\mathcal{B}_{e} is a chain consisting of nn 1\mathbb{P}^{1}’s: e=C1C2Cn\mathcal{B}_{e}=C_{1}\cup C_{2}\cup\cdots\cup C_{n} with CiCi+1C_{i}\cap C_{i+1} a point and otherwise disjoint.

Brieskorn [Brieskorn], following suggestions of Grothendieck, related the singularity of the nilpotent cone along the subregular orbits with Kleinian singularities, and he also realized the semi-universal deformation of this singularity inside 𝔤\mathfrak{g}. Assume GG is simply-laced. One can construct a transversal slice Se𝒩S_{e}\subset\mathcal{N} through ee of dimension two such that SeS_{e} consists of regular elements except ee. Then SeS_{e} is a normal surface with a Kleinian singularity at ee of the same type as the Dynkin diagram of GG. 444A Kleinian singularity is a surface singularity analytically isomorphic to the singularity at (0,0)(0,0) of the quotient of 𝔸2\mathbb{A}^{2} by a finite subgroup of SL2\textup{SL}_{2}.. The preimage S~e:=π1(Se)𝒩~\widetilde{S}_{e}:=\pi^{-1}(S_{e})\subset\widetilde{\mathcal{N}} turns out to be the minimal resolution of SeS_{e}, and hence e\mathcal{B}_{e} is the union of exceptional divisors. Upon identifying the components of e\mathcal{B}_{e} with simple roots of GG, the intersection matrix of the exceptional divisors is exactly the negative of the Cartan matrix of GG. Slodowy [Slodowy] extended the above picture to non-simply-laced groups, and we refer to his book [Slodowy] for a beautiful account of the connection between the subregular orbit and Kleinian singularities.

1.4. Geometric Properties of Springer fibers

1.4.1. Connectivity

The Springer fibers e\mathcal{B}_{e} are connected. See [StConj, p.131, Prop 1], [Spaltenstein, Ch II, Cor 1.7], and Exercise 1.7.11.

1.4.2. Equidimensionality

Spaltenstein [Spa77], [Spaltenstein, Ch II, Prop 1.12] showed that all irreducible components of e\mathcal{B}_{e} have the same dimension.

1.4.3. The dimension formula

Let ded_{e} be the dimension of e\mathcal{B}_{e}. Steinberg [St, Thm 4.6] and Springer [Springer] showed that

(2) de=12(dim𝒩dimGe)=12(dimCG(e)r).d_{e}=\frac{1}{2}(\dim\mathcal{N}-\dim G\cdot e)=\frac{1}{2}(\dim C_{G}(e)-r).

1.4.4. Centralizer action

The group G×𝔾mG\times\mathbb{G}_{m} acts on 𝒩\mathcal{N} (where 𝔾m\mathbb{G}_{m} by dilation). Let G~e=StabG×𝔾m(e)\widetilde{G}_{e}=\textup{Stab}_{G\times\mathbb{G}_{m}}(e) be the stabilizer. Then G~e\widetilde{G}_{e} acts on e\mathcal{B}_{e}. Note that G~e\widetilde{G}_{e} always surjective onto 𝔾m\mathbb{G}_{m} with kernel CG(e)C_{G}(e).

The action of G~e\widetilde{G}_{e} on H(e)\textup{H}^{*}({\mathcal{B}_{e}}) factors through the finite group Ae=π0(G~e)=π0(Ge)A_{e}=\pi_{0}(\widetilde{G}_{e})=\pi_{0}(G_{e}). Note that AeA_{e} depends not only on the isogeny class of GG, but on the isomorphism class of GG. For example, for ee regular, Ae=π0(ZG)A_{e}=\pi_{0}(ZG) where ZGZG is the center of GG. The action of AeA_{e} on H(e)\textup{H}^{*}({\mathcal{B}_{e}}) further factors through the image of AeA_{e} in the adjoint group GadG^{\textup{ad}}.

1.4.5. Purity

Springer [SpringerPure] proved that the cohomology of e\mathcal{B}_{e} is always pure (in the sense of Hodge theory when k=k=\mathbb{C}, or in the sense of Frobenius weights when kk is a finite field).

1.4.6.

Let e𝒩e\in\mathcal{N}. Consider the restriction map ie:H()H(e)i^{*}_{e}:\textup{H}^{*}({\mathcal{B}})\to\textup{H}^{*}({\mathcal{B}_{e}}) induced by the inclusion e\mathcal{B}_{e}\hookrightarrow\mathcal{B}. Then the image of iei^{*}_{e} is exactly the invariants of H(e)\textup{H}^{*}({\mathcal{B}_{e}}) under AeA_{e}. This is a theorem of Hotta and Springer [HottaSpringer, Theorem 1.1]. In particular, when GG is of type AA, such restriction maps are always surjective.

1.4.7. Parity vanishing

De Concini, Lusztig and Procesi [DLP] proved that Hi(e)\textup{H}^{i}({\mathcal{B}_{e}}) vanishes for all odd ii and any e𝒩e\in\mathcal{N}. When k=k=\mathbb{C}, they prove a stronger statement: Hi(e,)\textup{H}^{i}({\mathcal{B}_{e},\mathbb{Z}}) vanishes for odd ii and is torsion-free for even ii.

1.5. The Springer correspondence

Let WW be the Weyl group of GG. In 1976, Springer [Springer] made the fundamental observation that there is natural WW-action on H(e)\textup{H}^{*}({\mathcal{B}_{e}}), even though WW does not act on e\mathcal{B}_{e} as automorphisms of varieties.

3 Theorem (Springer [Springer, Thm 6.10]).
  1. (1)

    For each nilpotent element ee, there is a natural graded action of WW on H(e)\textup{H}^{*}({\mathcal{B}_{e}}) that commutes with the action of AeA_{e}.

  2. (2)

    For each nilpotent element ee and each irreducible representation ρ\rho of AeA_{e}, the multiplicity space M(e,ρ):=HomAe(ρ,H2de(e))M(e,\rho):=\textup{Hom}_{A_{e}}(\rho,\textup{H}^{2d_{e}}({\mathcal{B}_{e}})) is either zero or an irreducible representation of WW under the action in part (1).

  3. (3)

    Each irreducible representation χ\chi of WW appears as M(e,ρ)M(e,\rho) for a unique pair (e,ρ)(e,\rho) up to GG-conjugacy. The assignment χ(e,ρ)\chi\mapsto(e,\rho) thus gives an injection

    (4) Irr(W){(e,ρ)}/G.\textup{Irr}(W)\hookrightarrow\{(e,\rho)\}/G.

1.5.1. Convention

In fact there are two natural actions of WW on H(e)\textup{H}^{*}({\mathcal{B}_{e}}) that differ by tensoring with the sign representation of WW. In these notes we use the action that is normalized by the following properties. The trivial representation of WW corresponds to regular nilpotent ee and the trivial ρ\rho. The sign representation of WW corresponds to e=0e=0. Note however that Springer’s original paper [Springer] uses the other action.

1.5.2. The case e=0e=0

Taking e=0e=0, Springer’s theorem gives a graded action of WW on H()\textup{H}^{*}({\mathcal{B}}). What is this action? First, this action can be seen geometrically by considering G/TG/T instead of =G/B\mathcal{B}=G/B. In fact, since NG(T)/T=WN_{G}(T)/T=W, the right action of NG(T)N_{G}(T) on G/TG/T induces an action of WW on G/TG/T, which then induces an action of WW on H(G/T)\textup{H}^{*}({G/T}). Since the projection G/TG/BG/T\to G/B is an affine space bundle, it follows that H()H(G/T)\textup{H}^{*}({\mathcal{B}})\cong\textup{H}^{*}({G/T}). It can be shown that under this isomorphism, the action of WW on H(G/T)\textup{H}^{*}({G/T}) corresponds exactly to Springer’s action on H()\textup{H}^{*}({\mathcal{B}}).

Let S=Sym(𝕏(T))S=\textup{Sym}(\mathbb{X}^{*}(T)\otimes\mathbb{Q}_{\ell}) be the graded symmetric algebra where 𝕏(T)\mathbb{X}^{*}(T) has degree 22. The reflection representation of WW on 𝕏(T)\mathbb{X}^{*}(T) then induces a graded action of WW on SS. Recall Borel’s presentation of the cohomology ring of the flag variety

(5) H(,)S/(S+W)\textup{H}^{*}({\mathcal{B},\mathbb{Q}_{\ell}})\cong S/(S^{W}_{+})

where S+SS_{+}\subset S is the ideal spanned by elements of positive degree, and (S+W)(S^{W}_{+}) denotes the ideal of SS generated by WW-invariants on S+S_{+}. Then (5) is in fact an isomorphism of WW-modules (see [Springer, Prop 7.2]). By a theorem of Chevalley, S/(S+W)S/(S^{W}_{+}) is isomorphic to the regular representation of WW, therefore, as a WW-module, H()\textup{H}^{*}({\mathcal{B}}) is also isomorphic to the regular representation of WW.

6 Remark.

The target set in (4) can be canonically identified with the set of isomorphism classes of irreducible GG-equivariant local systems on nilpotent orbits. In fact, for an irreducible GG-equivariant local systems \mathcal{L} on a nilpotent orbit 𝒪𝒩\mathcal{O}\subset\mathcal{N}, its stalk at e𝒪e\in\mathcal{O} gives an irreducible representation ρ\rho of the centralizer CG(e)C_{G}(e) which factors through AeA_{e}. Note that the notion of GG-equivariance changes when GG varies in a fixed isogeny class. It is possible to extend the above injection (4) into a bijection by supplementing Irr(W)\textup{Irr}(W) with Irr(W)\textup{Irr}(W^{\prime}) for a collection of smaller Weyl groups. This is called the generalized Springer correspondence discovered by Lusztig [LuIC].

Springer’s original proof of Theorem 3 uses trigonometric sums over 𝔤(𝔽q)\mathfrak{g}(\mathbb{F}_{q}) and, when kk has characteristic zero, his proof uses reduction to finite fields. The following theorem due to Borho and MacPherson [BM] can be used to give a direct proof of the Springer correspondence for all base fields kk of large characteristics or characteristic zero. To state it, we need to use the language of constructible (complexes of) \mathbb{Q}_{\ell}-sheaves and perverse sheaves, for which we refer to the standard reference [BBD] and de Cataldo’s lectures [dC] in this volume.

7 Theorem.

The complex 𝒮:=𝐑π[dim𝒩]\mathcal{S}:=\mathbf{R}\pi_{*}\mathbb{Q}_{\ell}[\dim\mathcal{N}] is a perverse sheaf on 𝒩\mathcal{N} whose endomorphism ring is canonically isomorphic to the group algebra [W]\mathbb{Q}_{\ell}[W]. In particular, WW acts on the stalks of 𝐑π\mathbf{R}\pi_{*}\mathbb{Q}_{\ell}, i.e., WW acts on H(e)\textup{H}^{*}({\mathcal{B}_{e}}) for all e𝒩e\in\mathcal{N}.

We sketch three constructions of the WW-action on 𝐑π[dim𝒩]\mathbf{R}\pi_{*}\mathbb{Q}_{\ell}[\dim\mathcal{N}].

1.5.3. Construction via middle extension

This construction (or rather the version where 𝔤\mathfrak{g} is replaced by GG) is due to Lusztig [LuGreen, §3]. The dimension formula for Springer fibers (2) imply that

  • The map π:𝒩~𝒩\pi:\widetilde{\mathcal{N}}\to\mathcal{N} is semismall. 555A proper surjective map f:XYf:X\to Y of irreducible varieties is called semismall (resp. small) if for any d1d\geq 1, {yY|dimf1(y)d}\{y\in Y|\dim f^{-1}(y)\geq d\} has codimension at least 2d2d (resp. 2d+12d+1) in YY.

There is an extension of the dimension formula (2) for the dimension of ~X\widetilde{\mathcal{B}}_{X} valid for all elements X𝔤X\in\mathfrak{g}. Using this formula one can show that

  • The map π𝔤:𝔤~𝔤\pi_{\mathfrak{g}}:\widetilde{\mathfrak{g}}\to\mathfrak{g} is small.

As a well-known fact in the theory of perverse sheaves, the smallness of π𝔤\pi_{\mathfrak{g}} (together with the fact that 𝔤~\widetilde{\mathfrak{g}} is smooth) implies that 𝒮𝔤:=𝐑π𝔤,[dim𝔤]\mathcal{S}_{\mathfrak{g}}:=\mathbf{R}\pi_{\mathfrak{g},*}\mathbb{Q}_{\ell}[\dim\mathfrak{g}] is a perverse sheaf which is the middle extension of its restriction to any open dense subset of 𝔤\mathfrak{g}. Over the regular semisimple locus 𝔤rs𝔤\mathfrak{g}^{\textup{rs}}\subset\mathfrak{g}, π𝔤\pi_{\mathfrak{g}} is a WW-torsor, therefore 𝒮𝔤|𝔤rs\mathcal{S}_{\mathfrak{g}}|_{\mathfrak{g}^{rs}} is a local system shifted in degree dim𝔤-\dim\mathfrak{g} that admits an action of WW. By the functoriality of middle extension, 𝒮𝔤\mathcal{S}_{\mathfrak{g}} admits an action of WW. Taking stalks of 𝒮𝔤\mathcal{S}_{\mathfrak{g}}, we get an action of WW on H(~X)\textup{H}^{*}({\widetilde{\mathcal{B}}_{X}}) for all X𝔤X\in\mathfrak{g}.

In particular, for a nilpotent element ee, we get an action of WW on H(e)=H(~e)\textup{H}^{*}({\mathcal{B}_{e}})=\textup{H}^{*}({\widetilde{\mathcal{B}}_{e}}), because ~e\widetilde{\mathcal{B}}_{e} and e\mathcal{B}_{e} have the same reduced structure. This is the action defined by Springer in his original paper [Springer], which differs from our action by tensoring with the sign character of WW.

1.5.4. Construction via Fourier transform

By the semismallness of π\pi, the complex 𝒮=𝐑π[dim𝒩]\mathcal{S}=\mathbf{R}\pi_{*}\mathbb{Q}_{\ell}[\dim\mathcal{N}] is also a perverse sheaf. However, it is not the middle extension from an open subset of 𝒩\mathcal{N}. There is a notion of Fourier transform for 𝔾m\mathbb{G}_{m}-equivariant sheaves on affine spaces [Laumon]. One can show that 𝒮\mathcal{S} is isomorphic to the Fourier transform of 𝒮𝔤\mathcal{S}_{\mathfrak{g}} and vice versa. The WW-action on 𝒮𝔤\mathcal{S}_{\mathfrak{g}} then induces an action of WW on 𝒮\mathcal{S} by the functoriality of Fourier transform. Taking the stalk of 𝒮\mathcal{S} at ee we get an action of WW on H(e)\textup{H}^{*}({\mathcal{B}_{e}}). This action is normalized according to our convention in §1.5.1.

1.5.5. Construction via correspondences

Consider the Steinberg variety St𝔤=𝔤~×𝔤𝔤~\textup{St}_{\mathfrak{g}}=\widetilde{\mathfrak{g}}\times_{\mathfrak{g}}\widetilde{\mathfrak{g}} which classifies triples (X,B1,B2)(X,B_{1},B_{2}), where B1,B2B_{1},B_{2} are Borel subgroups of GG and XLieB1LieB2X\in\textup{Lie}\ B_{1}\cap\textup{Lie}\ B_{2}. The irreducible components of St𝔤\textup{St}_{\mathfrak{g}} are indexed by elements in the Weyl group: for wWw\in W, letting Stw\textup{St}_{w} be the closure of the graph of the ww-action on 𝔤~rs\widetilde{\mathfrak{g}}^{\textup{rs}}, then Stw\textup{St}_{w} is an irreducible component of St and these exhaust all irreducible components of St. The formalism of cohomological correspondences allows us to get an endomorphism of the complex 𝒮𝔤=𝐑π𝔤,[dim𝔤]\mathcal{S}_{\mathfrak{g}}=\mathbf{R}\pi_{\mathfrak{g},*}\mathbb{Q}_{\ell}[\dim\mathfrak{g}] from each Stw\textup{St}_{w}. It is nontrivial to show that these endomorphisms together form an action of WW on 𝒮𝔤\mathcal{S}_{\mathfrak{g}}. The key ingredient in the argument is still the smallness of the map π𝔤\pi_{\mathfrak{g}}. After the WW-action on 𝒮𝔤\mathcal{S}_{\mathfrak{g}} is defined, one then define the Springer action on H(e)\textup{H}^{*}({\mathcal{B}_{e}}) by either twisting the action of WW on the stalk 𝒮𝔤,e\mathcal{S}_{\mathfrak{g},e} by the sign representation as in §1.5.3, or by using Fourier transform as in §1.5.4. We refer to [GS, Remark 3.3.4] for some discussion of this construction. See also [CG, §3.4] for a similar but different construction using limits of Stw\textup{St}_{w} in the nilpotent Steinberg variety 𝒩~×𝒩𝒩~\widetilde{\mathcal{N}}\times_{\mathcal{N}}\widetilde{\mathcal{N}}.

Note that the above three constructions all allow one to show that End(𝒮)[W]\textup{End}(\mathcal{S})\cong\mathbb{Q}_{\ell}[W], hence giving a proof of Theorem 7.

1.5.6. Construction via monodromy

We sketch a construction of Slodowy [Slo4, §4] which works for k=k=\mathbb{C}. This construction was conjectured to give the same action of WW on H(e)\textup{H}^{*}({\mathcal{B}_{e}}) as the one in Theorem 3. A similar construction by Rossmann appeared in [Rossmann, §2], in which the author identified his action with that constructed by Kazhdan and Lusztig in [KL80], and the latter was known to be the same as Springer’s action. Thus all these constructions give the same WW-action as in Theorem 3.

Let e𝒩e\in\mathcal{N} and let Se𝔤S_{e}\subset\mathfrak{g} be a transversal slice to the orbit of ee. Upon choosing an 𝔰𝔩2\mathfrak{sl}_{2}-triple (e,h,f)(e,h,f) containing ee, there is a canonical choice of such a transversal slice Se=e+𝔤fS_{e}=e+\mathfrak{g}_{f}, where 𝔤f\mathfrak{g}_{f} is the centralizer of ff in 𝔤\mathfrak{g}. Now consider the following diagram where the squares are Cartesian except for the rightmost one

(8) e\textstyle{\mathcal{B}_{e}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}S~enil\textstyle{\widetilde{S}^{\textup{nil}}_{e}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}S~e\textstyle{\widetilde{S}_{e}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔤~\textstyle{\widetilde{\mathfrak{g}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π𝔤\scriptstyle{\pi_{\mathfrak{g}}}χ~\scriptstyle{\widetilde{\chi}}𝔱\textstyle{\mathfrak{t}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}{e}\textstyle{\{e\}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Senil\textstyle{S^{\textup{nil}}_{e}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Se\textstyle{S_{e}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔤\textstyle{\mathfrak{g}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}χ\scriptstyle{\chi}𝔠\textstyle{\mathfrak{c}}

Here the rightmost square is (1), Senil=Se𝒩S^{\textup{nil}}_{e}=S_{e}\cap\mathcal{N} and S~e\widetilde{S}_{e} and S~enil\widetilde{S}^{\textup{nil}}_{e} are the preimages of SeS_{e} and SenilS^{\textup{nil}}_{e} under π𝔤\pi_{\mathfrak{g}}. Let VeSeV_{e}\subset S_{e} be a small ball around ee and let V0𝔠V_{0}\subset\mathfrak{c} be an even smaller ball around 0𝔠0\in\mathfrak{c}. Let Ue=Veχ1(V0)SeU_{e}=V_{e}\cap\chi^{-1}(V_{0})\subset S_{e}. Then the diagram (8) restricts to a diagram

(9) e\textstyle{\mathcal{B}_{e}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}U~enil\textstyle{\widetilde{U}^{\textup{nil}}_{e}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}U~e\textstyle{\widetilde{U}_{e}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}χ~e\scriptstyle{\widetilde{\chi}_{e}}V~0\textstyle{\widetilde{V}_{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}{e}\textstyle{\{e\}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Uenil\textstyle{U^{\textup{nil}}_{e}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ue\textstyle{U_{e}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}V0\textstyle{V_{0}}

Here Uenil=Ue𝒩U^{\textup{nil}}_{e}=U_{e}\cap\mathcal{N}, and V~0,U~e\widetilde{V}_{0},\widetilde{U}_{e} and U~enil\widetilde{U}^{\textup{nil}}_{e} are the preimages of V0,UeV_{0},U_{e} and UenilU^{\textup{nil}}_{e} under the vertical maps in (8). The key topological facts here are

  • The inclusion eU~enil\mathcal{B}_{e}\hookrightarrow\widetilde{U}^{\textup{nil}}_{e} admits a deformation retract, hence it is a homotopy equivalence;

  • The map χ~e:U~eV~0\widetilde{\chi}_{e}:\widetilde{U}_{e}\to\widetilde{V}_{0} is a trivializable fiber bundle (in the sense of differential topology).

Now a general fiber of χ~e\widetilde{\chi}_{e} admits a homotopy action of WW by the second point above because the rightmost square in (9) is Cartesian over V0𝔠rsV_{0}\cap\mathfrak{c}^{\textup{rs}} and the map V~0V0\widetilde{V}_{0}\to V_{0} is a WW-torsor over V0𝔠rsV_{0}\cap\mathfrak{c}^{\textup{rs}}. By the first point above, e\mathcal{B}_{e} has the same homotopy type with U~enil=χ~e1(0)\widetilde{U}^{\textup{nil}}_{e}=\widetilde{\chi}^{-1}_{e}(0), hence e\mathcal{B}_{e} also has the same homotopy type as a general fiber of χ~e\widetilde{\chi}_{e} because χ~e\widetilde{\chi}_{e} is a fiber bundle. Combining these facts, we get an action of WW on the homotopy type of e\mathcal{B}_{e}, which is a stronger structure than an action of WW on the cohomology of e\mathcal{B}_{e}. A consequence of this construction is that the WW-action on H(e)\textup{H}^{*}({\mathcal{B}_{e}}) in Theorem 3 preserves the ring structure.

1.5.7. Proof of Theorem 3 assuming Theorem 7

We decompose the perverse sheaf 𝒮\mathcal{S} into isotypical components under the WW-action

𝒮=χIrr(W)Vχ𝒮χ\mathcal{S}=\bigoplus_{\chi\in\textup{Irr}(W)}V_{\chi}\otimes\mathcal{S}_{\chi}

where VχV_{\chi} is the space on which WW acts via the irreducible representation χ\chi, and 𝒮χ=HomW(Vχ,𝒮)\mathcal{S}_{\chi}=\textup{Hom}_{W}(V_{\chi},\mathcal{S}) is a perverse sheaf on 𝒩\mathcal{N}. Since End(𝒮)[W]\textup{End}(\mathcal{S})\cong\mathbb{Q}_{\ell}[W], we conclude that each 𝒮χ\mathcal{S}_{\chi} is nonzero and that

(10) Hom(𝒮χ,𝒮χ)={χ=χ0otherwise\textup{Hom}(\mathcal{S}_{\chi},\mathcal{S}_{\chi^{\prime}})=\begin{cases}\mathbb{Q}_{\ell}&\chi=\chi^{\prime}\\ 0&\textup{otherwise}\end{cases}

The decomposition theorem [BBD, Th 6.2.5] implies that each 𝒮χ\mathcal{S}_{\chi} is a semisimple perverse sheaf. Therefore (10) implies that 𝒮χ\mathcal{S}_{\chi} is simple. Hence 𝒮χ\mathcal{S}_{\chi} is of the form IC(𝒪¯,)\textup{IC}(\overline{\mathcal{O}},\mathcal{L}) where 𝒪𝒩\mathcal{O}\subset\mathcal{N} is a nilpotent orbit and \mathcal{L} is an irreducible GG-equivariant local system on 𝒪\mathcal{O}. Moreover, since Hom(𝒮χ,𝒮χ)=0\textup{Hom}(\mathcal{S}_{\chi},\mathcal{S}_{\chi^{\prime}})=0 for χχ\chi\neq\chi^{\prime}, the simple perverse sheaves {𝒮χ}χIrr(W)\{\mathcal{S}_{\chi}\}_{\chi\in\textup{Irr}(W)} are non-isomorphic to each other. This proves part (3) of Theorem 3 by interpreting the right side of (4) as the set of isomorphism classes of irreducible GG-equivariant local systems on nilpotent orbits. If 𝒮χ=IC(𝒪¯,)\mathcal{S}_{\chi}=\textup{IC}(\overline{\mathcal{O}},\mathcal{L}) and e𝒪e\in\mathcal{O}, the semismallness of π\pi allows us to identify the stalk e\mathcal{L}_{e} with an AeA_{e}-isotypic subspace of H2de(e)\textup{H}^{2d_{e}}({\mathcal{B}_{e}}). This proves part (2) of Theorem 3. ∎

We give some further examples of the Springer correspondence.

1.5.8. Type AA

When G=SLnG=\textup{SL}_{n}, all AeA_{e} are trivial. The Springer correspondence sets a bijection between irreducible representations of W=SnW=S_{n} and nilpotent orbits of 𝔤=𝔰𝔩n\mathfrak{g}=\mathfrak{sl}_{n}, both parametrized by partitions of nn. In §1.3.6 we have seen that if ee has Jordan type λ\lambda, the top dimensional cohomology H2de(e)\textup{H}^{2d_{e}}({\mathcal{B}_{e}}) has a basis indexed by the standard Young tableaux of λ\lambda, the latter also indexing a basis of the irreducible representation of SnS_{n} corresponding to the partition λ\lambda. For example, for G=SL3G=\textup{SL}_{3}, the Springer correspondences reads

  • trivial representation \leftrightarrow regular orbit, partition 3=33=3;

  • two-dimensional representation \leftrightarrow subregular orbit, partition 3=2+13=2+1;

  • sign representation \leftrightarrow {0}, partition 3=1+1+13=1+1+1.

1.5.9. The subregular orbit and the reflection representation

Consider the case ee is a subregular nilpotent element. In this case, the component group AeA_{e} can be identified with the automorphism group of the Dynkin diagram of GG^{\prime} introduced in §1.3.8 (see [Slodowy, §7.5, Proposition]). After identifying the irreducible components of e\mathcal{B}_{e} with the vertices of the Dynkin diagram of GG^{\prime}, the action of AeA_{e} on H2(e)\textup{H}^{2}({\mathcal{B}_{e}}) is by permuting the basis given by irreducible components in the same way as its action on the Dynkin diagram of GG^{\prime}. For example, when G=G2G=G_{2}, we may write e=C1C2C3C4\mathcal{B}_{e}=C_{1}\cup C_{2}\cup C_{3}\cup C_{4} with C1C_{1}, C2C_{2}, C3C_{3} each intersecting C4C_{4} in a point and otherwise disjoint. The group AeA_{e} is isomorphic to S3S_{3}, and its action on H2(e)\textup{H}^{2}({\mathcal{B}_{e}}) fixes the fundamental class of C4C_{4} and permutes the fundamental classes of C1,C2C_{1},C_{2} and C3C_{3}.

Note that H2(e)Ae\textup{H}^{2}({\mathcal{B}_{e}})^{A_{e}} always has dimension rr, the rank of GG. In fact, as a WW-module, H2(e)Ae\textup{H}^{2}({\mathcal{B}_{e}})^{A_{e}} is isomorphic to the reflection representation of WW on 𝔱\mathfrak{t}^{*}. In other words, under the Springer correspondence, the pair (e=subregular,ρ=1)(e=\textup{subregular},\rho=1) corresponds to the reflection representation of WW.

1.6. Comments and generalizations

1.6.1. Extended symmetry

The WW-action on H(e)\textup{H}^{*}({\mathcal{B}_{e}}) can be extended to an action of a larger algebra in various ways, if we use more sophisticated cohomology theories. On the equivariant cohomology HG~e(e)\textup{H}^{*}_{\widetilde{G}_{e}}(\mathcal{B}_{e}), there is an action of the graded affine Hecke algebra (see Lusztig [LuGrHk]). On the G~e\widetilde{G}_{e}-equivariant KK-group of e\mathcal{B}_{e}, there is an action of the affine Hecke algebra (see Kazhdan-Lusztig [KLAffHk] and Chriss-Ginzburg [CG]).

1.6.2. The group version

There are obvious analogs of the Springer resolution and the Grothendieck alteration when 𝒩\mathcal{N} and 𝔤\mathfrak{g} are replaced with the unipotent variety 𝒰G\mathcal{U}\subset G and GG itself. When char(k)\textup{char}(k) is large, the exponential map identifies 𝒩\mathcal{N} with 𝒰\mathcal{U} in a GG-equivariant manner, hence the theories of Springer fibers for nilpotent elements and unipotent elements are identical. The group version of the perverse sheaf 𝒮𝔤\mathcal{S}_{\mathfrak{g}} and its irreducible direct summands are precursors of character sheaves, a theory developed by Lusztig ([LuChShI], [LuChShII], [LuChShIV] and [LuChShV]) to study characters of the finite groups G(𝔽q)G(\mathbb{F}_{q}).

1.6.3. Partial Springer resolutions

We may define analogs of e\mathcal{B}_{e} in partial flag varieties. Let 𝒫\mathcal{P} be a partial flag variety of GG classifying parabolic subgroups PP of GG of a given type. There are two analogs of the map π:𝒩~𝒩\pi:\widetilde{\mathcal{N}}\to\mathcal{N} one may consider.

First, instead of considering 𝒩~=T\widetilde{\mathcal{N}}=T^{*}\mathcal{B}, we may consider T𝒫T^{*}\mathcal{P}, which classifies pairs (e,P)𝒩×𝒫(e,P)\in\mathcal{N}\times\mathcal{P} such that eLie𝔫Pe\in\textup{Lie}\ \mathfrak{n}_{P}, where 𝔫P\mathfrak{n}_{P} is the nilpotent radical of LieP\textup{Lie}\ P. Let τ𝒫:T𝒫𝒩\tau_{\mathcal{P}}:T^{*}\mathcal{P}\to\mathcal{N} be the first projection. In general this map is not surjective, its image is the closure of a nilpotent orbit 𝒪𝒫\mathcal{O}_{\mathcal{P}}. The orbit 𝒪P\mathcal{O}_{P} is characterized by the property that its intersection with 𝔫P\mathfrak{n}_{P} is dense in 𝔫P\mathfrak{n}_{P}, for any P𝒫P\in\mathcal{P}. This is called the Richardson class associated to parabolic subgroups of type 𝒫\mathcal{P}. When GG is of type AA, each nilpotent class 𝒪\mathcal{O} is the Richardson class associated to parabolic subgroups of some type 𝒫\mathcal{P} (not unique in general). The map T𝒫𝒪¯T^{*}\mathcal{P}\to\overline{\mathcal{O}} is a resolution of singularities. For general GG, not every nilpotent orbit is Richardson.

Second, we may consider the subscheme 𝒩~𝒫𝒩×𝒫\widetilde{\mathcal{N}}_{\mathcal{P}}\subset\mathcal{N}\times\mathcal{P} classifying pairs (e,P)(e,P) such that e𝒩Pe\in\mathcal{N}_{P}, where 𝒩PLieP\mathcal{N}_{P}\subset\textup{Lie}\ P is the nilpotent cone of PP. The projection π𝒫:𝒩~𝒫𝒩\pi_{\mathcal{P}}:\widetilde{\mathcal{N}}_{\mathcal{P}}\to\mathcal{N} is now surjective, and is a partial resolution of singularities. The Springer resolution π\pi can be factored as

π:𝒩~=𝒩~ν𝒫𝒩~𝒫π𝒫𝒩.\pi:\widetilde{\mathcal{N}}=\widetilde{\mathcal{N}}_{\mathcal{B}}\xrightarrow{\nu_{\mathcal{P}}}\widetilde{\mathcal{N}}_{\mathcal{P}}\xrightarrow{\pi_{\mathcal{P}}}\mathcal{N}.

We have an embedding T𝒫𝒩~𝒫T^{*}\mathcal{P}\hookrightarrow\widetilde{\mathcal{N}}_{\mathcal{P}}. We may consider the fibers of either τ𝒫\tau_{\mathcal{P}} or π𝒫\pi_{\mathcal{P}} as parabolic analogs of Springer fibers. We call them partial Springer fibers. The Springer action of WW on the cohomology of e\mathcal{B}_{e} has an analog for partial Springer fibers. For more information, we refer the readers to [BM].

1.6.4. Hessenberg varieties

The Grothendieck alteration π𝔤:𝔤~𝔤\pi_{\mathfrak{g}}:\widetilde{\mathfrak{g}}\to\mathfrak{g} admits a generalization where 𝔤\mathfrak{g} is replaced with another linear representation of GG.

Fix a Borel subgroup BB of GG. Let (V,ρ)(V,\rho) be a representation of GG and V+VV^{+}\subset V be a subspace which is stable under BB. Now we use the pair (V,V+)(V,V^{+}) instead of the pair (𝔤,𝔟)(\mathfrak{g},\mathfrak{b}), we get a generalization of the Grothendieck alteration. More precisely, let V~V×\widetilde{V}\subset V\times\mathcal{B} be the subscheme consisting of pairs (v,gB)V×(v,gB)\in V\times\mathcal{B} such that vρ(g)V+v\in\rho(g)V^{+}. Let πV:V~V\pi_{V}:\widetilde{V}\to V be the first projection. The fibers of πV\pi_{V} are called Hessenberg varieties.

Hessenberg varieties appear naturally in the study of certain affine Springer fibers, as we will see in §2.4. For more information on Hessenberg varieties, see [GKM] and [OY].

1.7. Exercises

1.7.1.

For G=SLnG=\textup{SL}_{n}, determine the sizes of the Jordan blocks of a regular and subregular nilpotent element of 𝔤\mathfrak{g}.

1.7.2.

For G=SL2G=\textup{SL}_{2} and SL3\textup{SL}_{3}, calculate the coordinate ring of the non-reduced Springer fiber ~e\widetilde{\mathcal{B}}_{e} for a regular nilpotent element ee. Show also that the Springer fiber e\mathcal{B}_{e} is indeed a reduced point.

Hint: if you write ee as an upper triangular matrix, then ~e\widetilde{\mathcal{B}}_{e} lies in the big Bruhat cell of the flag variety \mathcal{B}, from which you get coordinates for your calculation.

1.7.3.

Verify the statement in §1.3.7: consider G=Sp(V)G=\textup{Sp}(V), V=v1,v2,v3,v4V=\langle{v_{1},v_{2},v_{3},v_{4}}\rangle with the symplectic form given by vi,v5i=1\langle{v_{i},v_{5-i}}\rangle=1 if i=1,2i=1,2 and vi,vj=0\langle{v_{i},v_{j}}\rangle=0 for i+j5i+j\neq 5. Let e:v4v10,v20,v30e:v_{4}\mapsto v_{1}\to 0,v_{2}\mapsto 0,v_{3}\mapsto 0. Then any flag 0V1V2V1V0\subset V_{1}\subset V_{2}\subset V_{1}^{\bot}\subset V in e\mathcal{B}_{e} must satisfy

v1V2v1,v2,v3.\langle{v_{1}}\rangle\subset V_{2}\subset\langle{v_{1},v_{2},v_{3}}\rangle.

Moreover, this is the only condition for a flag 0V1V2V1V0\subset V_{1}\subset V_{2}\subset V_{1}^{\bot}\subset V to lie in e\mathcal{B}_{e}.

1.7.4.

Let e𝒩e\in\mathcal{N}. Let BGB\subset G be a Borel subgroup.

  1. (1)

    Let α\alpha be a simple root. Let PαBP_{\alpha}\supset B be a parabolic subgroup whose Levi factor has semisimple rank one with roots ±α\pm\alpha. Let 𝒫αG/Pα\mathcal{P}_{\alpha}\cong G/P_{\alpha} be the partial flag variety of GG classifying parabolics conjugate to PαP_{\alpha}. Restricting the projection 𝒫α\mathcal{B}\to\mathcal{P}_{\alpha} to e\mathcal{B}_{e}, we get a map

    πα:eπα(e).\pi_{\alpha}:\mathcal{B}_{e}\to\pi_{\alpha}(\mathcal{B}_{e}).

    Show that the pullback πα\pi^{*}_{\alpha} on cohomology induces an isomorphism

    (11) H(πα(e))H(e)sα\textup{H}^{*}({\pi_{\alpha}(\mathcal{B}_{e})})\cong\textup{H}^{*}({\mathcal{B}_{e}})^{s_{\alpha}}

    where sαWs_{\alpha}\in W is the simple reflection associated with α\alpha, which acts on H(e)\textup{H}^{*}({\mathcal{B}_{e}}) via Springer’s action.

  2. (2)

    Can you generalize (11) to other partial flag varieties?

1.7.5.

Let G=SL3G=\textup{SL}_{3} and let e𝒩e\in\mathcal{N} be a subregular element. Calculate the action of S3S_{3} on the two dimensional H2(e)\textup{H}^{2}({\mathcal{B}_{e}}) in terms of the basis given by the fundamental classes of the two irreducible components (see §1.3.4).

Hint: for this and the next problem, you may find Exercise 1.7.4 useful.

1.7.6.

Describe the Springer fibers for Sp4\textup{Sp}_{4}. Calculate the Springer correspondence for G=Sp4G=\textup{Sp}_{4} explicitly.

1.7.7.

Using the dimension formula for e\mathcal{B}_{e}, verify that the Springer resolution π\pi is semismall.

1.7.8.

Let X𝔤X\in\mathfrak{g} and let X=Xs+XnX=X_{s}+X_{n} be the Jordan decomposition of XX. Let H=CG(Xs)GH=C_{G}(X_{s})\subset G. This is a Levi subgroup of GG. Let H\mathcal{B}^{H} be the flag variety of HH and let XnH\mathcal{B}^{H}_{X_{n}} be the Springer fiber of XnX_{n} viewed as a nilpotent element in LieH\textup{Lie}\ H. Show that dim~X=dimXnH\dim\widetilde{\mathcal{B}}_{X}=\dim\mathcal{B}^{H}_{X_{n}}.

1.7.9.

Use Exercise 1.7.8 and the dimension formula (2) to derive a formula for the dimension of ~X\widetilde{\mathcal{B}}_{X} for all elements X𝔤X\in\mathfrak{g}. Use your formula to prove that the Grothendieck alteration π𝔤\pi_{\mathfrak{g}} is small.

1.7.10.

Show that 𝒩\mathcal{N} is rationally smooth; i.e., its intersection cohomology complex is isomorphic to the shifted constant sheaf [dim𝒩]\mathbb{Q}_{\ell}[\dim\mathcal{N}].

Hint: the largest direct summand in 𝐑π\mathbf{R}\pi_{*}\mathbb{Q}_{\ell} is the shifted IC sheaf of 𝒩\mathcal{N}, and it is also the restriction of a direct summand of 𝐑π𝔤,\mathbf{R}\pi_{\mathfrak{g},*}\mathbb{Q}_{\ell}.

1.7.11.

Show that the Springer fibers e\mathcal{B}_{e} are connected.

Hint: the H0\textup{H}^{0} of the Springer fibers are packed in some sheaf.

1.7.12.

Denote the simple roots of GG by {α1,,αr}\{\alpha_{1},\cdots,\alpha_{r}\}. A parabolic subgroup PGP\subset G is called of type ii if the roots of its Levi quotient LPL_{P} are ±αi\pm\alpha_{i}.

  1. (1)

    Let 1ir1\leq i\leq r and let PP be a parabolic subgroup of type ii. Let 𝔫P\mathfrak{n}_{P} be the nilpotent radical of LieP\textup{Lie}\ P. Show that 𝔫P𝒪subreg\mathfrak{n}_{P}\cap\mathcal{O}_{\operatorname{subreg}} is dense in 𝔫P\mathfrak{n}_{P}.

  2. (2)

    Let e𝒪subrege\in\mathcal{O}_{\operatorname{subreg}}. Show that for each ii, there are finitely many parabolics PP of type ii such that e𝔫Pe\in\mathfrak{n}_{P}. For each such PP, the subvariety CP:={Be|BP}C_{P}:=\{B\in\mathcal{B}_{e}|B\subset P\} of e\mathcal{B}_{e} is isomorphic to 1\mathbb{P}^{1}, and is called a curve of type ii.

  3. (3)

    For parabolics PQP\neq Q of type ii and jj, show that CPCQC_{P}\cap C_{Q} is either empty or a point.

  4. (4)

    Let PP be a parabolic subgroup of type ii such that e𝔫Pe\in\mathfrak{n}_{P}. For any 1jr,ij1\leq j\leq r,i\neq j, CPC_{P} intersects exactly αi,αj-\langle{\alpha^{\vee}_{i},\alpha_{j}}\rangle curves of type jj.

  5. (5)

    Show that e\mathcal{B}_{e} has the configuration described in §1.3.8.

  6. (6)

    Use Exercise 1.7.4 to calculate the Springer action of WW on H2(e)\textup{H}^{2}({\mathcal{B}_{e}}).

2. Lecture II: Affine Springer fibers

Affine Springer fibers are analogs of Springer fibers for loop groups. They were introduced by Kazhdan and Lusztig [KL]. Roughly speaking, in the case of classical groups, instead of classifying flags in a kk-vector space fixed by a kk-linear transformation, affine Springer fibers classify (chains of) lattices in an FF-vector space fixed by an FF-linear transformation, where F=k((t))F=k(\!(t)\!). The cohomology groups of affine Springer fibers carry actions of the affine Weyl group.

The setup in this section is the same as in §1.1.

2.1. Loop group, parahoric subgroups and the affine flag variety

Let F=k((t))F=k(\!(t)\!) be the field of formal Laurent series in one variable tt. Then FF has a discrete valuation valF:F×\textup{val}_{F}:F^{\times}\to\mathbb{Z} such that valF(t)=1\textup{val}_{F}(t)=1 and its valuation ring is 𝒪F=k[[t]]\mathcal{O}_{F}=k[\![t]\!].

2.1.1. The loop group

The positive loop group L+GL^{+}G is a group-valued functor on kk-algebras. For any kk-algebra RR, we define LG(R):=G(R[[t]])LG(R):=G(R[\![t]\!]). It turns out that LGLG is representable by a scheme over kk which is not of finite type.

For example, when G=GLnG=\textup{GL}_{n}, an element in LG(R)=GLn(R[[t]])LG(R)=\textup{GL}_{n}(R[\![t]\!]) is given by n2n^{2} formal Laurent series aij=s0aij(s)tsa_{ij}=\sum_{s\geq 0}a^{(s)}_{ij}t^{s} (1i,jn1\leq i,j\leq n), with aij(s)Ra^{(s)}_{ij}\in R, subject to one condition that det((aij))\det((a_{ij})) (which is a polynomial in the aij(s)a^{(s)}_{ij} of degree nn) is invertible in RR. Therefore in this case LGLG is an open subscheme in the infinite dimensional affine space with coordinates aij(s)a^{(s)}_{ij}, 1i,jn1\leq i,j\leq n and s0s\geq 0.

Similarly we may define the loop group LGLG to be the functor LG(R)=G(R((t)))LG(R)=G(R(\!(t)\!)) on kk-algebras RR. The functor LGLG is no longer representable by a scheme, but rather by an ind-scheme. An ind-scheme is an inductive limit limmXm\varinjlim_{m}X_{m} in the category of schemes, i.e., {Xm}\{X_{m}\} form an inductive system of schemes over kk, and limmXm\varinjlim_{m}X_{m} is the functor RlimmXm(R)R\mapsto\varinjlim_{m}X_{m}(R). When G=GLnG=\textup{GL}_{n}, we may define XmX_{m} to be the subfunctor of LGLG such that Xm(R)X_{m}(R) consists of nn-by-nn invertible matrices with entries in tmR[[t]]R((t))t^{-m}R[\![t]\!]\subset R(\!(t)\!). Then the same argument as in the L+GL^{+}G case shows that XmX_{m} is representable by a scheme over kk. For m<mm<m^{\prime}, we have a natural closed embedding XmXmX_{m}\hookrightarrow X_{m^{\prime}}, and LGLG in this case is the inductive limit limmXm\varinjlim_{m}X_{m}. For general GG, see [BL, §1] and [Faltings, §2].

2.1.2. The affine Grassmannian

The affine Grassmannian GrG\textup{Gr}_{G} of GG is defined as the sheafification of the presheaf RLG(R)/L+G(R)R\mapsto LG(R)/L^{+}G(R) in the category of kk-algebras under the fpqc topology. In particular, we have GrG(k)=G(F)/G(𝒪F)\textup{Gr}_{G}(k)=G(F)/G(\mathcal{O}_{F}).

When G=GLnG=\textup{GL}_{n}, the affine Grassmannian GrG\textup{Gr}_{G} can be identified with the moduli space of projective R[[t]]R[\![t]\!]-submodules ΛR((t))n\Lambda\subset R(\!(t)\!)^{n} such that

(12) (tmR[[t]])nΛ(tmR[[t]])n(t^{m}R[\![t]\!])^{n}\subset\Lambda\subset(t^{-m}R[\![t]\!])^{n}

for some m0m\geq 0. Such an R[[t]]R[\![t]\!]-module Λ\Lambda is called a lattice in R((t))nR(\!(t)\!)^{n}. For fixed mm, let XmX_{m} be the subfunctor of GrG\textup{Gr}_{G} classifying those Λ\Lambda such that (12) holds, then XmX_{m} is representable by a projective scheme over kk. The natural closed embeddings XmXm+1X_{m}\hookrightarrow X_{m+1} make {Xm}\{X_{m}\} into an inductive system of projective schemes, and GrG\textup{Gr}_{G} is representable by the ind-scheme limmXm\varinjlim_{m}X_{m}.

Let us elaborate on the bijection between GrG(k)\textup{Gr}_{G}(k) and the lattices in Fn=k((t))nF^{n}=k(\!(t)\!)^{n}. Let 𝒪FnFn\mathcal{O}_{F}^{n}\subset F^{n} be the standard lattice. Let 𝔏n\mathfrak{L}_{n} be the set of lattices in FnF^{n} (in the case R=kR=k a lattice is simply an 𝒪F\mathcal{O}_{F}-submodules of FnF^{n} of rank nn). The group LG(k)=G(F)LG(k)=G(F) acts on 𝔏n\mathfrak{L}_{n} by LGg:ΛgΛLG\ni g:\Lambda\mapsto g\Lambda. This action is transitive and the stabilizer of the standard lattice 𝒪Fn\mathcal{O}^{n}_{F} is L+G(k)=G(𝒪F)L^{+}G(k)=G(\mathcal{O}_{F}). Therefore this action induces a G(F)G(F)-equivariant bijection

(13) GrG(k)=G(F)/G(𝒪F)𝔏n.\textup{Gr}_{G}(k)=G(F)/G(\mathcal{O}_{F})\stackrel{{\scriptstyle\sim}}{{\to}}\mathfrak{L}_{n}.

For general GG, GrG\textup{Gr}_{G} is always representable by an ind-scheme of the form limmXm\varinjlim_{m}X_{m} where XmX_{m} are projective schemes over kk, and the transition maps XmXm+1X_{m}\hookrightarrow X_{m+1} are closed embeddings. We have a canonical exhaustion of GrG\textup{Gr}_{G} by projective schemes given by the affine Schubert stratification, which we now recall. The action of L+GL^{+}G on GrG\textup{Gr}_{G} by left translation has orbits indexed by dominant cocharacters λ𝕏(T)+\lambda\in\mathbb{X}_{*}(T)^{+}. We denote by GrG,λ\textup{Gr}_{G,\lambda} the L+GL^{+}G-orbit through λ(t)\lambda(t). Let GrG,λ\textup{Gr}_{G,\leq\lambda} be the closure of GrG,λ\textup{Gr}_{G,\lambda}. Then GrG,λ\textup{Gr}_{G,\leq\lambda} is a projective scheme and GrG\textup{Gr}_{G} is the union of GrG,λ\textup{Gr}_{G,\leq\lambda}. For more details on the affine Grassmannian, we refer to [BL, §2], [Faltings, §2] and Zhu’s lectures [Zhu].

2.1.3. Parahoric subgroups

The subgroup L+GL^{+}G of LGLG is an example of a class of subgroups called parahoric subgroups. Fix a Borel subgroup BGB\subset G and let 𝐈L+G\mathbf{I}\subset L^{+}G be the preimage of BB under the map L+GGL^{+}G\to G given by reduction modulo tt. Then 𝐈\mathbf{I} is an example of an Iwahori subgroup of LGLG. General Iwahori subgroups are conjugates of 𝐈\mathbf{I} in LGLG. Like L+GL^{+}G, Iwahori subgroups are group subschemes of LGLG of infinite type. Parahoric subgroups are connected group subschemes of LGLG containing an Iwahori subgroup with finite codimension. A precise definition of parahoric subgroups involves a fair amount of Bruhat-Tits theory, which we refer the readers to the original papers of Bruhat and Tits [BT1], and the survey paper [Tits].

Just as the conjugacy classes of parabolic subgroups of GG are in bijection with subsets of the Dynkin diagram of GG, the LGLG-conjugacy classes of parahoric subgroups of LGLG are in bijection with proper subsets of the vertices of the extended Dynkin diagram Dyn~(G)\widetilde{\operatorname{Dyn}}(G) of GG, which has one more vertex than the Dynkin diagram of GG. See Kac’s book [Kac, §4.8], Bourbaki [Bourbaki, Ch VI] for extended Dynkin diagrams and the expository paper of Gross [Gross] for connection with parahoric subgroups.

Each 𝐏\mathbf{P} admits a canonical exact sequence of group schemes

1𝐏+𝐏L𝐏11\to\mathbf{P}^{+}\to\mathbf{P}\to L_{\mathbf{P}}\to 1

where 𝐏+\mathbf{P}^{+} is the pro-unipotent radical of 𝐏\mathbf{P} and L𝐏L_{\mathbf{P}} is a reductive group over kk, called the Levi quotient of 𝐏\mathbf{P}. If 𝐏\mathbf{P} corresponds to a subset JJ of the vertices of Dyn~(G)\widetilde{\operatorname{Dyn}}(G), then the Dynkin diagram of the Levi quotient L𝐏L_{\mathbf{P}} is the sub-diagram of Dyn~(G)\widetilde{\operatorname{Dyn}}(G) spanned by JJ.

2.1.4. Affine flag varieties

For each parahoric subgroup 𝐏LG\mathbf{P}\subset LG we may define the corresponding affine partial flag variety Fl𝐏\textup{Fl}_{\mathbf{P}} as the fpqc sheafification of the functor RLG(R)/𝐏(R)R\mapsto LG(R)/\mathbf{P}(R) on the category of kk-algebras. This functor is also representable by an ind-scheme limmXm\varinjlim_{m}X_{m} where each XmX_{m} is a projective scheme over kk and the transition maps are closed embeddings. The affine Grassmannian GrG\textup{Gr}_{G} is a special case of Fl𝐏\textup{Fl}_{\mathbf{P}} for 𝐏=L+G\mathbf{P}=L^{+}G.

Consider the special case 𝐏=𝐈\mathbf{P}=\mathbf{I} is an Iwahori subgroup of LGLG. When GG is simply-connected, 𝐈\mathbf{I} is its own normalizer, and we may identify Fl𝐈\textup{Fl}_{\mathbf{I}} as the moduli space of Iwahori subgroups of LGLG, hence giving an intrinsic definition of the affine flag variety. We usually denote Fl𝐈\textup{Fl}_{\mathbf{I}} by Fl or FlG\textup{Fl}_{G} and call it the affine flag variety of GG, with the caveat that Fl𝐈\textup{Fl}_{\mathbf{I}} is canonically independent of the choice of 𝐈\mathbf{I} only when GG is simply-connected.

Let 𝐏𝐐\mathbf{P}\subset\mathbf{Q} be two parahoric subgroups of LGLG. Then we have a natural projection Fl𝐏Fl𝐐\textup{Fl}_{\mathbf{P}}\to\textup{Fl}_{\mathbf{Q}}. The fibers of this projection are isomorphic to the partial flag variety of L𝐐L_{\mathbf{Q}} corresponding to its parabolic subgroup given by the image of 𝐏L𝐐\mathbf{P}\to L_{\mathbf{Q}}. In particular, there is a natural projection FlGGrG\textup{Fl}_{G}\to\textup{Gr}_{G} whose fibers are isomorphic to the flag variety \mathcal{B}.

2.1.5. The case of SLn\textup{SL}_{n}

We have seen in §2.1.2 that the affine Grassmannian of GLn\textup{GL}_{n} has an interpretation as the moduli space of lattices. In fact, parahoric subgroups of LGLG and the associated affine partial flag varieties can also be described using lattices. Here we consider the case G=SLnG=\textup{SL}_{n}.

Recall that the set of lattices in FnF^{n} is denoted by 𝔏n\mathfrak{L}_{n}. For any two lattices Λ1,Λ2𝔏n\Lambda_{1},\Lambda_{2}\in\mathfrak{L}_{n} we may define their relative length to be the integer

[Λ1:Λ2]:=dimk(Λ1/Λ1Λ2)dimk(Λ2/Λ1Λ2).[\Lambda_{1}:\Lambda_{2}]:=\dim_{k}(\Lambda_{1}/\Lambda_{1}\cap\Lambda_{2})-\dim_{k}(\Lambda_{2}/\Lambda_{1}\cap\Lambda_{2}).

Let J/nJ\subset\mathbb{Z}/n\mathbb{Z} be a non-empty subset. Let J~\widetilde{J} be the preimage of JJ under the projection /n\mathbb{Z}\to\mathbb{Z}/n\mathbb{Z}. A periodic JJ-chain of lattices is a function Λ:J~𝔏n\Lambda:\widetilde{J}\to\mathfrak{L}_{n} sending each iJ~i\in\widetilde{J} to a lattice Λi𝔏n\Lambda_{i}\in\mathfrak{L}_{n} such that

  • [Λi:𝒪Fn]=i[\Lambda_{i}:\mathcal{O}_{F}^{n}]=i for all iJ~i\in\widetilde{J};

  • ΛiΛj\Lambda_{i}\subset\Lambda_{j} for i<ji<j in J~\widetilde{J};

  • Λi=tΛi+n\Lambda_{i}=t\Lambda_{i+n} for all iJ~i\in\widetilde{J}.

Let 𝔏J\mathfrak{L}_{J} be the set of periodic JJ-chains of lattices. For each {Λi}iJ~𝔏J\{\Lambda_{i}\}_{i\in\widetilde{J}}\in\mathfrak{L}_{J}, let 𝐏{Λi}iJLG\mathbf{P}_{\{\Lambda_{i}\}_{i\in J}}\subset LG to be the simultaneous stabilizers of all Λi\Lambda_{i}’s. Then 𝐏{Λi}iJ\mathbf{P}_{\{\Lambda_{i}\}_{i\in J}} is a parahoric subgroup of LGLG. We call such a parahoric subgroup of type JJ. In fact all parahoric subgroups of LGLG arise from a unique periodic JJ-chain of lattices, for a unique non-empty J/nJ\subset\mathbb{Z}/n\mathbb{Z}. Therefore we get a bijection between J𝔏J\sqcup_{J}\mathfrak{L}_{J} and the set of parahoric subgroups of LGLG. In particular, L+GL^{+}G is the parahoric subgroup corresponding to the periodic {0}\{0\}-chain of lattices given by Λi=ti/n𝒪Fn\Lambda_{i}=t^{i/n}\mathcal{O}^{n}_{F}, where iJ~=ni\in\widetilde{J}=n\mathbb{Z}.

The extended Dynkin diagram of GG is a loop with nn nodes which we index cyclically by the set /n\mathbb{Z}/n\mathbb{Z}, such that 0 corresponds to the extra node compared to the usual Dynkin diagram. Parahoric subgroups of type JJ\neq\varnothing corresponds to the proper subset /nJ\mathbb{Z}/n\mathbb{Z}-J of the nodes of the extended Dynkin diagram.

One can find the moduli space FlJ\textup{Fl}_{J} of periodic JJ-chains of lattices such that FlJ(k)=𝔏J\textup{Fl}_{J}(k)=\mathfrak{L}_{J}. Fixing any parahoric subgroup 𝐏\mathbf{P} of type JJ, FlJ\textup{Fl}_{J} can be identified with the affine partial flag variety Fl𝐏\textup{Fl}_{\mathbf{P}}. In particular, the affine flag variety Fl for G=SLnG=\textup{SL}_{n} can be identified with the moduli space of periodic full chains of lattices, i.e., a sequence of lattices Λ1Λ0Λ1\cdots\Lambda_{-1}\subset\Lambda_{0}\subset\Lambda_{1}\cdots in FnF^{n} with [Λi:𝒪Fn]=i[\Lambda_{i}:\mathcal{O}^{n}_{F}]=i and Λi=tΛi+n\Lambda_{i}=t\Lambda_{i+n} for all ii\in\mathbb{Z}.

2.1.6. The case of Sp2n\textup{Sp}_{2n}

Now consider G=Sp2n=Sp(V)G=\textup{Sp}_{2n}=\textup{Sp}(V), where V=k2nV=k^{2n} is equipped with a symplectic form. We extended the symplectic form on VV FF-linearly to a symplectic form ,\langle{-,-}\rangle on VkFV\otimes_{k}F. For a lattice Λ𝔏2n\Lambda\in\mathfrak{L}_{2n}, define its symplectic dual to be the set Λ:={vVkF|v,Λ𝒪F}\Lambda^{\vee}:=\{v\in V\otimes_{k}F|\langle{v,\Lambda}\rangle\subset\mathcal{O}_{F}\}. Then Λ\Lambda^{\vee} is again a lattice in VkFV\otimes_{k}F. The operation ΛΛ\Lambda\mapsto\Lambda^{\vee} defines an involution on 𝔏2n\mathfrak{L}_{2n}.

Let J/2nJ\subset\mathbb{Z}/2n\mathbb{Z} be a non-empty subset stable under multiplication by 1-1. Let J~\widetilde{J}\subset\mathbb{Z} be the preimage of JJ under the natural projection /2n\mathbb{Z}\to\mathbb{Z}/2n\mathbb{Z}. A periodic self-dual JJ-chain of lattices in VkFV\otimes_{k}F is a periodic JJ-chain of lattices (i.e., an element in 𝔏J\mathfrak{L}_{J} in the notation of §2.1.5) satisfying the extra condition that

Λi=Λi, for all iJ~.\Lambda^{\vee}_{i}=\Lambda_{-i},\textup{ for all }i\in\widetilde{J}.

Denote the set of periodic self-dual JJ-chains of lattices in VkFV\otimes_{k}F by 𝔏JSp(V)\mathfrak{L}^{\textup{Sp}(V)}_{J}. This is a set with an action of G(F)=Sp(VkF)G(F)=\textup{Sp}(V\otimes_{k}F). For any {Λi}iJ~𝔏JSp(V)\{\Lambda_{i}\}_{i\in\widetilde{J}}\in\mathfrak{L}^{\textup{Sp}(V)}_{J}, the simultaneous stabilizer of the Λi\Lambda_{i}’s is a parahoric subgroup of LGLG, and every parahoric subgroup of LGLG arises this way. For a parahoric subgroup 𝐏\mathbf{P} of type JJ, the corresponding affine partial flag variety Fl𝐏\textup{Fl}_{\mathbf{P}} can be identified with the moduli space of periodic self-dual JJ-chains of lattices so that Fl𝐏(k)𝔏JSp(V)\textup{Fl}_{\mathbf{P}}(k)\cong\mathfrak{L}^{\textup{Sp}(V)}_{J} as G(F)G(F)-sets. The readers are invited to work out the similar story for orthogonal groups, see Exercise 2.8.1.

2.2. Affine Springer fibers

2.2.1. Affine Springer fibers in the affine Grassmannian

For any kk-algebra RR, we denote 𝔤kR\mathfrak{g}\otimes_{k}R by 𝔤(R)\mathfrak{g}(R). In particular, 𝔤(F)=𝔤kF\mathfrak{g}(F)=\mathfrak{g}\otimes_{k}F is the Lie algebra of the loop group LGLG. For gLGg\in LG, let Ad(g)\textup{Ad}(g) denote its adjoint action on 𝔤(F)\mathfrak{g}(F).

Let γ𝔤(F):=𝔤kF\gamma\in\mathfrak{g}(F):=\mathfrak{g}\otimes_{k}F be a regular semisimple element 666Here we are dealing with a Lie algebra 𝔤\mathfrak{g} over the non-algebraically-closed field FF. An element γ𝔤(F)\gamma\in\mathfrak{g}(F) is regular semisimple if it is regular semisimple as an element in 𝔤(F¯)\mathfrak{g}(\overline{F}), see the footnote in §1.2.2. Equivalently, γ\gamma is regular semisimple if its image in 𝔠(F)\mathfrak{c}(F) lies in 𝔠rs(F)\mathfrak{c}^{\textup{rs}}(F).. We consider the subfunctor of GrG\textup{Gr}_{G} whose value on a kk-algebra RR is given by

(14) 𝒳~γ(R)={[g]GrG(R)|Ad(g1)γ𝔤(R[[t]])}.\widetilde{\mathcal{X}}_{\gamma}(R)=\{[g]\in\textup{Gr}_{G}(R)|\textup{Ad}(g^{-1})\gamma\in\mathfrak{g}(R[\![t]\!])\}.

Then 𝒳~γ\widetilde{\mathcal{X}}_{\gamma} is a closed sub-ind-scheme of GrG\textup{Gr}_{G}. Let 𝒳γ=𝒳~γred\mathcal{X}_{\gamma}=\widetilde{\mathcal{X}}^{\textup{red}}_{\gamma} be the underlying reduced ind-scheme of 𝒳~γ\widetilde{\mathcal{X}}_{\gamma}. We call 𝒳γ\mathcal{X}_{\gamma} the affine Springer fiber of γ\gamma in the affine Grassmannian GrG\textup{Gr}_{G}.

2.2.2. Alternative definition in terms of lattices

We consider the case G=GLnG=\textup{GL}_{n}. Let γ𝔤(F)=𝔤𝔩n(F)\gamma\in\mathfrak{g}(F)=\mathfrak{gl}_{n}(F) be a regular semisimple matrix. As in 2.1.2 we identify GrG\textup{Gr}_{G} with the moduli space of lattices in FnF^{n}, or more precisely GrG(R)\textup{Gr}_{G}(R) is the set of lattices in R((t))nR(\!(t)\!)^{n}. Then 𝒳~γ(R)\widetilde{\mathcal{X}}_{\gamma}(R) can be identified with those lattices ΛR((t))n\Lambda\subset R(\!(t)\!)^{n} such that γΛΛ\gamma\Lambda\subset\Lambda, i.e., those stable under the endomorphism of R((t))nR(\!(t)\!)^{n} given by γ\gamma.

When G=SLnG=\textup{SL}_{n}, GrG(R)\textup{Gr}_{G}(R) classifies lattices Λ\Lambda in R((t))nR(\!(t)\!)^{n} such that [Λ:R[[t]]n]=0[\Lambda:R[\![t]\!]^{n}]=0. The affine Springer fiber 𝒳~γ\widetilde{\mathcal{X}}_{\gamma} in this case is cut out by the same condition γΛΛ\gamma\Lambda\subset\Lambda.

When G=Sp2nG=\textup{Sp}_{2n}, GrG(R)\textup{Gr}_{G}(R) classifies lattices Λ\Lambda in R((t))2nR(\!(t)\!)^{2n} such that Λ=Λ\Lambda^{\vee}=\Lambda, see §2.1.6. The affine Springer fiber 𝒳~γ\widetilde{\mathcal{X}}_{\gamma} in this case is also cut out by the same condition γΛΛ\gamma\Lambda\subset\Lambda.

We give the simplest examples of affine Springer fibers.

2.2.3.

Let γ𝔱(𝒪F)\gamma\in\mathfrak{t}(\mathcal{O}_{F}) such that the reduction γ¯𝔱\overline{\gamma}\in\mathfrak{t} is regular semisimple. For each cocharacter λ:𝔾mT\lambda:\mathbb{G}_{m}\to T, the element tλ:=λ(t)T(F)t^{\lambda}:=\lambda(t)\in T(F) gives a point [tλ]GrG(k)[t^{\lambda}]\in\textup{Gr}_{G}(k) which lies in 𝒳~γ\widetilde{\mathcal{X}}_{\gamma} since Ad(tλ)γ=γ𝔤(𝒪F)\textup{Ad}(t^{-\lambda})\gamma=\gamma\in\mathfrak{g}(\mathcal{O}_{F}). The reduced ind-scheme 𝒳γ\mathcal{X}_{\gamma} is in fact the discrete set {[tλ]}\{[t^{\lambda}]\} which is in bijection with 𝕏(T)\mathbb{X}_{*}(T). More canonically, there is an action of the loop group LTLT on 𝒳~γ\widetilde{\mathcal{X}}_{\gamma} given by its left translation action on GrG\textup{Gr}_{G}. This action factors through the quotient GrT=LT/L+T\textup{Gr}_{T}=LT/L^{+}T and realizes 𝒳~γ\widetilde{\mathcal{X}}_{\gamma} as a GrT\textup{Gr}_{T}-torsor.

2.2.4.

Consider the case G=SL2G=\textup{SL}_{2} and γ=(t00t)\gamma=\left(\begin{array}[]{cc}t&0\\ 0&-t\end{array}\right). Then 𝒳γ\mathcal{X}_{\gamma} is an infinite chain of 1\mathbb{P}^{1}’s. More precisely, for each nn\in\mathbb{Z}, we consider the subscheme CnC_{n} of GrG\textup{Gr}_{G} classifying lattices ΛF2\Lambda\subset F^{2} such that tn𝒪Ftn+1𝒪FΛtn1𝒪Ftn𝒪Ft^{n}\mathcal{O}_{F}\oplus t^{-n+1}\mathcal{O}_{F}\subset\Lambda\subset t^{n-1}\mathcal{O}_{F}\oplus t^{-n}\mathcal{O}_{F}. Then Cn1C_{n}\cong\mathbb{P}^{1}. We have 𝒳γ=nCn\mathcal{X}_{\gamma}=\cup_{n\in\mathbb{Z}}C_{n} is an infinite chain of 1\mathbb{P}^{1}’s. The components CnC_{n} and Cn+1C_{n+1} intersect at one point tn𝒪Ftn𝒪Ft^{-n}\mathcal{O}_{F}\oplus t^{n}\mathcal{O}_{F} and otherwise the components are disjoint.

Here is a way to calculate the kk-points of 𝒳γ\mathcal{X}_{\gamma}. We use the Iwasawa decomposition for G(F)G(F):

G(F)=nN(F)(tn00tn)G(𝒪F).G(F)=\bigsqcup_{n\in\mathbb{Z}}N(F)\left(\begin{array}[]{cc}t^{n}&0\\ 0&t^{-n}\end{array}\right)G(\mathcal{O}_{F}).

According to this decomposition, any point in GrG\textup{Gr}_{G} can be represented by

(15) g=(1x01)(tn00tn)g=\left(\begin{array}[]{cc}1&x\\ 0&1\end{array}\right)\left(\begin{array}[]{cc}t^{n}&0\\ 0&t^{-n}\end{array}\right)

for some xFx\in F and a unique nn\in\mathbb{Z}, and xx has a well-defined image in F/t2n𝒪FF/t^{2n}\mathcal{O}_{F}. Since

Ad(g1)γ=g1γg=(t2t12nx0t)\textup{Ad}(g^{-1})\gamma=g^{-1}\gamma g=\left(\begin{array}[]{cc}t&2t^{1-2n}x\\ 0&-t\end{array}\right)

the condition Ad(g1)γ𝔤(𝒪F)\textup{Ad}(g^{-1})\gamma\in\mathfrak{g}(\mathcal{O}_{F}) is the same as requiring xt2n1𝒪Fx\in t^{2n-1}\mathcal{O}_{F}. Therefore 𝒳γ(k)=nYn\mathcal{X}_{\gamma}(k)=\sqcup_{n\in\mathbb{Z}}Y_{n} where YnY_{n} consists of elements of the form (15) with xt2n1𝒪F/t2n𝒪Fx\in t^{2n-1}\mathcal{O}_{F}/t^{2n}\mathcal{O}_{F}. Therefore each YnY_{n} can be identified with kk. It is easy to check that YnCn(k)Y_{n}\subset C_{n}(k).

2.2.5.

Consider the case G=SL2G=\textup{SL}_{2} and γ=(0t2t0)\gamma=\left(\begin{array}[]{cc}0&t^{2}\\ t&0\end{array}\right). Then 𝒳γ\mathcal{X}_{\gamma} consists exactly of those lattices ΛGrG\Lambda\in\textup{Gr}_{G} such that t𝒪F𝒪FΛ𝒪Ft1𝒪Ft\mathcal{O}_{F}\oplus\mathcal{O}_{F}\subset\Lambda\subset\mathcal{O}_{F}\oplus t^{-1}\mathcal{O}_{F}. Therefore 𝒳γ1\mathcal{X}_{\gamma}\cong\mathbb{P}^{1}. Details of these calculations are left to the reader, see Exercises 2.8.2.

2.2.6. Invariance under conjugation

If γ,γ𝔤(F)\gamma,\gamma^{\prime}\in\mathfrak{g}(F) are related by Ad(g)γ=γ\textup{Ad}(g)\gamma=\gamma^{\prime} for some gG(F)g\in G(F), then the left multiplication by gg on GrG\textup{Gr}_{G} restricts to an isomorphism 𝒳~γ𝒳~γ\widetilde{\mathcal{X}}_{\gamma}\cong\widetilde{\mathcal{X}}_{\gamma^{\prime}}, hence also 𝒳γ𝒳γ\mathcal{X}_{\gamma}\cong\mathcal{X}_{\gamma^{\prime}}. Therefore the isomorphism type of 𝒳γ\mathcal{X}_{\gamma} is invariant under G(F)G(F)-conjugation on γ\gamma. Recall we have map χ:𝔤𝔠:=𝔤G𝔱W\chi:\mathfrak{g}\to\mathfrak{c}:=\mathfrak{g}\sslash G\cong\mathfrak{t}\sslash W. For a regular semisimple point a𝔠rs(F)a\in\mathfrak{c}^{\textup{rs}}(F), the fiber χ1(a)\chi^{-1}(a) is a single G(F)G(F)-conjugacy class (here we are using that the residue field of FF is algebraically closed). Therefore, the isomorphism type of 𝒳γ\mathcal{X}_{\gamma} depends only on a=χ(γ)𝔠rs(F)a=\chi(\gamma)\in\mathfrak{c}^{\textup{rs}}(F).

Unlike Springer fibers, 𝒳γ\mathcal{X}_{\gamma} can be empty for certain γ\gamma. The affine Springer fiber 𝒳γ\mathcal{X}_{\gamma} is nonempty if and only if a=χ(γ)𝔠(𝒪F)a=\chi(\gamma)\in\mathfrak{c}(\mathcal{O}_{F}). In fact, if gG(𝒪F)𝒳γ(k)gG(\mathcal{O}_{F})\in\mathcal{X}_{\gamma}(k), then Ad(g1)γ𝔤(𝒪F)\textup{Ad}(g^{-1})\gamma\in\mathfrak{g}(\mathcal{O}_{F}) hence χ(γ)=χ(Ad(g1)γ)𝔠(𝒪F)\chi(\gamma)=\chi(\textup{Ad}(g^{-1})\gamma)\in\mathfrak{c}(\mathcal{O}_{F}). Conversely, we have a Kostant section ϵ:𝔠𝔤\epsilon:\mathfrak{c}\to\mathfrak{g} of χ\chi which identifies 𝔠\mathfrak{c} with e+𝔤fe+\mathfrak{g}_{f}, where (e,h,f)(e,h,f) is a regular 𝔰𝔩2\mathfrak{sl}_{2}-triple in 𝔤\mathfrak{g}. Therefore, for any a𝔠(𝒪F)𝔠rs(F)a\in\mathfrak{c}(\mathcal{O}_{F})\cap\mathfrak{c}^{\textup{rs}}(F), ϵ(a)𝔤(𝒪F)\epsilon(a)\in\mathfrak{g}(\mathcal{O}_{F}), and 𝒳ϵ(a)\mathcal{X}_{\epsilon(a)} contains the unit coset in GrG\textup{Gr}_{G} hence nonempty; since 𝒳γ\mathcal{X}_{\gamma} is isomorphic to 𝒳ϵ(χ(γ))\mathcal{X}_{\epsilon(\chi(\gamma))}, it is also nonempty.

For a𝔠(𝒪F)𝔠rs(F)a\in\mathfrak{c}(\mathcal{O}_{F})\cap\mathfrak{c}^{\textup{rs}}(F), we also write 𝒳a\mathcal{X}_{a} for 𝒳ϵ(a)\mathcal{X}_{\epsilon(a)}. The above discussion shows that all Springer fibers 𝒳γ\mathcal{X}_{\gamma} are isomorphic to 𝒳a\mathcal{X}_{a} for a=χ(γ)a=\chi(\gamma).

2.2.7. Parahoric versions

For each parahoric subgroup 𝐏LG\mathbf{P}\subset LG, we may similarly define the closed sub-indscheme 𝒳~𝐏,γFl𝐏\widetilde{\mathcal{X}}_{\mathbf{P},\gamma}\subset\textup{Fl}_{\mathbf{P}} using the analog of the condition (14) with 𝔤(R[[t]])\mathfrak{g}(R[\![t]\!]) replaced by (Lie𝐏)^kR(\textup{Lie}\ \mathbf{P})\widehat{\otimes}_{k}R. The reduced ind-scheme 𝒳𝐏,γ=𝒳~𝐏,γredFl𝐏\mathcal{X}_{\mathbf{P},\gamma}=\widetilde{\mathcal{X}}_{\mathbf{P},\gamma}^{\textup{red}}\subset\textup{Fl}_{\mathbf{P}} is called the affine Springer fiber of γ\gamma of type 𝐏\mathbf{P}. In particular, when 𝐈\mathbf{I} is an Iwahori subgroup of LGLG, we denote 𝒳𝐈,γ\mathcal{X}_{\mathbf{I},\gamma} by 𝒴γ\mathcal{Y}_{\gamma}.

For 𝐏𝐐\mathbf{P}\subset\mathbf{Q} two parahoric subgroups of LGLG, the natural projection Fl𝐏Fl𝐐\textup{Fl}_{\mathbf{P}}\to\textup{Fl}_{\mathbf{Q}} induces a map 𝒳𝐏,γ𝒳𝐐,γ\mathcal{X}_{\mathbf{P},\gamma}\to\mathcal{X}_{\mathbf{Q},\gamma}. In particular we have a map 𝒴γ𝒳γ\mathcal{Y}_{\gamma}\to\mathcal{X}_{\gamma}.

2.3. Symmetry on affine Springer fibers

For the Springer fiber e\mathcal{B}_{e}, the centralizer CG(e)C_{G}(e) acts on it. In this subsection we investigate a similar structure for affine Springer fibers.

2.3.1. Centralizer action

Let GγG_{\gamma} be the centralizer of γ\gamma in GFG_{F} (the algebraic group over FF obtained from GG by base change). Then GγG_{\gamma} is an algebraic group over FF. Since γ\gamma is regular semisimple, GγG_{\gamma} is a torus over FF. One can define the loop group LGγLG_{\gamma} of GγG_{\gamma} as the functor RGγ(R((t)))R\mapsto G_{\gamma}(R(\!(t)\!)) on kk-algebras.

We claim that LGγLG_{\gamma} acts on the ind-scheme 𝒳~γ\widetilde{\mathcal{X}}_{\gamma}. This can be seen on the level of RR-points. Suppose hLGγ(R)=Gγ(R((t)))h\in LG_{\gamma}(R)=G_{\gamma}(R(\!(t)\!)) and [g]𝒳~γ(R)[g]\in\widetilde{\mathcal{X}}_{\gamma}(R). Then the coset [hg]GrG(R)[hg]\in\textup{Gr}_{G}(R) still satisfies

Ad((hg)1)γ=Ad(g1)Ad(h1)γ=Ad(g1)γ𝔤(R[[t]])\textup{Ad}((hg)^{-1})\gamma=\textup{Ad}(g^{-1})\textup{Ad}(h^{-1})\gamma=\textup{Ad}(g^{-1})\gamma\in\mathfrak{g}(R[\![t]\!])

using that hh is in the centralizer of γ\gamma. Therefore [hg]𝒳~γ(R)[hg]\in\widetilde{\mathcal{X}}_{\gamma}(R). The assignment [g][hg][g]\mapsto[hg] for hLGγh\in LG_{\gamma} and [g]𝒳~γ[g]\in\widetilde{\mathcal{X}}_{\gamma} defines an action of LGγLG_{\gamma} on 𝒳~γ\widetilde{\mathcal{X}}_{\gamma}. It induces an action of LGγLG_{\gamma} on the reduced structure 𝒳γ\mathcal{X}_{\gamma}.

2.3.2. The split case

We consider the case where γ𝔱rs(F)\gamma\in\mathfrak{t}^{\textup{rs}}(F). In this case Gγ=TkFG_{\gamma}=T\otimes_{k}F, and

LGγ=LT𝕏(T)L𝔾mLG_{\gamma}=LT\cong\mathbb{X}_{*}(T)\otimes_{\mathbb{Z}}L\mathbb{G}_{m}

where L𝔾mL\mathbb{G}_{m} is the loop group of the multiplicative group 𝔾m\mathbb{G}_{m}. For any kk-algebra RR, L𝔾m(R)=R((t))×L\mathbb{G}_{m}(R)=R(\!(t)\!)^{\times}. It is easy to see that an element a=iaitiR((t))a=\sum_{i}a_{i}t^{i}\in R(\!(t)\!) is invertible if and only if aa starts with finitely many nilpotent coefficients and the first non-nilpotent coefficient is invertible in RR. When RR is reduced, the leading coefficient of aa must be invertible in RR, which implies R((t))×=tR[[t]]×R(\!(t)\!)^{\times}=t^{\mathbb{Z}}\cdot R[\![t]\!]^{\times}, and R[[t]]×={(a0,a1,)|a0R×,aiR,i1}R[\![t]\!]^{\times}=\{(a_{0},a_{1},\cdots)|a_{0}\in R^{\times},a_{i}\in R,\forall i\geq 1\}. We see that the reduced ind-scheme (L𝔾m)red×L+𝔾m(L\mathbb{G}_{m})^{\textup{red}}\cong\mathbb{Z}\times L^{+}\mathbb{G}_{m}, and that L+𝔾m𝔾m×𝔸L^{+}\mathbb{G}_{m}\cong\mathbb{G}_{m}\times\mathbb{A}^{\mathbb{N}} as schemes, where 𝔸=Speck[x1,x2,]\mathbb{A}^{\mathbb{N}}=\textup{Spec}\ k[x_{1},x_{2},\cdots]. Therefore, when γ𝔱rs(F)\gamma\in\mathfrak{t}^{\textup{rs}}(F), we have (LT)red𝕏(T)×L+T(LT)^{\textup{red}}\cong\mathbb{X}_{*}(T)\times L^{+}T, and L+TL^{+}T is an affine scheme of infinite type.

2.3.3. The lattice Λγ\Lambda_{\gamma}

For a general regular semisimple γ𝔤(F)\gamma\in\mathfrak{g}(F), let 𝕏(Gγ):=HomF(𝔾m,Gγ)\mathbb{X}_{*}(G_{\gamma}):=\textup{Hom}_{F}(\mathbb{G}_{m},G_{\gamma}) be the FF-rational cocharacter lattice of the torus GγG_{\gamma}. For each λ𝕏(Gγ)\lambda\in\mathbb{X}_{*}(G_{\gamma}) viewed as a homomorphism 𝔾mGγ\mathbb{G}_{m}\to G_{\gamma} defined over FF, we may consider the element λ(t)\lambda(t). The assignment λλ(t)\lambda\mapsto\lambda(t) defines an injective homomorphism

𝕏(Gγ)Gγ(F).\mathbb{X}_{*}(G_{\gamma})\hookrightarrow G_{\gamma}(F).

whose image is denoted by Λγ\Lambda_{\gamma}. It can be shown that the quotient Λγ\(LGγ)red\Lambda_{\gamma}\backslash(LG_{\gamma})^{\textup{red}} is an affine scheme that is a finite disjoint union of 𝔾ma×𝔸\mathbb{G}_{m}^{a}\times\mathbb{A}^{\mathbb{N}} for some integer aa.

2.3.4. The case G=GLnG=\textup{GL}_{n}

We continue with the setup of §2.2.2. We assume that char(k)>n\textup{char}(k)>n. Then the characteristic polynomial P(x)=xn+a1xn1++anF[x]P(x)=x^{n}+a_{1}x^{n-1}+\cdots+a_{n}\in F[x] of γ\gamma is separable. The FF-algebra F[x]/(P(x))F[x]/(P(x)) is then a product of fields F1××FmF_{1}\times\cdots\times F_{m}, with i=1m[Fi:F]=n\sum_{i=1}^{m}[F_{i}:F]=n. Each field extension Fi/FF_{i}/F is obtained by adjoining a root of an irreducible factor Pi(x)P_{i}(x) of P(x)P(x), and FiF_{i} is necessarily of the form k((t1/ei))k(\!(t^{1/e_{i}})\!) since char(k)>n\textup{char}(k)>n. Then the centralizer GγG_{\gamma} is isomorphic to the product of induced tori

Gγi=1mResFFi𝔾m.G_{\gamma}\cong\prod_{i=1}^{m}\textup{Res}^{F_{i}}_{F}\mathbb{G}_{m}.

We have 𝕏(Gγ)m\mathbb{X}_{*}(G_{\gamma})\cong\mathbb{Z}^{m}, and the map 𝕏(Gγ)Gγ(F)\mathbb{X}_{*}(G_{\gamma})\to G_{\gamma}(F) is given by

m(d1,,dm)(td1,,tdm)F1×××Fm×\mathbb{Z}^{m}\ni(d_{1},\cdots,d_{m})\mapsto(t^{d_{1}},\cdots,t^{d_{m}})\in F^{\times}_{1}\times\cdots\times F^{\times}_{m}

The quotient Λγ\Gγ(F)\Lambda_{\gamma}\backslash G_{\gamma}(F) is isomorphic to i=1mFi×/t\prod_{i=1}^{m}F^{\times}_{i}/t^{\mathbb{Z}}. Since each FiF_{i} is isomorphic to k((t1/ei))k(\!(t^{1/e_{i}})\!), we have an exact sequence 1𝒪Fi×Fi×/t/ei01\to\mathcal{O}^{\times}_{F_{i}}\to F^{\times}_{i}/t^{\mathbb{Z}}\to\mathbb{Z}/e_{i}\mathbb{Z}\to 0, and hence the quotient Λγ\(LGγ)red\Lambda_{\gamma}\backslash(LG_{\gamma})^{\textup{red}} contains the group scheme i=1mLFi+𝔾m\prod_{i=1}^{m}L^{+}_{F_{i}}\mathbb{G}_{m} with finite index. Here LFi+𝔾mL^{+}_{F_{i}}\mathbb{G}_{m} is isomorphic to L+𝔾mL^{+}\mathbb{G}_{m} as a scheme, except that we are renaming the uniformizer t1/eit^{1/e_{i}}.

Alternatively, we may fix a uniformizer tiFit_{i}\in F_{i} (for example take ti=t1/eit_{i}=t^{1/e_{i}}) and let Λ~γ=t1××tmiFi×=Gγ(F)\widetilde{\Lambda}_{\gamma}=t_{1}^{\mathbb{Z}}\times\cdots\times t^{\mathbb{Z}}_{m}\subset\prod_{i}F_{i}^{\times}=G_{\gamma}(F). The lattice Λ~γ\widetilde{\Lambda}_{\gamma} will be useful in calculating orbital integrals, see §3.3.2.

2.3.5. The case G=SL2G=\textup{SL}_{2}

Let G=SL2G=\textup{SL}_{2} and γ=(0tn10)\gamma=\left(\begin{array}[]{cc}0&t^{n}\\ 1&0\end{array}\right) where n1n\geq 1 is odd. Then Gγ(F)G_{\gamma}(F) consists of matrices (abtnba)\left(\begin{array}[]{cc}a&bt^{n}\\ b&a\end{array}\right) with a,bFa,b\in F and a2tnb2=1a^{2}-t^{n}b^{2}=1. Note that this equation forces a,b𝒪Fa,b\in\mathcal{O}_{F}, hence Gγ(F)=Gγ(𝒪F)G_{\gamma}(F)=G_{\gamma}(\mathcal{O}_{F}). The torus GγG_{\gamma} is non-split and splits over the quadratic extension E=F(t1/2)E=F(t^{1/2}). The lattice Λγ=HomF(𝔾m,Gγ)=0\Lambda_{\gamma}=\textup{Hom}_{F}(\mathbb{G}_{m},G_{\gamma})=0. Writing a=i0aitia=\sum_{i\geq 0}a_{i}t^{i} and bi=i0bitib_{i}=\sum_{i\geq 0}b_{i}t^{i}, we see that a0=±1a_{0}=\pm 1, and once bb and a0a_{0} are fixed, the higher coefficients of aa can be solved uniquely using the Taylor expansion of (1+tnb2)1/2(1+t^{n}b^{2})^{1/2}. Therefore (LGγ)redL+Gγ(LG_{\gamma})^{\textup{red}}\cong L^{+}G_{\gamma} is isomorphic to {±1}×𝔸\{\pm 1\}\times\mathbb{A}^{\mathbb{N}}, given by (a,b)(a0,b0,b1,)(a,b)\mapsto(a_{0},b_{0},b_{1},\cdots).

2.3.6. Symmetry on affine Springer fibers

Ngô has found a more precise statement about the action of LGγLG_{\gamma} on 𝒳~γ\widetilde{\mathcal{X}}_{\gamma}, namely the action factors through a canonical finite-dimensional quotient. We sketch the story following [NgoFL, §3.3]. Let a=χ(γ)𝔠(F)rsa=\chi(\gamma)\in\mathfrak{c}(F)^{\textup{rs}} be the image of γ\gamma under χ:𝔤𝔠\chi:\mathfrak{g}\to\mathfrak{c}. We assume a𝔠(𝒪F)a\in\mathfrak{c}(\mathcal{O}_{F}) for otherwise 𝒳γ\mathcal{X}_{\gamma} is empty.

There is a smooth affine group scheme JJ over 𝔠\mathfrak{c} called the regular centralizer group scheme. It is characterized by the property that its pullback to 𝔤\mathfrak{g} via χ\chi, denoted χJ\chi^{*}J, maps into the universal centralizer group scheme II over 𝔤\mathfrak{g}, and this map is an isomorphism over the regular locus 𝔤reg\mathfrak{g}^{\textup{reg}}. Let JaJ_{a} be pullback of JJ under the map a:Spec𝒪F𝔠a:\textup{Spec}\ \mathcal{O}_{F}\to\mathfrak{c}. Then JaJ_{a} is a smooth affine group scheme over 𝒪F\mathcal{O}_{F} whose FF-fiber is the torus GγG_{\gamma} (i.e., JaJ_{a} is an integral model of GγG_{\gamma} over 𝒪F\mathcal{O}_{F}). We may form the positive loop group L+JaL^{+}J_{a} of JaJ_{a} as well as its affine Grassmannian Pa:=LGγ/L+JaP_{a}:=LG_{\gamma}/L^{+}J_{a} (also called the local Picard group). The reduced group scheme ParedP^{\textup{red}}_{a} is finite-dimensional and locally of finite type. Ngô showed that the action of LGγLG_{\gamma} on 𝒳~γ\widetilde{\mathcal{X}}_{\gamma} (and hence on 𝒳γ\mathcal{X}_{\gamma}) factors through the local Picard group PaP_{a}, and it does not factor through any further quotient. For related statement, see §2.5.8.

2.4. Further examples of affine Springer fibers

In this subsection we give more examples illustrating the rich geometry of affine Springer fibers. We omit the calculations that lead to the geometric descriptions.

In all examples below, the affine Springer fibers are homogeneous in the sense that 𝒳γ\mathcal{X}_{\gamma} is equipped an extra 𝔾m\mathbb{G}_{m}-action coming from the loop rotation on GrG\textup{Gr}_{G} by dilation on the uniformizer tt. For more information on homogeneous affine Springer fibers and their application to representation theory, see [OY].

2.4.1.

When G=SL2G=\textup{SL}_{2} and γ=(0tm+1tm0)\gamma=\left(\begin{array}[]{cc}0&t^{m+1}\\ t^{m}&0\end{array}\right). When mm is even, 𝒳γ\mathcal{X}_{\gamma} consists exactly of those lattices Λ\Lambda such that

tm/2𝒪Ftm/2𝒪FΛtm/2𝒪Ftm/2𝒪F.t^{m/2}\mathcal{O}_{F}\oplus t^{m/2}\mathcal{O}_{F}\subset\Lambda\subset t^{-m/2}\mathcal{O}_{F}\oplus t^{-m/2}\mathcal{O}_{F}.

In this case, 𝒳γ\mathcal{X}_{\gamma} coincides with the closure of the L+GL^{+}G-orbit in GrG\textup{Gr}_{G} corresponding to the coweight mα/2m\alpha^{\vee}/2.

When mm is odd, 𝒳γ\mathcal{X}_{\gamma} consists exactly of those lattices Λ\Lambda such that

t(m+1)/2𝒪Ft(m1)/2𝒪FΛt(m1)/2𝒪Ft(m+1)/2𝒪F.t^{(m+1)/2}\mathcal{O}_{F}\oplus t^{(m-1)/2}\mathcal{O}_{F}\subset\Lambda\subset t^{-(m-1)/2}\mathcal{O}_{F}\oplus t^{-(m+1)/2}\mathcal{O}_{F}.

In this case, consider instead the affine Grassmannian GrG\textup{Gr}_{G^{\prime}} of G=PGL2G^{\prime}=\textup{PGL}_{2}, which contains GrG\textup{Gr}_{G} as a component. Then 𝒳γ\mathcal{X}_{\gamma} can be identified with the closure of the L+GL^{+}G^{\prime}-orbit in GrG\textup{Gr}_{G^{\prime}} corresponding to the coweight mα/2m\alpha^{\vee}/2.

In either case, we have dim𝒳γ=m\dim\mathcal{X}_{\gamma}=m.

2.4.2. The Lusztig-Smelt examples

Let G=GLnG=\textup{GL}_{n} and γ𝔤(F)\gamma\in\mathfrak{g}(F) with characteristic polynomial P(x)=xntm=0P(x)=x^{n}-t^{m}=0, where (m,n)=1(m,n)=1. If a lattice ΛFn\Lambda\subset F^{n} is stable under γ\gamma, it carries an action of the ring R=𝒪F[X]/(Xntm)R=\mathcal{O}_{F}[X]/(X^{n}-t^{m}). Let K=Frac(R)K=\textup{Frac}(R), then K=k((s))K=k(\!(s)\!) with x=smx=s^{m} and t=snt=s^{n}. Then the integral closure of RR in KK is R~:=k[[s]]\widetilde{R}:=k[\![s]\!]. The action of γ\gamma on FnF^{n} makes it a one-dimensional KK-vector space. We fix a KK-linear isomorphism FnKF^{n}\cong K, under which a lattice in FnF^{n} stable under RR is simply a fractional RR-ideal, i.e., a finitely generated RR-submodule of K=Frac(R)K=\textup{Frac}(R). Then 𝒳γ\mathcal{X}_{\gamma} can be identified with the moduli space of fractional RR-ideals.

The centralizer Gγ(F)G_{\gamma}(F) is simply K×K^{\times}, which acts on the set of fractional RR-ideals by multiplication. This action clearly factors through K×/R×K^{\times}/R^{\times}, which is the group of kk-points of the local Picard group scheme PγP_{\gamma}.

There is an action of 𝔾m\mathbb{G}_{m} on KK (by field automorphisms) given by scaling ss (so sis^{i} gets weight ii under this action). This induces an action of 𝔾m\mathbb{G}_{m} on 𝒳γ\mathcal{X}_{\gamma}. The fixed points of 𝔾m\mathbb{G}_{m} on 𝒳γ\mathcal{X}_{\gamma} correspond to fractional ideals generated by monomials of ss. More precisely, if a fractional RR-ideal ΛK=k((s))\Lambda\subset K=k(\!(s)\!) is fixed by 𝔾m\mathbb{G}_{m}, define MΛ={i|siΛ}M_{\Lambda}=\{i\in\mathbb{Z}|s^{i}\in\Lambda\} which is a subset of \mathbb{Z} stable under adding mm and nn, because Λ\Lambda is an R=k[[sm,sn]]R=k[\![s^{m},s^{n}]\!]-module. Therefore MΛM_{\Lambda}\subset\mathbb{Z} is a finitely generated module for the monoid Am,n:=0m+0n0A_{m,n}:=\mathbb{Z}_{\geq 0}m+\mathbb{Z}_{\geq 0}n\subset\mathbb{Z}_{\geq 0}. The assignment ΛMΛ\Lambda\mapsto M_{\Lambda} gives a bijection

𝒳γ𝔾m(k){Am,n-submodules M}.\mathcal{X}_{\gamma}^{\mathbb{G}_{m}}(k)\stackrel{{\scriptstyle\sim}}{{\to}}\{\mbox{$A_{m,n}$-submodules $M\subset\mathbb{Z}$}\}.

Any Am,nA_{m,n}-submodule of \mathbb{Z} contains all sufficiently large integers. Therefore any two such Am,nA_{m,n}-module MM and MM^{\prime} differ by finitely many elements, and we can define [M:M]=#(M\M)#(M\M)[M:M^{\prime}]=\#(M\backslash M^{\prime})-\#(M^{\prime}\backslash M). Fox any fixed ii\in\mathbb{Z}, we have a total of 1n+m(n+mn)\frac{1}{n+m}\binom{n+m}{n} fixed points with [M:0]=i[M:\mathbb{Z}_{\geq 0}]=i. For a fixed point pMp_{M} corresponding to an Am,nA_{m,n}-module MM, consider the subvariety CM={p𝒳γ|lim𝔾mz0zp=pM}C_{M}=\{p\in\mathcal{X}_{\gamma}|\lim_{\mathbb{G}_{m}\ni z\to 0}z\cdot p=p_{M}\}. Then CMC_{M} is isomorphic to an affine space whose dimension can be expressed combinatorially in terms of MM. The cells CMC_{M} give a stratification of 𝒳γ\mathcal{X}_{\gamma}. This gives a way to compute the Poincaré polynomial of connected components of 𝒳γ\mathcal{X}_{\gamma}. For more details, and the similar picture for 𝒴γ\mathcal{Y}_{\gamma}, see [LS].

2.4.3.

We look at the geometry of 𝒴γ\mathcal{Y}_{\gamma} in a special case of §2.4.2. We consider the case G=GL3G=\textup{GL}_{3} and γ3t2=0\gamma^{3}-t^{2}=0. Introducing the variable ss with t=s3t=s^{3} and γ=s2\gamma=s^{2} as before, then R=k[[s2,s3]]R=k[\![s^{2},s^{3}]\!] with fraction field K=k((s))K=k(\!(s)\!). The affine Springer fiber 𝒴γ\mathcal{Y}_{\gamma} classifies a chain of fractional RR-ideals Λ0Λ1Λ2s3Λ0\Lambda_{0}\subset\Lambda_{1}\subset\Lambda_{2}\subset s^{-3}\Lambda_{0}. We consider a component of 𝒴γ0𝒴γ\mathcal{Y}^{0}_{\gamma}\subset\mathcal{Y}_{\gamma}, classifying chains {Λi}\{\Lambda_{i}\} as above with [Λi:k[[s]]]=i[\Lambda_{i}:k[\![s]\!]]=i, 0i20\leq i\leq 2.

We first study the 𝔾m\mathbb{G}_{m}-fixed points on 𝒴γ0\mathcal{Y}^{0}_{\gamma}. For each 𝔾m\mathbb{G}_{m}-fixed RR-submodule of k((s))k(\!(s)\!), we denote its associated module for the monoid A3,2A_{3,2} (see §2.4.2) by a sequence of integers. For example, (0,1,2,)(0,1,2,\cdots) stands for the standard lattice k[[s]]k[\![s]\!]. Note that the sequence for any 𝔾m\mathbb{G}_{m}-fixed RR-fractional ideal is either consecutive (n,n+1,n+2,)(n,n+1,n+2,\cdots) or has at most one gap at the second place (n,n+2,n+3,)(n,n+2,n+3,\cdots). We have the following four fixed points in 𝒴γ0\mathcal{Y}^{0}_{\gamma}:

  1. (1)

    q:(0,1,2,)(1,0,1,)(2,1,0,)q:(0,1,2,\cdots)\subset(-1,0,1,\cdots)\subset(-2,-1,0,\cdots);

  2. (2)

    p0:(1,1,2,)(1,0,1,)(2,1,0,)p_{0}:(-1,1,2,\cdots)\subset(-1,0,1,\cdots)\subset(-2,-1,0,\cdots);

  3. (3)

    p1:(0,1,2,)(2,0,1,)(2,1,0,)p_{1}:(0,1,2,\cdots)\subset(-2,0,1,\cdots)\subset(-2,-1,0,\cdots);

  4. (4)

    p2:(0,1,2,)(1,0,1,)(3,1,0,)p_{2}:(0,1,2,\cdots)\subset(-1,0,1,\cdots)\subset(-3,-1,0,\cdots).

Our indexing scheme is that pip_{i} is obtained from qq by changing the lattice Λi\Lambda_{i}.

Then 𝒴γ0\mathcal{Y}^{0}_{\gamma} is the union of three irreducible components C0C1C2C_{0}\cup C_{1}\cup C_{2}, and each component CiC_{i} is isomorphic to 1\mathbb{P}^{1}. They all contain qq and that is the only intersection between any two of them. We have piCip_{i}\in C_{i} for i=0,1,2i=0,1,2.

There is a natural way to index affine partial flag varieties of GG by subsets J{0,1,2}J\subset\{0,1,2\}, as we saw in §2.1.5. Let 𝒳J,γ\mathcal{X}_{J,\gamma} be the affine Springer fiber of γ\gamma in FlJ\textup{Fl}_{J}. Under the projection 𝒴γ𝒳J,γ\mathcal{Y}_{\gamma}\to\mathcal{X}_{J,\gamma}, the curves CiC_{i} for iJi\notin J collapse to a point, and the other curves map isomorphically onto their images.

2.4.4.

Let G=Sp(V)G=\textup{Sp}(V) where VV is a symplectic space of dimension 2n2n over kk, and assume char(k)2\textup{char}(k)\neq 2. Fix a decomposition V=UUV=U\oplus U^{*} into Lagrangian subspaces of VV, such that the symplectic form restricts to the natural pairing on U×UU\times U^{*}. Consider γ=(0tXY0)𝔤(F)\gamma=\left(\begin{array}[]{cc}0&tX\\ Y&0\end{array}\right)\in\mathfrak{g}(F) where XSym2(U)X\in\textup{Sym}^{2}(U) (viewed as a self-adjoint map ξ:UU\xi:U^{*}\to U) and YSym2(U)Y\in\textup{Sym}^{2}(U^{*}) (viewed as a self-adjoint map η:UU\eta:U\to U^{*}). The condition that γ\gamma is regular semisimple is equivalent to that: (1) both ξ\xi and η\eta are isomorphisms; (2) ξηGL(U)\xi\eta\in\textup{GL}(U) is regular semisimple, or equivalently ηξGL(U)\eta\xi\in\textup{GL}(U^{*}) is regular semisimple.

The affine Springer fiber 𝒳γ\mathcal{X}_{\gamma} classifies self-dual lattices ΛVkF\Lambda\subset V\otimes_{k}F which are stable under γ\gamma (see §2.2.2). Consider the action of 𝔾m\mathbb{G}_{m} on VkF=UkFUkFV\otimes_{k}F=U\otimes_{k}F\oplus U^{*}\otimes_{k}F such that UtiUt^{i} has weight 2i12i-1 and UtiU^{*}t^{i} has weight 2i2i. This induces a 𝔾m\mathbb{G}_{m}-action on GrG\textup{Gr}_{G} and on 𝒳γ\mathcal{X}_{\gamma}. We first consider the fixed points GrG𝔾m\textup{Gr}_{G}^{\mathbb{G}_{m}}. A lattice ΛGrG\Lambda\in\textup{Gr}_{G} is fixed under 𝔾m\mathbb{G}_{m} if and only if it is the tt-adic completion of

Span{(AiBi)ti;i}\textup{Span}\{(A_{i}\oplus B_{i})t^{i};i\in\mathbb{Z}\}

where A1A0A1U\cdots\subset A_{-1}\subset A_{0}\subset A_{1}\subset\cdots\subset U is a filtration of UU such that AN=0A_{N}=0 for N0N\ll 0 and AN=UA_{N}=U for N0N\gg 0, and B1B0B1U\cdots\subset B_{-1}\subset B_{0}\subset B_{1}\subset\cdots\subset U^{*} is a similar filtration of UU^{*}, such that

(16) Bi1=Ai,iB_{i-1}=A^{\bot}_{-i},\forall i\in\mathbb{Z}

under the duality pairing between UU and UU^{*}. The last condition reflects the fact that Λ\Lambda is self-dual under the symplectic form.

A lattice Λ𝒳γ𝔾m\Lambda\in\mathcal{X}_{\gamma}^{\mathbb{G}_{m}} is then determined by two filtrations AA_{\bullet} of UU and BB_{\bullet} of UU^{*}, dual in the sense (16), with the extra condition that

ηAiBi;ξBiAi+1,i.\eta A_{i}\subset B_{i};\quad\xi B_{i}\subset A_{i+1},\quad\forall i\in\mathbb{Z}.

We summarize the data into the following diagram

A1\textstyle{A_{-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}η\scriptstyle{\eta}A0\textstyle{A_{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}η\scriptstyle{\eta}A1\textstyle{A_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}η\scriptstyle{\eta}\textstyle{\cdots}\textstyle{\cdots}A0\textstyle{A^{\bot}_{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ξ\scriptstyle{\xi}A1\textstyle{A^{\bot}_{-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ξ\scriptstyle{\xi}A2\textstyle{A_{-2}^{\bot}}

For example, when dimV=4\dim V=4, hence dimU=2\dim U=2, there are two possibilities. The first possibility is A1=0A0=UA_{-1}=0\subset A_{0}=U, which corresponds to the standard lattice Vk𝒪FV\otimes_{k}\mathcal{O}_{F}. The second possibility is A1=0A0A1=UA_{-1}=0\subset A_{0}\subset A_{1}=U, and A0UA_{0}\subset U is a line satisfying ξA0A0\xi A^{\bot}_{0}\subset A_{0}, i.e., A0UA^{\bot}_{0}\subset U^{*} is an isotropic line under the quadratic form XX. There are two such lines A0A^{\bot}_{0}, giving two other 𝔾m\mathbb{G}_{m}-fixed points. Therefore, 𝒳γ𝔾m\mathcal{X}_{\gamma}^{\mathbb{G}_{m}} consists of 3 points. We have 𝒳γ=C1C2\mathcal{X}_{\gamma}=C_{1}\cup C_{2} where Ci1C_{i}\cong\mathbb{P}^{1}, the two components intersect at the standard lattice, and each CiC_{i} contains one of the remaining 𝔾m\mathbb{G}_{m}-fixed points.

2.4.5. The Bernstein-Kazhdan example

In [KL, Appendix], Bernstein and Kazhdan gave the first example of an irreducible component of an affine Springer fiber which was not a rational variety. We keep the same notation as in §2.4.4. Let Fl𝐏\textup{Fl}_{\mathbf{P}} be the partial affine flag variety of G=Sp(V)G=\textup{Sp}(V) classifying pairs of lattices ΛΛ\Lambda^{\prime}\subset\Lambda such that Λ=Λ\Lambda^{\vee}=\Lambda and Λ=t1Λ\Lambda^{\prime\vee}=t^{-1}\Lambda. Let γ\gamma be as in §2.4.4. Then the same 𝔾m\mathbb{G}_{m} acts on 𝒳𝐏,γ\mathcal{X}_{\mathbf{P},\gamma}, and the fixed points ΛΛ\Lambda^{\prime}\subset\Lambda can be described by two pairs of filtrations (A,B)(A_{\bullet},B_{\bullet}) and (A,B)(A^{\prime}_{\bullet},B^{\prime}_{\bullet}), where (A,B)(A_{\bullet},B_{\bullet}) is the kind of filtration of UU and UU^{*} as described in §2.4.4, and (A,B)(A^{\prime}_{\bullet},B^{\prime}_{\bullet}) is similar except that (16) is replaced by Bi=AiB^{\prime}_{i}=A_{-i}^{\prime\bot}. Moreover, the inclusion ΛΛ\Lambda^{\prime}\subset\Lambda is equivalent to AiAiA^{\prime}_{i}\subset A_{i} and BiBiB^{\prime}_{i}\subset B_{i}, for all ii.

Consider for example dimV=6\dim V=6 and we fix the dimension of the filtrations:

A1=0,dimA0=2,A1=U;dimA0=1,A1=U.A_{-1}=0,\quad\dim A_{0}=2,\quad A_{1}=U;\quad\dim A^{\prime}_{0}=1,\quad A^{\prime}_{1}=U.

Such filtrations (A,B;A,B)(A_{\bullet},B_{\bullet};A^{\prime}_{\bullet},B^{\prime}_{\bullet}) are determined by the complete flag 0A0A0U0\subset A^{\prime}_{0}\subset A_{0}\subset U satisfying ξA0A0\xi A^{\bot}_{0}\subset A_{0} and that ηA0A0\eta A^{\prime}_{0}\subset A^{\prime\bot}_{0}. In other words, A0A^{\bot}_{0} is an isotropic line in UU^{*} under the quadratic form XX, and A0A^{\prime}_{0} is an isotropic line in UU under the quadratic form YY. The pair (A0,A0)(A^{\bot}_{0},A^{\prime}_{0}) determines a point in Q(X)×Q(Y)(U)×(U)Q(X)\times Q(Y)\subset\mathbb{P}(U)\times\mathbb{P}(U^{*}), the product of conics defined by XX and YY. The incidence relation A0A0A^{\prime}_{0}\subset A_{0} defines a curve of bidegree (2,2)(2,2) in Q(X)×Q(Y)1×1Q(X)\times Q(Y)\cong\mathbb{P}^{1}\times\mathbb{P}^{1}, which is then a curve of genus one. Therefore, a connected component of 𝒳𝐏,γ𝔾m\mathcal{X}^{\mathbb{G}_{m}}_{\mathbf{P},\gamma} is a curve of genus one. Consider the points in 𝒳𝐏,γ\mathcal{X}_{\mathbf{P},\gamma} that contract to this curve, and take its closure ZZ. One can show that dimZ=3=dim𝒳𝐏,γ\dim Z=3=\dim\mathcal{X}_{\mathbf{P},\gamma}. Hence 𝒳𝐏,γ\mathcal{X}_{\mathbf{P},\gamma} contains an irreducible component ZZ which is irrational. We refer to the appendix of [KL] for more details.

2.4.6. “Subregular” affine Springer fibers

When 𝒳γ\mathcal{X}_{\gamma} or 𝒴γ\mathcal{Y}_{\gamma} is one-dimensional, we may call them subregular affine Springer fibers, by analog with subregular Springer fibers discussed in §1.3.8. If dim𝒴γ=1\dim\mathcal{Y}_{\gamma}=1, it is a union of 1\mathbb{P}^{1}’s, hence we can define its dual graph. In [KL, Prop 7.7], the dual graphs of the subregular affine Springer fibers in Fl are classified, and they are almost always the extended Dynkin diagrams of simply-laced groups, except that they can also be infinite chains in type AA (see Example 2.2.4).

2.5. Geometric Properties of affine Springer fibers

2.5.1. Non-reducedness

The ind-scheme 𝒳~γ\widetilde{\mathcal{X}}_{\gamma} is never reduced if GG is nontrivial and γ\gamma is regular semisimple in 𝔤(F)\mathfrak{g}(F). For example, in the case considered in §2.2.3, 𝒳~γ\widetilde{\mathcal{X}}_{\gamma} is isomorphic to GrT=LT/L+T\textup{Gr}_{T}=LT/L^{+}T. We have seen in §2.3.1 that for non-reduced rings RR, elements in LT(R)=R((t))×LT(R)=R(\!(t)\!)^{\times} can have nilpotent leading coefficients. Therefore GrT(R)\textup{Gr}_{T}(R) is not just 𝕏(T)\mathbb{X}_{*}(T), which is GrTred\textup{Gr}^{\textup{red}}_{T}. This shows that GrT\textup{Gr}_{T} is non-reduced, hence 𝒳~γ\widetilde{\mathcal{X}}_{\gamma} is non-reduced.

The next theorem is the fundamental finiteness statement about 𝒳γ\mathcal{X}_{\gamma}.

17 Theorem (Essentially Kazhdan and Lusztig [KL, Prop 2.1]).

Let γ𝔤(F)\gamma\in\mathfrak{g}(F) be a regular semisimple element. Then

  1. (1)

    There exists a closed subscheme Z𝒳γZ\subset\mathcal{X}_{\gamma} which is projective over kk, such that 𝒳γ=ΛγZ\mathcal{X}_{\gamma}=\cup_{\ell\in\Lambda_{\gamma}}\ell\cdot Z.

  2. (2)

    The ind-scheme 𝒳γ\mathcal{X}_{\gamma} is a scheme locally of finite type over kk.

  3. (3)

    The action of Λγ\Lambda_{\gamma} on 𝒳γ\mathcal{X}_{\gamma} is free, and the quotient Λγ\𝒳γ\Lambda_{\gamma}\backslash\mathcal{X}_{\gamma} (as an fppf sheaf on kk-algebras) is representable by a proper algebraic space over kk.

We sketch a proof of this theorem below in three steps.

2.5.2. First reduction

We show that part (1) of the theorem implies (2) and (3). Let ZZ be a projective subscheme as in (1). To show (2), we would like to show that any x𝒳γ(k)x\in\mathcal{X}_{\gamma}(k) has an open neighborhood which is a scheme of finite type. By the Λγ\Lambda_{\gamma}-action we may assume xZ(k)x\in Z(k). Since ZZ is of finite type, the set Σ:={Λγ|ZZ}\Sigma:=\{\ell\in\Lambda_{\gamma}|Z\cap\ell\cdot Z\neq\varnothing\} is finite. Let U=𝒳γΣZU=\mathcal{X}_{\gamma}-\cup_{\ell\notin\Sigma}\ell\cdot Z, then UU is an open neighborhood of ZZ, hence an open neighborhood of xx. Moreover, UU is contained in the finite union ΣZ\cup_{\ell\in\Sigma}\ell\cdot Z, hence contained in some Schubert variety GrG,λ\textup{Gr}_{G,\leq\lambda}. Hence UU is an open subset of the projective scheme 𝒳γGrG,λ\mathcal{X}_{\gamma}\cap\textup{Gr}_{G,\leq\lambda}, therefore UU is itself a scheme of finite type. To show (3), note that the fppf sheaf quotient Λγ\𝒳γ\Lambda_{\gamma}\backslash\mathcal{X}_{\gamma} is a separated algebraic space because it is the quotient of 𝒳γ\mathcal{X}_{\gamma} by the étale equivalence relation Λγ×𝒳γ𝒳γ×𝒳γ\Lambda_{\gamma}\times\mathcal{X}_{\gamma}\subset\mathcal{X}_{\gamma}\times\mathcal{X}_{\gamma} (given by the action and projection maps). By (1), there is a surjection ZΛγ\𝒳γZ\to\Lambda_{\gamma}\backslash\mathcal{X}_{\gamma} from a projective scheme ZZ, which implies that Λγ\𝒳γ\Lambda_{\gamma}\backslash\mathcal{X}_{\gamma} is proper.

2.5.3. Proof of (1) when γ\gamma lies in a split torus

We first consider the case where γ\gamma lies in a split torus. By G(F)G(F)-conjugation, we may assume γ𝔱(F)\gamma\in\mathfrak{t}(F). In this case Λγ𝕏(T)\Lambda_{\gamma}\cong\mathbb{X}_{*}(T). Fix a Borel subgroup BB containing TT and let NN be the unipotent radical of BB. The Iwasawa decomposition of LGLG gives

GrG=LNΛγL+G/L+G=λ𝕏(T)LNλ(t)L+G/L+G.\textup{Gr}_{G}=LN\cdot\Lambda_{\gamma}L^{+}G/L^{+}G=\sqcup_{\lambda\in\mathbb{X}_{*}(T)}LN\cdot\lambda(t)L^{+}G/L^{+}G.

Let X:=𝒳γ(LNL+G/L+G)𝒳γX:=\mathcal{X}_{\gamma}\cap(LN\cdot L^{+}G/L^{+}G)\subset\mathcal{X}_{\gamma}. It is enough to show that XX lies in some affine Schubert variety Grλ\textup{Gr}_{\leq\lambda}, for then its closure Z:=X¯Z:=\overline{X} in Grλ\textup{Gr}_{\leq\lambda} satisfies the condition in (1). For later use, we note that the translations X\ell\cdot X for Λγ\ell\in\Lambda_{\gamma} are disjoint and cover 𝒳γ\mathcal{X}_{\gamma}.

Fixing an ordering of the positive roots of TT with respect to BB, we may write an element uN(F)u\in N(F) uniquely as

(18) u=α>0xα(cα)u=\prod_{\alpha>0}x_{\alpha}(c_{\alpha})

where cαFc_{\alpha}\in F and xα:𝔾aNx_{\alpha}:\mathbb{G}_{a}\to N is the root group corresponding to α\alpha. To show that XX is contained in an affine Schubert variety, it suffices to give a lower bound for the valuations of cαc_{\alpha} appearing in (18) for any uN(F)u\in N(F) such that [u]X(k)[u]\in X(k). We may expand Ad(u1)γ𝔟(F)\textup{Ad}(u^{-1})\gamma\in\mathfrak{b}(F) in terms of the root decomposition

Ad(u1)γ=γ+α>0Pα(γ;c)eα\textup{Ad}(u^{-1})\gamma=\gamma+\sum_{\alpha>0}P_{\alpha}(\gamma;c)e_{\alpha}

where eα𝔤αe_{\alpha}\in\mathfrak{g}_{\alpha} is a fixed basis for each root space and Pα(γ;c)P_{\alpha}(\gamma;c) is an FF-valued polynomial function in {cα}α>0\{c_{\alpha}\}_{\alpha>0} and linear in γ\gamma. Induction on the height of α\alpha shows that Pα(γ;c)P_{\alpha}(\gamma;c) takes the following form

(19) Pα(γ;c)=α,γcα+β<αβ,γPαβ(c)P_{\alpha}(\gamma;c)=\langle{\alpha,\gamma}\rangle c_{\alpha}+\sum_{\beta<\alpha}\langle{\beta,\gamma}\rangle P^{\beta}_{\alpha}(c)

where Pαβ(c)P^{\beta}_{\alpha}(c) is a polynomial involving only {cα}α<α\{c_{\alpha^{\prime}}\}_{\alpha^{\prime}<\alpha}, and homogeneous of degree α\alpha (we define degcα:=α𝕏(T)\deg c_{\alpha^{\prime}}:=\alpha^{\prime}\in\mathbb{X}^{*}(T)).

Let n=maxα>0{valFα,γ}n=\max_{\alpha>0}\{\textup{val}_{F}\langle{\alpha,\gamma}\rangle\}. This is finite because γ\gamma is regular semisimple. If [u]X(k)[u]\in X(k), i.e., Ad(u1)γ𝔤(𝒪F)\textup{Ad}(u^{-1})\gamma\in\mathfrak{g}(\mathcal{O}_{F}), then induction on the height of α\alpha shows that

valF(cα)(2ht(α)1)n,\textup{val}_{F}(c_{\alpha})\geq-(2\textup{ht}(\alpha)-1)n,

which gives the desired lower bound and shows that XX lies in an affine Schubert variety.

2.5.4. Proof of (1) in the general case

In the general case, we give a simplified argument compared to the original one in [KL], following the same idea. We make a base change to F=k((t1/m))F^{\prime}=k(\!(t^{1/m})\!) over which γ\gamma can be conjugated into a split torus. Let GrG\textup{Gr}^{\prime}_{G} be the affine Grassmannian of GG defined using the field FF^{\prime} in place of FF (so that GrG(k)=G(F)/G(𝒪F)\textup{Gr}^{\prime}_{G}(k)=G(F^{\prime})/G(\mathcal{O}^{\prime}_{F})), and let 𝒳γGrG\mathcal{X}^{\prime}_{\gamma}\subset\textup{Gr}^{\prime}_{G} be the corresponding affine Springer fiber. Then both GrG\textup{Gr}^{\prime}_{G} and 𝒳γ\mathcal{X}^{\prime}_{\gamma} carry an action of Γ:=Gal(F/F)μm\Gamma:=\textup{Gal}(F^{\prime}/F)\cong\mu_{m} induced from its action on FF^{\prime}, and we have a closed embedding 𝒳γ(𝒳γ)Γ\mathcal{X}_{\gamma}\hookrightarrow(\mathcal{X}^{\prime}_{\gamma})^{\Gamma}.

Let Λγ=𝕏(GγF)Gγ(F)\Lambda^{\prime}_{\gamma}=\mathbb{X}_{*}(G_{\gamma}\otimes F^{\prime})\hookrightarrow G_{\gamma}(F^{\prime}) be the lattice constructed using the field FF^{\prime}. There is an action of Γ\Gamma on 𝕏(GγF)\mathbb{X}_{*}(G_{\gamma}\otimes F^{\prime}) with fixed points 𝕏(Gγ)Λγ\mathbb{X}_{*}(G_{\gamma})\cong\Lambda_{\gamma}. This action induces an action of Γ\Gamma on Λγ\Lambda^{\prime}_{\gamma}, but it does not respect the embedding 𝕏(GγF)Gγ(F)\mathbb{X}_{*}(G_{\gamma}\otimes F^{\prime})\hookrightarrow G_{\gamma}(F^{\prime}). The fixed points (Λγ)Γ=𝕏(GγF)Γ(\Lambda^{\prime}_{\gamma})^{\Gamma}=\mathbb{X}_{*}(G_{\gamma}\otimes F^{\prime})^{\Gamma} may not lie in Gγ(F)G_{\gamma}(F), however it always contains Λγ\Lambda_{\gamma} with finite index.

From the proof in the split case in §2.5.3 we have a finite type locally closed subscheme X𝒳γX^{\prime}\subset\mathcal{X}^{\prime}_{\gamma} coming from the Iwasawa decomposition, such that 𝒳γ\mathcal{X}^{\prime}_{\gamma} can be decomposed as the disjoint union ΛγX\sqcup_{\ell\in\Lambda^{\prime}_{\gamma}}\ell\cdot X^{\prime} (not as schemes but as constructible sets). We identify Λγ\ell\in\Lambda^{\prime}_{\gamma} with an element in 𝕏(GγF)\mathbb{X}_{*}(G_{\gamma}\otimes F^{\prime}) then σ()\sigma(\ell) makes sense for σΓ\sigma\in\Gamma. One checks that σ(X)=σ()X\sigma(\ell\cdot X^{\prime})=\sigma(\ell)\cdot X^{\prime} for Λγ\ell\in\Lambda^{\prime}_{\gamma} even though σ\sigma does not respect the embedding ΛγGγ(F)\Lambda^{\prime}_{\gamma}\to G_{\gamma}(F^{\prime}). Therefore (𝒳γ)Γ(Λγ)ΓX(\mathcal{X}^{\prime}_{\gamma})^{\Gamma}\subset(\Lambda^{\prime}_{\gamma})^{\Gamma}\cdot X^{\prime}. Choosing representatives CC for the finite coset space (Λγ)Γ/Λγ(\Lambda^{\prime}_{\gamma})^{\Gamma}/\Lambda_{\gamma}, we see that (𝒳γ)Γ(\mathcal{X}^{\prime}_{\gamma})^{\Gamma} is contained in Λγ(CX)\Lambda_{\gamma}\cdot(C\cdot X^{\prime}). Since CXC\cdot X^{\prime} is of finite type, Z=CX¯(𝒳γ)ΓZ^{\prime}=\overline{C\cdot X^{\prime}}\cap(\mathcal{X}^{\prime}_{\gamma})^{\Gamma} is a projective subscheme of 𝒳γ\mathcal{X}^{\prime}_{\gamma} whose Λγ\Lambda_{\gamma}-translations cover (𝒳γ)Γ(\mathcal{X}^{\prime}_{\gamma})^{\Gamma}. Finally the projective subscheme Z=Z𝒳γZ=Z^{\prime}\cap\mathcal{X}_{\gamma} of 𝒳γ\mathcal{X}_{\gamma} satisfies the requirement of (1). This finishes the proof of Theorem 17. ∎

2.5.5. Reduction to Levi

The proof of Theorem 17 in the split case in §2.5.3 gives more information. In the Iwasawa decomposition, let SλGrGS_{\lambda}\subset\textup{Gr}_{G} be the LNLN-orbit of λ(t)\lambda(t), for λ𝕏(T)\lambda\in\mathbb{X}_{*}(T). This is called a semi-infinite orbit, because it has infinite dimension and also has infinite codimension in GrG\textup{Gr}_{G}. Let Cλ:=𝒳γSλC_{\lambda}:=\mathcal{X}_{\gamma}\cap S_{\lambda}, then Cλ=λ(t)XC_{\lambda}=\lambda(t)\cdot X in the notation of §2.5.3. The formula (19) implies that CλC_{\lambda}\neq\varnothing (or equivalently XX\neq\varnothing, or 𝒳γ\mathcal{X}_{\gamma}\neq\varnothing) if and only if α,γ𝒪F\langle{\alpha,\gamma}\rangle\in\mathcal{O}_{F} for all roots α\alpha, and if so, CλC_{\lambda} is isomorphic to an almost affine space (namely an iterated 𝔸1\mathbb{A}^{1}-bundle) of dimension

dimCλ=α>0valF(α,γ)=12valFΔ(γ)\dim C_{\lambda}=\sum_{\alpha>0}\textup{val}_{F}(\langle{\alpha,\gamma}\rangle)=\frac{1}{2}\textup{val}_{F}\Delta(\gamma)

Here Δ(γ)\Delta(\gamma) is the determinant of the adjoint action of γ\gamma on 𝔤(F)/𝔱(F)\mathfrak{g}(F)/\mathfrak{t}(F). Therefore 𝒳γ\mathcal{X}_{\gamma} can be decomposed into almost affine spaces of the same dimension indexed by 𝕏(T)\mathbb{X}_{*}(T). However, this decomposition is not a stratification: the closure C¯λ\overline{C}_{\lambda} of CλC_{\lambda} will intersect other CλC_{\lambda^{\prime}} but certainly not a union of such CλC_{\lambda^{\prime}}’s.

The decomposition 𝒳γ=Cλ\mathcal{X}_{\gamma}=\sqcup C_{\lambda} in the split case has a generalization. Suppose PP is a parabolic subgroup of GG with unipotent radical NPN_{P} and a Levi subgroup MPM_{P}. Let 𝔪P=LieMP\mathfrak{m}_{P}=\textup{Lie}\ M_{P} and suppose γ𝔪P(F)\gamma\in\mathfrak{m}_{P}(F) is regular semisimple as an element in 𝔤(F)\mathfrak{g}(F). Using the generalized Iwasawa decomposition G(F)=NP(F)MP(F)G(𝒪F)G(F)=N_{P}(F)M_{P}(F)G(\mathcal{O}_{F}), there is a well-defined map GrG(k)GrMP(k)\textup{Gr}_{G}(k)\to\textup{Gr}_{M_{P}}(k) sending nmG(𝒪F)nmG(\mathcal{O}_{F}) to mMP(𝒪F)mM_{P}(\mathcal{O}_{F}), for nNP(F)n\in N_{P}(F) and mMP(F)m\in M_{P}(F). However this map does not give a map of ind-schemes. Nevertheless the fibers of this map have natural structure of infinite dimensional affine spaces. Restricting this map to 𝒳γ\mathcal{X}_{\gamma} we get

τ:𝒳γ(k)𝒳γMP(k)\tau:\mathcal{X}_{\gamma}(k)\to\mathcal{X}^{M_{P}}_{\gamma}(k)

where 𝒳γMPGrMP\mathcal{X}^{M_{P}}_{\gamma}\subset\textup{Gr}_{M_{P}} is the affine Springer fiber for γ\gamma and the group MPM_{P}. The fibers of τ\tau, if non-empty, are almost affine spaces of dimension 12valFΔMPG(γ)\frac{1}{2}\textup{val}_{F}\Delta^{G}_{M_{P}}(\gamma), where

(20) ΔMPG(γ):=det(ad(γ)|𝔤(F)/𝔪P(F)).\Delta^{G}_{M_{P}}(\gamma):=\det(\textup{ad}(\gamma)|\mathfrak{g}(F)/\mathfrak{m}_{P}(F)).

Assume TMPT\subset M_{P}, then the connected components of GrMP\textup{Gr}_{M_{P}} are indexed by 𝕏(T)/RMP\mathbb{X}_{*}(T)/R^{\vee}_{M_{P}} where RMPR^{\vee}_{M_{P}} is the coroot lattice of MPM_{P}. If we decompose 𝒳γMP\mathcal{X}^{M_{P}}_{\gamma} into connected components 𝒳γMP(λ)\mathcal{X}^{M_{P}}_{\gamma}(\lambda) for 𝕏(T)/RMP\mathbb{X}_{*}(T)/R^{\vee}_{M_{P}} and taking their preimages under τ\tau, we get a decomposition 𝒳γ=𝒳γ,λ\mathcal{X}_{\gamma}=\sqcup\mathcal{X}_{\gamma,\lambda} into locally closed sub-ind-schemes indexed also by λ𝕏(T)/RMP\lambda\in\mathbb{X}_{*}(T)/R^{\vee}_{M_{P}}. One can show that 𝒳γ,λ\mathcal{X}_{\gamma,\lambda}, if non-empty, is an almost affine space bundle over 𝒳γMP(λ)\mathcal{X}^{M_{P}}_{\gamma}(\lambda) with fiber dimension 12valFΔMPG(γ)\frac{1}{2}\textup{val}_{F}\Delta^{G}_{M_{P}}(\gamma).

2.5.6. Connectivity and equidimensionality

When GG is simply-connected, Fl is connected, and in this case the affine Springer fiber 𝒴γ\mathcal{Y}_{\gamma} is also connected. See [KL, §4, Lemma 2]. As a consequence, when GG is simply-connected, 𝒳𝐏,γ\mathcal{X}_{\mathbf{P},\gamma} is connected for all parahoric 𝐏\mathbf{P} because the natural projection 𝒴γ𝒳𝐏,γ\mathcal{Y}_{\gamma}\to\mathcal{X}_{\mathbf{P},\gamma} is surjective.

In [KL], it is also shown that 𝒴γ\mathcal{Y}_{\gamma} is equidimensional. The argument there is similar to Spaltenstein’s the proof of the connectivity and equidimensionality for Springer fibers in [Spa77].

2.5.7. The dimension formula

By Theorem 17, the dimension of 𝒳γ\mathcal{X}_{\gamma} is well-defined, and is the dimension of Lγ\𝒳γL_{\gamma}\backslash\mathcal{X}_{\gamma} as an algebraic space. To state a formula for dim𝒳γ\dim\mathcal{X}_{\gamma}, we need some more notation.

Consider the adjoint action ad(γ):𝔤(F)𝔤(F)\textup{ad}(\gamma):\mathfrak{g}(F)\to\mathfrak{g}(F). The kernel of this map is 𝔤γ(F)\mathfrak{g}_{\gamma}(F), and the induced endomorphism ad(γ)¯\overline{\textup{ad}(\gamma)} on 𝔤(F)/𝔤γ(F)\mathfrak{g}(F)/\mathfrak{g}_{\gamma}(F) is invertible. Let Δ(γ)F×\Delta(\gamma)\in F^{\times} be the determinant of ad(γ)¯\overline{\textup{ad}(\gamma)}. This is consistent with our earlier definition of Δ(γ)\Delta(\gamma) in the case γ\gamma lies in a split torus 𝔱(F)\mathfrak{t}(F).

On the other hand, recall 𝕏(Gγ)\mathbb{X}_{*}(G_{\gamma}) is the group of FF-rational cocharacters of GγG_{\gamma}, which is also the rank of the maximal FF-split subtorus of GγG_{\gamma}. Let

c(γ)=rrk𝕏(Gγ).c(\gamma)=r-\textup{rk}_{\mathbb{Z}}\mathbb{X}_{*}(G_{\gamma}).

Then c(γ)c(\gamma) is also the rank of the maximal FF-anisotropic subtorus GγG_{\gamma}.

21 Theorem (Bezrukavnikov [Bez], conjectured by Kazhdan-Lusztig [KL]).

Let γ𝔤(F)\gamma\in\mathfrak{g}(F) be a regular semisimple element. Then we have

(22) dim𝒳γ=12(valFΔ(γ)c(γ)).\dim\mathcal{X}_{\gamma}=\frac{1}{2}(\textup{val}_{F}\Delta(\gamma)-c(\gamma)).

2.5.8. Sketch of proof

A key role in the proof is played by the notion of regular points of 𝒳γ\mathcal{X}_{\gamma}. We have an evaluation map ev:𝒳γ[𝔤/G]\textup{ev}:\mathcal{X}_{\gamma}\to[\mathfrak{g}/G] sending [g]𝒳γGrG[g]\in\mathcal{X}_{\gamma}\subset\textup{Gr}_{G} to the reduction of Ad(g1)γ\textup{Ad}(g^{-1})\gamma modulo tt, which is well-defined up to the adjoint action by GG. We say [g]𝒳γ[g]\in\mathcal{X}_{\gamma} is a regular point if ev([g])\textup{ev}([g]) lies in the open substack [𝔤reg/G][\mathfrak{g}^{\textup{reg}}/G] of [𝔤/G][\mathfrak{g}/G]. Let 𝒳γreg𝒳γ\mathcal{X}^{\textup{reg}}_{\gamma}\subset\mathcal{X}_{\gamma} be the open sub-ind-scheme of 𝒳γ\mathcal{X}_{\gamma} consisting of regular points. It can be shown that 𝒳γreg\mathcal{X}^{\textup{reg}}_{\gamma} is non-empty. Denote the preimage of 𝒳γreg\mathcal{X}^{\textup{reg}}_{\gamma} in 𝒴γ\mathcal{Y}_{\gamma} by 𝒴γreg\mathcal{Y}^{\textup{reg}}_{\gamma}, then the projection map 𝒴γreg𝒳γreg\mathcal{Y}^{\textup{reg}}_{\gamma}\to\mathcal{X}^{\textup{reg}}_{\gamma} is an isomorphism. Since 𝒴γreg\mathcal{Y}^{\textup{reg}}_{\gamma} is equidimensional as mentioned in §2.5.6, we see that dim𝒳γreg=dim𝒴γreg=dim𝒴γ\dim\mathcal{X}^{\textup{reg}}_{\gamma}=\dim\mathcal{Y}^{\textup{reg}}_{\gamma}=\dim\mathcal{Y}_{\gamma}. Of course we have dim𝒳γdim𝒴γ\dim\mathcal{X}_{\gamma}\leq\dim\mathcal{Y}_{\gamma}, therefore we must have dim𝒳γreg=dim𝒳γ=dim𝒴γ\dim\mathcal{X}^{\textup{reg}}_{\gamma}=\dim\mathcal{X}_{\gamma}=\dim\mathcal{Y}_{\gamma}. It remains to calculate the dimension of 𝒳γreg\mathcal{X}^{\textup{reg}}_{\gamma}.

Recall we have defined the local Picard group PaP_{a} in §2.3.6. The action of PaP_{a} on 𝒳γ\mathcal{X}_{\gamma} preserves 𝒳γreg\mathcal{X}^{\textup{reg}}_{\gamma}, and in fact 𝒳γreg\mathcal{X}^{\textup{reg}}_{\gamma} is a torsor under PaP_{a}. Therefore it suffices to compute the dimension of Pa=LGγ/L+JaP_{a}=LG_{\gamma}/L^{+}J_{a}.

Consider the projection χ𝔱:𝔱𝔠=𝔱W\chi_{\mathfrak{t}}:\mathfrak{t}\to\mathfrak{c}=\mathfrak{t}\sslash W. Pullback along a=χ(γ):Spec𝒪F𝔠a=\chi(\gamma):\textup{Spec}\ \mathcal{O}_{F}\to\mathfrak{c} we get a finite morphism χa:χ𝔱1(a)=SpecASpec𝒪F\chi_{a}:\chi^{-1}_{\mathfrak{t}}(a)=\textup{Spec}\ A\to\textup{Spec}\ \mathcal{O}_{F}, for a finite flat 𝒪F\mathcal{O}_{F}-algebra AA. There is a close relationship between the group scheme JaJ_{a} and Ja=(Res𝒪FA(TkA))WJ^{\prime}_{a}=(\textup{Res}^{A}_{\mathcal{O}_{F}}(T\otimes_{k}A))^{W}, where the Weyl group WW acts diagonally on both TT and AA. In fact JaJ_{a} and JaJ^{\prime}_{a} are equal up to connected components in their special fibers. In particular, dimPa=dimPa\dim P_{a}=\dim P^{\prime}_{a}, where Pa=LGγ/L+JaP^{\prime}_{a}=LG_{\gamma}/L^{+}J^{\prime}_{a}. It is not hard to see that

dimPa=dim(𝔱A~)Wdim(𝔱A)W=dim(𝔱(A~/A))W\dim P^{\prime}_{a}=\dim(\mathfrak{t}\otimes\widetilde{A})^{W}-\dim(\mathfrak{t}\otimes A)^{W}=\dim(\mathfrak{t}\otimes(\widetilde{A}/A))^{W}

where A~\widetilde{A} is the normalization of AA. From this one deduces the dimension formula (22). ∎

23 Remark.

Using the relation between affine Springer fibers and Hitchin fibers, Ngô [NgoFL, Cor 4.16.2] showed that 𝒳γreg\mathcal{X}^{\textup{reg}}_{\gamma} is in fact dense in 𝒳γ\mathcal{X}_{\gamma}. In particular, each irreducible component of 𝒳γ\mathcal{X}_{\gamma} is a rational variety. However, this rationality property is false for affine Springer fibers in more general affine partial flag varieties, as we saw in Bernstein-Kazhdan’s example in §2.4.5.

2.5.9. Purity

It was conjectured by Goresky, Kottwitz and MacPherson [GKM] that the cohomology of affine Springer fibers should be pure (in the sense of Frobenius weights if k=𝔽¯pk=\overline{\mathbb{F}}_{p}, or in the sense of Hodge structures if k=k=\mathbb{C}). The purity of affine Springer fibers would allow the authors of [GKM] to prove the Fundamental Lemma for unramified elements using localization techniques in equivariant cohomology. This purity conjecture is still open in general. In [GKMPurity], a class of affine Springer fibers called equivalued were shown to be pure.

2.5.10. Invariance under perturbation

Suppose a,a𝔠(𝒪F)𝔠rs(F)a,a^{\prime}\in\mathfrak{c}(\mathcal{O}_{F})\cap\mathfrak{c}^{\textup{rs}}(F). We say aamodtNa\equiv a^{\prime}\mod t^{N} if aa and aa^{\prime} have the same image under the map 𝔠(𝒪F)𝔠(𝒪F/tN)\mathfrak{c}(\mathcal{O}_{F})\to\mathfrak{c}(\mathcal{O}_{F}/t^{N}). In [NgoFL, Prop 3.5.1], it is shown that for fixed a𝔠(𝒪F)𝔠rs(F)a\in\mathfrak{c}(\mathcal{O}_{F})\cap\mathfrak{c}^{\textup{rs}}(F), there exists some NN (depending on aa) such that whenever aamodtNa^{\prime}\equiv a\mod t^{N}, we have isomorphisms

𝒳~a𝒳~a and PaPa\widetilde{\mathcal{X}}_{a}\cong\widetilde{\mathcal{X}}_{a^{\prime}}\textup{ and }P_{a}\cong P_{a^{\prime}}

in a way compatible with the actions. Therefore, we may say that 𝒳a\mathcal{X}_{a} varies locally constantly with aa under the tt-adic topology on 𝔠(𝒪F)\mathfrak{c}(\mathcal{O}_{F}).

For example, consider the case G=GLnG=\textup{GL}_{n} and let a𝔠(𝒪F)a\in\mathfrak{c}(\mathcal{O}_{F}) correspond to a characteristic polynomial P(x)=xn+a1xn1++anP(x)=x^{n}+a_{1}x^{n-1}+\cdots+a_{n} whose roots are in 𝒪F\mathcal{O}_{F} and are distinct modulo tt. Then for any aamodta^{\prime}\equiv a\mod t, the characteristic polynomial of aa^{\prime} also has distinct roots modulo tt. In this case, 𝒳~a\widetilde{\mathcal{X}}_{a} and 𝒳~a\widetilde{\mathcal{X}}_{a^{\prime}} are both torsors under PaPaGrTP_{a}\cong P_{a^{\prime}}\cong\textup{Gr}_{T}.

2.6. Affine Springer representations

In this subsection we introduce an analog of Springer’s WW-action on H(e)\textup{H}^{*}({\mathcal{B}_{e}}) in the affine situation.

2.6.1. The affine Weyl group

We view WW as a group of automorphisms of the cocharacter lattice 𝕏(T)\mathbb{X}_{*}(T), where TT is a fixed maximal torus of GG. The extended affine Weyl group W~\widetilde{W} is the semidirect product

W~=𝕏(T)W.\widetilde{W}=\mathbb{X}_{*}(T)\rtimes W.

When GG is simply-connected, so that 𝕏(T)\mathbb{X}_{*}(T) is spanned by coroots, W~\widetilde{W} is a Coxeter group with simple reflections {s0,s1,,sr}\{s_{0},s_{1},\cdots,s_{r}\} in bijection with the nodes of the extended Dynkin diagram of GG. In general, W~\widetilde{W} is a semidirect product of the affine Weyl group Waff=(R)WW_{\textup{aff}}=(\mathbb{Z}R^{\vee})\rtimes W (where R\mathbb{Z}R^{\vee} is the coroot lattice), which is a Coxeter group, and an abelian group Ω𝕏(T)/R\Omega\cong\mathbb{X}_{*}(T)/\mathbb{Z}R^{\vee}. The group W~\widetilde{W} naturally acts on the affine space 𝕏(T)\mathbb{X}_{*}(T)_{\mathbb{R}} by affine transformations, where 𝕏(T)\mathbb{X}_{*}(T) acts by translations.

24 Theorem (Lusztig [L96], Sage [Sage]).

There is a canonical action of W~\widetilde{W} on H(𝒴γ)\textup{H}_{*}({\mathcal{Y}_{\gamma}}).

Since 𝒴γ\mathcal{Y}_{\gamma} is not of finite type, the \ell-adic homology H(𝒴γ)\textup{H}_{*}({\mathcal{Y}_{\gamma}}) is understood as the inductive limit limnH(𝒴γ,n)\varinjlim_{n}\textup{H}_{*}({\mathcal{Y}_{\gamma,n}}), whenever we present 𝒴γ\mathcal{Y}_{\gamma} as a union of projective subschemes 𝒴γ,n\mathcal{Y}_{\gamma,n}.

2.6.2. Sketch of the construction of the W~\widetilde{W}-action

We consider only the case GG is simply-connected so that W~=Waff\widetilde{W}=W_{\textup{aff}} is generated by affine simple reflections s0,,srs_{0},\cdots,s_{r}. For each parahoric subgroup 𝐏LG\mathbf{P}\subset LG we have a corresponding affine Springer fiber 𝒳~𝐏,γ\widetilde{\mathcal{X}}_{\mathbf{P},\gamma}. For 𝐏\mathbf{P} containing a fixed Iwahori subgroup 𝐈\mathbf{I}, we have a projection π𝐏,γ:𝒴~γ𝒳~𝐏,γ\pi_{\mathbf{P},\gamma}:\widetilde{\mathcal{Y}}_{\gamma}\to\widetilde{\mathcal{X}}_{\mathbf{P},\gamma}.

Let L𝐏L_{\mathbf{P}} be the Levi quotient of 𝐏\mathbf{P} and 𝔩𝐏=LieL𝐏\mathfrak{l}_{\mathbf{P}}=\textup{Lie}\ L_{\mathbf{P}}. We have an evaluation map ev𝐏,γ:𝒳~𝐏,γ[𝔩𝐏/L𝐏]\textup{ev}_{\mathbf{P},\gamma}:\widetilde{\mathcal{X}}_{\mathbf{P},\gamma}\to[\mathfrak{l}_{\mathbf{P}}/L_{\mathbf{P}}] defined as follows. For [g]Fl𝐏[g]\in\textup{Fl}_{\mathbf{P}} such that Ad(g1)γLie𝐏\textup{Ad}(g^{-1})\gamma\in\textup{Lie}\ \mathbf{P}, we send the coset [g]=g𝐏[g]=g\mathbf{P} to the image of Ad(g1)γ\textup{Ad}(g^{-1})\gamma under the projection Lie𝐏𝔩𝐏\textup{Lie}\ \mathbf{P}\to\mathfrak{l}_{\mathbf{P}}. This is well-defined up to the adjoint action of L𝐏L_{\mathbf{P}}. We have a Cartesian diagram

𝒴~γ\textstyle{\widetilde{\mathcal{Y}}_{\gamma}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π𝐏,γ\scriptstyle{\pi_{\mathbf{P},\gamma}}ev𝐈,γ\scriptstyle{\textup{ev}_{\mathbf{I},\gamma}}[𝔩~𝐏/L𝐏]\textstyle{[\widetilde{\mathfrak{l}}_{\mathbf{P}}/L_{\mathbf{P}}]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π𝔩𝐏\scriptstyle{\pi_{\mathfrak{l}_{\mathbf{P}}}}𝒳~𝐏,γ\textstyle{\widetilde{\mathcal{X}}_{\mathbf{P},\gamma}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ev𝐏,γ\scriptstyle{\textup{ev}_{\mathbf{P},\gamma}}[𝔩𝐏/L𝐏]\textstyle{[\mathfrak{l}_{\mathbf{P}}/L_{\mathbf{P}}]}

where π𝔩𝐏\pi_{\mathfrak{l}_{\mathbf{P}}} is the Grothendieck alteration for the reductive group 𝔩𝐏\mathfrak{l}_{\mathbf{P}}. By the Springer theory for L𝐏L_{\mathbf{P}}, we have a W(L𝐏)W(L_{\mathbf{P}})-action on the direct image complex 𝐑π𝔩𝐏,𝔻\mathbf{R}\pi_{\mathfrak{l}_{\mathbf{P},*}}\mathbb{D} (where 𝔻\mathbb{D} stands for the dualizing complex for [𝔩~𝐏/L𝐏][\widetilde{\mathfrak{l}}_{\mathbf{P}}/L_{\mathbf{P}}]). By proper base change, we get an action of W(L𝐏)W(L_{\mathbf{P}}) on 𝐑π𝐏,γ,𝔻𝒴~γ\mathbf{R}\pi_{\mathbf{P},\gamma,*}\mathbb{D}_{\widetilde{\mathcal{Y}}_{\gamma}}, and hence on H(𝒴γ)\textup{H}_{*}({\mathcal{Y}_{\gamma}}). Taking a standard parahoric 𝐏\mathbf{P} corresponding to the ii-th node in the extended Dynkin diagram, then W(L𝐏)=siW(L_{\mathbf{P}})=\langle{s_{i}}\rangle, and we get an involution sis_{i} acting on H(𝒴γ)\textup{H}_{*}({\mathcal{Y}_{\gamma}}). To check the braid relation between sis_{i} and sjs_{j} for neighboring nodes ii and jj, we may choose a standard parahoric 𝐏\mathbf{P} such that W(L𝐏)=si,sjW(L_{\mathbf{P}})=\langle{s_{i},s_{j}}\rangle and the braid relation holds because W(L𝐏)W(L_{\mathbf{P}}) acts on H(𝒴γ)\textup{H}_{*}({\mathcal{Y}_{\gamma}}). This shows that W~\widetilde{W} acts on H(𝒴γ)\textup{H}_{*}({\mathcal{Y}_{\gamma}}). ∎

Despite the simplicity of the construction of the W~\widetilde{W}-action on the homology of affine Springer fibers, the calculation of these actions are quite difficult. One new feature here is that the action of W~\widetilde{W} on H(𝒴γ)\textup{H}_{*}({\mathcal{Y}_{\gamma}}) may not be semisimple, as we shall see in the next example.

2.6.3. An example in SL2\textup{SL}_{2}

Consider the case G=SL2G=\textup{SL}_{2} and the element γ=(0t2t0)\gamma=\left(\begin{array}[]{cc}0&t^{2}\\ t&0\end{array}\right). This is a subregular case. The affine Springer fiber 𝒴γ\mathcal{Y}_{\gamma} has two irreducible components C0C_{0} and C1C_{1} both isomorphic to 1\mathbb{P}^{1}. Here C0C_{0} parametrizes chains of lattices {Λi}\{\Lambda_{i}\} where Λ0\Lambda_{0} is the standard lattice 𝒪F2\mathcal{O}^{2}_{F} and Λ1Λ0\Lambda_{-1}\subset\Lambda_{0} is varying. The other component C1C_{1} parametrizes chains of lattices {Λi}\{\Lambda_{i}\} where Λ1=t𝒪F𝒪F\Lambda_{-1}=t\mathcal{O}_{F}\oplus\mathcal{O}_{F} and Λ0\Lambda_{0} is varying. The fundamental classes [C0],[C1][C_{0}],[C_{1}] give a natural basis for the top homology group H2(𝒴γ)\textup{H}_{2}({\mathcal{Y}_{\gamma}}). One can show that the action of the affine Weyl group W~=s0,s1\widetilde{W}=\langle{s_{0},s_{1}}\rangle on H2(𝒴γ)\textup{H}_{2}({\mathcal{Y}_{\gamma}}) takes the following form under the basis [C0][C_{0}] and [C1][C_{1}]:

s0=(1201);s1=(1021).s_{0}=\left(\begin{array}[]{cc}-1&2\\ 0&1\end{array}\right);\quad s_{1}=\left(\begin{array}[]{cc}1&0\\ 2&-1\end{array}\right).

We see that H2(𝒴γ)\textup{H}_{2}({\mathcal{Y}_{\gamma}}) is a nontrivial extension of the sign representation of W~\widetilde{W} by the trivial representation spanned by [C0]+[C1][C_{0}]+[C_{1}]. One can also canonically identify the W~\widetilde{W}-module H2(𝒴γ,)\textup{H}_{2}({\mathcal{Y}_{\gamma},\mathbb{Z}}) with the affine coroot lattice of the loop group LGLG.

2.7. Comments and generalizations

2.7.1. Relation with orbital integrals

As we will see in §3, the cohomology and point-counting of affine Springer fibers are closely related to orbital integrals on pp-adic groups G(𝔽q((t)))G(\mathbb{F}_{q}(\!(t)\!)).

2.7.2. Extended symmetry

The W~\widetilde{W}-action on H(𝒴γ)\textup{H}_{*}({\mathcal{Y}_{\gamma}}) can be extended to an action of the wreath product Sym(𝕏(T))W~\textup{Sym}(\mathbb{X}^{*}(T))\rtimes\widetilde{W}. For homogeneous affine Springer fibers 𝒴γ\mathcal{Y}_{\gamma} (those admitting a torus action coming from loop rotation), the equivariant cohomology group H𝔾m(𝒴γ)\textup{H}^{*}_{\mathbb{G}_{m}}(\mathcal{Y}_{\gamma}) admits an action of the graded double affine Hecke algebra, which is a deformation of Sym(𝕏(T))W~\textup{Sym}(\mathbb{X}^{*}(T))\rtimes\widetilde{W}. For details we refer to [OY]. Vasserot and Varagnolo [Vass] [VV] constructed an action of the double affine Hecke algebra on the KK-groups of affine Springer fibers.

2.7.3. The group version

Taking γG(F)\gamma\in G(F) instead of in 𝔤(F)\mathfrak{g}(F), one can similarly define the group version of affine Springer fibers, which we still denote by 𝒳~γ\widetilde{\mathcal{X}}_{\gamma} with reduced structure 𝒳γ\mathcal{X}_{\gamma}. For a kk-algebras RR, we have

(25) 𝒳~γ(R)={[g]GrG(R)|g1γgL+G(R)}.\widetilde{\mathcal{X}}_{\gamma}(R)=\{[g]\in\textup{Gr}_{G}(R)|g^{-1}\gamma g\in L^{+}G(R)\}.

However, in the group version, the definition above admits an interesting generalization. Recall the L+GL^{+}G-double cosets in LGLG are indexed by dominant cocharacters λ𝕏(T)+\lambda\in\mathbb{X}_{*}(T)^{+}. For λ𝕏(T)+\lambda\in\mathbb{X}_{*}(T)^{+} we denote the corresponding double coset by (LG)λ(LG)_{\lambda}, which is the preimage of the Schubert stratum GrG,λ\textup{Gr}_{G,\lambda} under the projection LGGrGLG\to\textup{Gr}_{G}. Similarly we may define (LG)λ(LG)_{\leq\lambda} to be the preimage of the closure GrG,λ\textup{Gr}_{G,\leq\lambda} of GrG,λ\textup{Gr}_{G,\lambda}. One can replace the condition g1γgL+G(R)g^{-1}\gamma g\in L^{+}G(R) in (25) by g1γg(LG)λ(R)g^{-1}\gamma g\in(LG)_{\lambda}(R) or g1γg(LG)λ(R)g^{-1}\gamma g\in(LG)_{\leq\lambda}(R), and take reduced structures to obtain reduced generalized affine Springer fibers 𝒳λ,γ\mathcal{X}_{\lambda,\gamma} and 𝒳λ,γ\mathcal{X}_{\leq\lambda,\gamma}. We have an open embedding 𝒳λ,γ𝒳λ,γ\mathcal{X}_{\lambda,\gamma}\hookrightarrow\mathcal{X}_{\leq\lambda,\gamma}, whose complement is the union of 𝒳μ,γ\mathcal{X}_{\mu,\gamma} for dominant coweights μλ\mu\leq\lambda. The motivation for introducing 𝒳λ,γ\mathcal{X}_{\leq\lambda,\gamma} is to give geometric interpretation of orbital integrals of spherical Hecke functions on G(F)G(F).

A.Bouthier has established the fundamental geometric properties of 𝒳λ,γ\mathcal{X}_{\lambda,\gamma}, parallel to Theorem 17 and Theorem 21.

26 Theorem (Bouthier [Bouthier]).

Let γG(F)\gamma\in G(F) be regular semisimple, and let λ𝕏(T)+\lambda\in\mathbb{X}_{*}(T)^{+}.

  1. (1)

    The generalized affine Springer fiber 𝒳λ,γ\mathcal{X}_{\lambda,\gamma} is non-empty if and only if [νγ]λ[\nu_{\gamma}]\leq\lambda, where [νγ]𝕏(T)+[\nu_{\gamma}]\in\mathbb{X}_{*}(T)^{+}_{\mathbb{Q}} is the Newton point of γ\gamma, see [KV, §2].

  2. (2)

    The ind-scheme 𝒳λ,γ\mathcal{X}_{\lambda,\gamma} is locally of finite type.

  3. (3)

    We have

    dim𝒳λ,γ=ρ,λ+12(valFΔ(γ)c(γ))\dim\mathcal{X}_{\lambda,\gamma}=\langle{\rho,\lambda}\rangle+\frac{1}{2}(\textup{val}_{F}\Delta(\gamma)-c(\gamma))

    where ρ𝕏(T)\rho\in\mathbb{X}^{*}(T)_{\mathbb{Q}} is half the sum of positive roots, and Δ(γ)\Delta(\gamma) and c(γ)c(\gamma) are defined similarly as in the Lie algebra situation.

The proof of this theorem uses the theory of Vinberg semigroups, which is a kind of compactification of GG.

2.7.4.

In [KV], Kottwitz and Viehmann defined two generalizations of affine Springer fibers for elements γ\gamma in the Lie algebra 𝔤(F)\mathfrak{g}(F).

2.7.5.

As an analog of Hessenberg varieties, one can also consider the following situation. Let (ρ,V)(\rho,V) be a linear representation of a reductive group GG over kk. Let ΛVF\Lambda\subset V\otimes F be an 𝒪F\mathcal{O}_{F}-lattice stable under G(𝒪F)G(\mathcal{O}_{F}). For vVFv\in V\otimes F we may define a sub-ind-scheme 𝒳~v\widetilde{\mathcal{X}}_{v} of GrG\textup{Gr}_{G}

𝒳~Λ,v(R)={[g]GrG(R)|ρ(g1)vΛ^kR}.\widetilde{\mathcal{X}}_{\Lambda,v}(R)=\{[g]\in\textup{Gr}_{G}(R)|\rho(g^{-1})v\in\Lambda\widehat{\otimes}_{k}R\}.

Let 𝒳Λ,v\mathcal{X}_{\Lambda,v} be the reduced structure of 𝒳~Λ,v\widetilde{\mathcal{X}}_{\Lambda,v}. The cohomology of these ind-schemes are related to orbital integrals that appear in relative trace formulae.

2.8. Exercises

2.8.1.

Let G=SO(V,q)G=\textup{SO}(V,q) for some vector space VV over kk equipped with a quadratic form qq. Give an interpretation of the parahoric subgroups and affine partial flag varieties of LGLG in terms of self-dual lattice chains in VkFV\otimes_{k}F, in the same style as in §2.1.6.

2.8.2.

Verify the descriptions of the affine Springer fibers for G=SL2G=\textup{SL}_{2} given in §2.2.4 and §2.2.5.

2.8.3.

Let G=SL2G=\textup{SL}_{2} and γ=(0tn10)𝔤(F)\gamma=\left(\begin{array}[]{cc}0&t^{n}\\ 1&0\end{array}\right)\in\mathfrak{g}(F). Describe 𝒴γ\mathcal{Y}_{\gamma}.

2.8.4.

Let G=SL2G=\textup{SL}_{2} and γ=(0tn10)𝔤(F)\gamma=\left(\begin{array}[]{cc}0&t^{n}\\ 1&0\end{array}\right)\in\mathfrak{g}(F). Construct a nontrivial 𝔾m\mathbb{G}_{m}-action on 𝒳γ\mathcal{X}_{\gamma} involving loop rotations (i.e., the action scales tt) and determine its fixed points.

2.8.5.

Let G=SL2G=\textup{SL}_{2} and γ=(tn00tn)𝔤(F)\gamma=\left(\begin{array}[]{cc}t^{n}&0\\ 0&-t^{n}\end{array}\right)\in\mathfrak{g}(F). Let TGT\subset G be the diagonal torus, then Gγ=TkFG_{\gamma}=T\otimes_{k}F. What is the regular locus 𝒳γreg\mathcal{X}^{\textup{reg}}_{\gamma} (see §2.5.8)? Study the L+TL^{+}T-orbits on 𝒳γ\mathcal{X}_{\gamma}.

2.8.6.

In the setup of §2.3.3, show that the action of Λγ\Lambda_{\gamma} on GrG\textup{Gr}_{G} is free, which implies that its action on 𝒳γ\mathcal{X}_{\gamma} is free. Show also that the permutation action of Λγ\Lambda_{\gamma} on the set of irreducible components of 𝒳γ\mathcal{X}_{\gamma} is free.

2.8.7.

For G=GLnG=\textup{GL}_{n}, let LGL\subset G be the Levi subgroup consisting of block diagonal matrices with sizes of blocks n1,,nsn_{1},\cdots,n_{s}, ini=n\sum_{i}n_{i}=n. Let γ=(γ1,,γs)𝔩(F)\gamma=(\gamma_{1},\cdots,\gamma_{s})\in\mathfrak{l}(F) be regular semisimple as an element in 𝔤(F)\mathfrak{g}(F). What is the invariant ΔLG(γ)\Delta^{G}_{L}(\gamma) (see (20)) in terms of familiar invariants of the characteristic polynomials of the γi\gamma_{i}?

2.8.8.

Let G=SL3G=\textup{SL}_{3} and γ=diag(x1t,x2t,x3t)𝔤(F)\gamma=\textup{diag}(x_{1}t,x_{2}t,x_{3}t)\in\mathfrak{g}(F), with xikx_{i}\in k pairwise distinct and x1+x2+x3=0x_{1}+x_{2}+x_{3}=0. Describe the affine Springer fibers 𝒳γ\mathcal{X}_{\gamma} and 𝒴γ\mathcal{Y}_{\gamma}.

Note: this is a good exercise if you have a whole day to kill.

2.8.9.

For G=Sp6G=\textup{Sp}_{6} and γ=(0tXY0)\gamma=\left(\begin{array}[]{cc}0&tX\\ Y&0\end{array}\right) as in Example 2.4.4, describe the 𝔾m\mathbb{G}_{m}-fixed points on 𝒳γ\mathcal{X}_{\gamma} and 𝒴γ\mathcal{Y}_{\gamma}.

2.8.10.

Verify the calculations in §2.6.3.

2.8.11.

Let G=SL2G=\textup{SL}_{2} and let γ=(t00t)\gamma=\left(\begin{array}[]{cc}t&0\\ 0&-t\end{array}\right). Describe the affine Springer fiber 𝒴γ\mathcal{Y}_{\gamma}. What is the action of W~=s0,s1\widetilde{W}=\langle{s_{0},s_{1}}\rangle on H2(𝒴γ)\textup{H}_{2}({\mathcal{Y}_{\gamma}}) in terms of the basis given by the irreducible components of 𝒴γ\mathcal{Y}_{\gamma}?

3. Lecture III: Orbital integrals

The significance of affine Springer fibers in representation theory is demonstrated by their close relationship with orbital integrals. Orbital integrals are certain integrals that appear in the harmonic analysis of pp-adic groups. Just as conjugacy classes of a finite group are fundamental to understanding its representations, orbital integrals are fundamental to understanding representations of pp-adic groups. In certain cases, orbital integrals can be interpreted as counting points on affine Springer fibers.

3.1. Integration on a pp-adic group

3.1.1. The setup

Let FF be a local non-archimedean field, i.e., FF is either a finite extension of p\mathbb{Q}_{p} or a finite extension of 𝔽p((t))\mathbb{F}_{p}(\!(t)\!). Then FF has a discrete valuation val:F×\textup{val}:F^{\times}\to\mathbb{Z} which we normalize to be surjective. Let 𝒪F\mathcal{O}_{F} be the valuation ring of FF and kk be the residue field. Therefore, unlike in the previous sections, kk is a finite field. We assume that char(k)\textup{char}(k) is large with respect to the groups in question.

3.1.2. Haar measure and integration

Let GG be an algebraic group over FF. The topological group G(F)G(F) is locally compact and totally disconnected. It has a right invariant Haar measure μG\mu_{G} which is unique up to a scalar. For a measurable subset SG(F)S\subset G(F), we denote its volume under μG\mu_{G} by vol(S,μG)\textup{vol}(S,\mu_{G}). Fixing a compact open subgroup K0G(F)K_{0}\subset G(F), we may normalize the Haar measure μG\mu_{G} so that K0K_{0} has volume 11. For example, if we choose an integral model 𝒢\mathcal{G} of GG over 𝒪F\mathcal{O}_{F}, we may take K0=𝒢(𝒪F)K_{0}=\mathcal{G}(\mathcal{O}_{F}).

With the Haar measure μG\mu_{G} one can integrate smooth (i.e., locally constant) compactly supported functions on G(F)G(F) with complex values. We denote this function space by 𝒮(G(F))\mathcal{S}(G(F)) (where 𝒮\mathcal{S} stands for Schwarz). For f𝒮(G(F))f\in\mathcal{S}(G(F)), the integral

G(F)fμG\int_{G(F)}f\mu_{G}

can be calculated as follows. One can find a subgroup KK0K\subset K_{0} of finite index such that ff is right KK-invariant, i.e., f(gx)=f(g)f(gx)=f(g) for all gG(F)g\in G(F) and xKx\in K (see Exercise 3.7.1). Then the integral above becomes a weighted counting in the coset G(F)/KG(F)/K:

G(F)fμG=vol(K,μG)[g]G(F)/Kf(g)=1[K0:K][g]G(F)/Kf(g).\int_{G(F)}f\mu_{G}=\textup{vol}(K,\mu_{G})\sum_{[g]\in G(F)/K}f(g)=\frac{1}{[K_{0}:K]}\sum_{[g]\in G(F)/K}f(g).

It is easy to check that the right side above is independent of the choice of KK as long as ff is right KK-invariant and KK has finite index in K0K_{0}.

3.1.3. Variant

Let HGH\subset G be a subgroup defined over FF together with a Haar measure μH\mu_{H} on it. Consider a function f𝒮(H(F)\G(F))f\in\mathcal{S}(H(F)\backslash G(F)), i.e., ff is a left H(F)H(F)-invariant, locally constant function on G(F)G(F) whose support is compact modulo H(F)H(F), we may define the integral

(27) H(F)\G(F)fμGμH.\int_{H(F)\backslash G(F)}f\frac{\mu_{G}}{\mu_{H}}.

This integral is calculated in the following way. Again we choose a finite index subgroup KG(F)K\subset G(F) such that ff is right KK-invariant. Then the integral (27) can be written as a weighted sum over double cosets H(F)\G(F)/KH(F)\backslash G(F)/K:

H(F)\G(F)fμGμH=1[K0:K][g]H(F)\G(F)/Kf(g)vol(H(F)gKg1,μH).\int_{H(F)\backslash G(F)}f\frac{\mu_{G}}{\mu_{H}}=\frac{1}{[K_{0}:K]}\sum_{[g]\in H(F)\backslash G(F)/K}\frac{f(g)}{\textup{vol}(H(F)\cap gKg^{-1},\mu_{H})}.

3.2. Orbital integrals

3.2.1. Definition of orbital integrals

We continue with the setup of §3.1. We denote the Lie algebra of GG by 𝔤(F)\mathfrak{g}(F) to emphasize that it is a vector space over FF. Let φ𝒮(𝔤(F))\varphi\in\mathcal{S}(\mathfrak{g}(F)) and γ𝔤(F)\gamma\in\mathfrak{g}(F). Consider the map G(F)𝔤(F)G(F)\to\mathfrak{g}(F) given by gAd(g1)γg\mapsto\textup{Ad}(g^{-1})\gamma. Then the composition f:gφ(Ad(g1)γ)f:g\mapsto\varphi(\textup{Ad}(g^{-1})\gamma) is a smooth function on G(F)G(F). Then ff is locally constant, left invariant under the centralizer Gγ(F)G_{\gamma}(F) of γ\gamma in G(F)G(F), and has compact support modulo Gγ(F)G_{\gamma}(F).

Fix Haar measures μG\mu_{G} on G(F)G(F) and μGγ\mu_{G_{\gamma}} on Gγ(F)G_{\gamma}(F). The following integral is then a special case of (27) (except that we write the integration variable gGγ(F)\G(F)g\in G_{\gamma}(F)\backslash G(F) explicit below while not in (27))

Oγ(φ):=Gγ(F)\G(F)φ(Ad(g1)γ)μGμGγ.O_{\gamma}(\varphi):=\int_{G_{\gamma}(F)\backslash G(F)}\varphi(\textup{Ad}(g^{-1})\gamma)\frac{\mu_{G}}{\mu_{G_{\gamma}}}.

Such integrals are called orbital integrals on the Lie algebra 𝔤(F)\mathfrak{g}(F). We may similarly define orbital integrals on the group G(F)G(F) by replacing γ\gamma with an element in G(F)G(F) and φ\varphi with an element in 𝒮(G(F))\mathcal{S}(G(F)).

3.2.2. Specific situation

For the rest of the section we will restrict to the following situation. Let GG be a split reductive group over FF. We may fix an integral model of GG by base changing the corresponding Chevalley group scheme from \mathbb{Z} to 𝒪F\mathcal{O}_{F}. In the following we will regard GG as a reductive group scheme over 𝒪F\mathcal{O}_{F}. We normalize the Haar measure μG\mu_{G} on G(F)G(F) by requiring that K0=G(𝒪F)K_{0}=G(\mathcal{O}_{F}) have volume 11.

The Lie algebra 𝔤(F)\mathfrak{g}(F) contains a canonical lattice 𝔤(𝒪F)\mathfrak{g}(\mathcal{O}_{F}) coming from the integral model over 𝒪F\mathcal{O}_{F}. We will be most interested in the orbital integral of the characteristic function φ=𝟏𝔤(𝒪F)\varphi=\mathbf{1}_{\mathfrak{g}(\mathcal{O}_{F})} of the lattice 𝔤(𝒪F)\mathfrak{g}(\mathcal{O}_{F}).

3.2.3. The centralizer of γ\gamma

Suppose γ\gamma is regular semisimple so that its centralizer GγG_{\gamma} is a torus over FF. Let Fur=F^kk¯F^{\textup{ur}}=F\widehat{\otimes}_{k}\overline{k}, which is a complete discrete valuation field whose residue field is algebraically closed. We continue to let tt denote a uniformizer of FF, which is also a uniformizer in FurF^{\textup{ur}}. Using tt, the construction in §2.3.3 gives an embedding 𝕏(GγFFur)Gγ(Fur)\mathbb{X}_{*}(G_{\gamma}\otimes_{F}F^{\textup{ur}})\hookrightarrow G_{\gamma}(F^{\textup{ur}}) whose image we still denote by Λγ\Lambda_{\gamma}. This embedding being Gal(k¯/k)\textup{Gal}(\overline{k}/k)-equivariant, Λγ\Lambda_{\gamma} carries an action of Gal(k¯/k)\textup{Gal}(\overline{k}/k). We think of Λγ\Lambda_{\gamma} as an étale group scheme over kk, then the notation Λγ(k)\Lambda_{\gamma}(k) makes sense, and it is just the Gal(k¯/k)\textup{Gal}(\overline{k}/k)-invariants in Λγ\Lambda_{\gamma} if we regard the latter as a plain group. Then Λγ(k)Gγ(F)\Lambda_{\gamma}(k)\subset G_{\gamma}(F) is a discrete and cocompact subgroup.

3.2.4. Centralizers in GLn\textup{GL}_{n}

Let G=GLnG=\textup{GL}_{n} and let γ\gamma be a regular semisimple element in 𝔤(F)\mathfrak{g}(F) which is not necessarily diagonalizable over FF. Assume either char(F)=0\textup{char}(F)=0 or char(k)>n\textup{char}(k)>n. As in §2.3.4, the characteristic polynomial P(x)=xn+a1xn1++anF[x]P(x)=x^{n}+a_{1}x^{n-1}+\cdots+a_{n}\in F[x] of γ\gamma is separable, hence the FF-algebra F[x]/(P(x))F[x]/(P(x)) is isomorphic to a product of fields F1××FmF_{1}\times\cdots\times F_{m}. We have

Gγ(F)F1×××Fm×.G_{\gamma}(F)\cong F^{\times}_{1}\times\cdots\times F^{\times}_{m}.

In this case, the lattice Λγ(k)Gγ(F)=F1×××Fm×\Lambda_{\gamma}(k)\subset G_{\gamma}(F)=F^{\times}_{1}\times\cdots\times F^{\times}_{m} consists of elements of the form (td1,,tdm)(t^{d_{1}},\cdots,t^{d_{m}}) for (d1,,dm)m(d_{1},\cdots,d_{m})\in\mathbb{Z}^{m}. The quotient Λγ(k)\Gγ(F)\Lambda_{\gamma}(k)\backslash G_{\gamma}(F) is isomorphic to i=1mFi×/t\prod_{i=1}^{m}F^{\times}_{i}/t^{\mathbb{Z}}. Each Fi×/tF^{\times}_{i}/t^{\mathbb{Z}} fits into an exact sequence

(28) 0𝒪Fi×Fi×/t/ei00\to\mathcal{O}^{\times}_{F_{i}}\to F^{\times}_{i}/t^{\mathbb{Z}}\to\mathbb{Z}/e_{i}\mathbb{Z}\to 0

where eie_{i} is the ramification degree of the extension Fi/FF_{i}/F, therefore the quotient Λγ(k)\Gγ(F)\Lambda_{\gamma}(k)\backslash G_{\gamma}(F) is compact.

3.2.5. Orbital integrals in terms of counting

Consider the following subset of G(F)/G(𝒪F)G(F)/G(\mathcal{O}_{F})

Xγ:={[g]G(F)/G(𝒪F)|Ad(g1)γ𝔤(𝒪F)}.X_{\gamma}:=\{[g]\in G(F)/G(\mathcal{O}_{F})|\textup{Ad}(g^{-1})\gamma\in\mathfrak{g}(\mathcal{O}_{F})\}.

This is a set-theoretic version of the affine Springer fiber.

The group Gγ(F)G_{\gamma}(F) acts on XγX_{\gamma} by the rule Gγ(F)h:[g][hg]G_{\gamma}(F)\ni h:[g]\mapsto[hg]. For any free abelian group LGγ(F)L\subset G_{\gamma}(F), its action on G(F)/G(𝒪F)G(F)/G(\mathcal{O}_{F}) by left translation is free (because the stabilizers are necessarily finite), hence it acts freely on XγX_{\gamma}.

More generally, for any discrete cocompact subgroup LGγ(F)L\subset G_{\gamma}(F), the quotient groupoid L\XγL\backslash X_{\gamma} is finitary, i.e., it has finitely many isomorphism classes and the automorphism group of each object is finite. For a finitary groupoid YY, we define the cardinality of YY to be

(29) #Y:=yOb(Y)/1#Aut(y)\#Y:=\sum_{y\in\textup{Ob}(Y)/\cong}\frac{1}{\#\textup{Aut}(y)}

The next lemma follows directly from the definitions, whose proof is left to the reader as Exercise 3.7.2.

30 Lemma.

Let γ\gamma be a regular semisimple element in 𝔤(F)\mathfrak{g}(F). Let LGγ(F)L\subset G_{\gamma}(F) be any discrete cocompact subgroup. We have

Oγ(𝟏𝔤(𝒪F))=1vol(Gγ(F)/L,μGγ)#(L\Xγ)O_{\gamma}(\mathbf{1}_{\mathfrak{g}(\mathcal{O}_{F})})=\frac{1}{\textup{vol}(G_{\gamma}(F)/L,\mu_{G_{\gamma}})}\#\left(L\backslash X_{\gamma}\right)

with the cardinality on the right side interpreted as in (29).

3.2.6. The case G=GLnG=\textup{GL}_{n} and fractional ideals

We continue with the situation in §3.2.4. Under the identification of G(F)/G(𝒪F)G(F)/G(\mathcal{O}_{F}) with the set of 𝒪F\mathcal{O}_{F}-lattices in FnF^{n} (see (13)), we have

Xγ{lattices ΛFn|γΛΛ}.X_{\gamma}\cong\{\textup{lattices }\Lambda\subset F^{n}|\gamma\Lambda\subset\Lambda\}.

The bijection sends [g]Xγ[g]\in X_{\gamma} to the lattice Λ=g𝒪Fn\Lambda=g\mathcal{O}^{n}_{F}.

We give another interpretation of XγX_{\gamma}. Let P(x)=xn+a1xn1++anP(x)=x^{n}+a_{1}x^{n-1}+\cdots+a_{n} be the characteristic polynomial of γ\gamma. Let

A=𝒪F[x]/(P(x))A=\mathcal{O}_{F}[x]/(P(x))

be the commutative 𝒪F\mathcal{O}_{F}-subalgebra of 𝔤(F)\mathfrak{g}(F) generated by γ\gamma. The ring of total fractions Frac(A)\textup{Frac}(A) is a finite étale FF-algebra of degree nn, and AA is an order in it. The canonical action of AA on FnF^{n} realizes FnF^{n} as a free Frac(A)\textup{Frac}(A)-module of rank 11. Recall a fractional AA-ideal is a finitely generated AA-submodule MFrac(A)M\subset\textup{Frac}(A). If we choose an element eFne\in F^{n} as a basis for the Frac(A)\textup{Frac}(A)-module structure, we get a bijection

(31) {fractional A-ideals}Xγ\{\mbox{fractional $A$-ideals}\}\leftrightarrow X_{\gamma}

which sends MFrac(A)M\subset\textup{Frac}(A) to the 𝒪F\mathcal{O}_{F}-lattice MeFnM\cdot e\subset F^{n}.

Using the algebra AA, we have a canonical isomorphism

Gγ(F)Frac(A)×.G_{\gamma}(F)\cong\textup{Frac}(A)^{\times}.

This isomorphism intertwines the action of Gγ(F)G_{\gamma}(F) on XγX_{\gamma} by left translation and the action of Frac(A)×\textup{Frac}(A)^{\times} on the set of fractional AA-ideals by multiplication.

When AA happens to be a product of Dedekind domains (i.e., AA is the maximal order in Frac(A)\textup{Frac}(A)), all fractional AA-ideals are principal, which is the same as saying that the action of Gγ(F)Frac(A)×G_{\gamma}(F)\cong\textup{Frac}(A)^{\times} on XγX_{\gamma} is transitive. In general, principal fractional ideals form a homogeneous space Frac(A)×/A×\textup{Frac}(A)^{\times}/A^{\times} under Gγ(F)G_{\gamma}(F); the difficulty in counting XγX_{\gamma} in general is caused by the singularity of the ring AA.

We normalize the Haar measure μGγ\mu_{G_{\gamma}} on Gγ(F)iFi×G_{\gamma}(F)\cong\prod_{i}F^{\times}_{i} so that i𝒪Fi×\prod_{i}\mathcal{O}^{\times}_{F_{i}} gets volume 11. Choose a uniformizer tiFit_{i}\in F_{i}, we may form the lattice L0=t1××tmGγ(F)L_{0}=t^{\mathbb{Z}}_{1}\times\cdots\times t^{\mathbb{Z}}_{m}\subset G_{\gamma}(F). Now Gγ(F)/L0i𝒪Fi×G_{\gamma}(F)/L_{0}\cong\prod_{i}\mathcal{O}_{F_{i}}^{\times} has volume 11. Therefore Lemma 30 gives

(32) Oγ(𝟏𝔤(𝒪F))=#(L0\Xγ).O_{\gamma}(\mathbf{1}_{\mathfrak{g}(\mathcal{O}_{F})})=\#(L_{0}\backslash X_{\gamma}).

Using (31), we may interpret (32) as saying that Oγ(𝟏𝔤(𝒪F))O_{\gamma}(\mathbf{1}_{\mathfrak{g}(\mathcal{O}_{F})}) is the number of fractional AA-ideals up to multiplication by the powers of the tit_{i}’s.

3.3. Relation with affine Springer fibers

From this subsection we restrict to the case FF is a local function field, i.e., F=k((t))F=k(\!(t)\!) for a finite field k=𝔽qk=\mathbb{F}_{q}. Let γ𝔤(F)\gamma\in\mathfrak{g}(F) be regular semisimple. The definitions of the affine Grassmannian and the affine Springer fiber 𝒳γ\mathcal{X}_{\gamma} we gave in §2 make sense when the base field kk is a finite field, so we have a sub-ind-scheme 𝒳γ\mathcal{X}_{\gamma} of GrG\textup{Gr}_{G}, both defined over kk.

The following lemma is clear from the definitions.

33 Lemma.

The set of kk-rational points 𝒳γ(k)\mathcal{X}_{\gamma}(k) is the same as the set XγX_{\gamma} defined in §3.2.5, both as subsets of GrG(k)=G(F)/G(𝒪F)\textup{Gr}_{G}(k)=G(F)/G(\mathcal{O}_{F}).

3.3.1.

If we base change from kk to k¯\overline{k}, by Theorem 17 we know that Λγ\𝒳γ,k¯\Lambda_{\gamma}\backslash\mathcal{X}_{\gamma,\overline{k}} is a proper algebraic space over k¯\overline{k}. The proof there actually shows that this algebraic space is defined over kk, which we denote by Λγ\𝒳γ\Lambda_{\gamma}\backslash\mathcal{X}_{\gamma}. We emphasize here that Λγ\Lambda_{\gamma} is viewed as an étale group scheme over kk whose k¯\overline{k}-points is the plain group used to be denoted Λγ\Lambda_{\gamma}.

In view of Lemma 30 and Lemma 33, it is natural to expect that the orbital integral Oγ(𝟏𝔤(𝒪F))O_{\gamma}(\mathbf{1}_{\mathfrak{g}(\mathcal{O}_{F})}) can be expressed as the number of kk-points on the quotient Λγ\𝒳γ\Lambda_{\gamma}\backslash\mathcal{X}_{\gamma}. Such a relationship takes its cleanest form when G=GLnG=\textup{GL}_{n}.

3.3.2. The case of GLn\textup{GL}_{n}

In the situation of §3.2.6, upon choosing uniformizers tiFit_{i}\in F_{i}, we defined the lattice L0=t1××tmL_{0}=t^{\mathbb{Z}}_{1}\times\cdots\times t^{\mathbb{Z}}_{m}. Base change from kk to k¯\overline{k}, we may similarly define a lattice Λ~γ\widetilde{\Lambda}_{\gamma} as in §2.3.4, using the same choice of uniformizers tiFit_{i}\in F_{i} (note that Fi^kk¯F_{i}\widehat{\otimes}_{k}\overline{k} may split into a product of fields, but ti1t_{i}\otimes 1 will project to a uniformizer in each factor). The Gal(k¯/k)\textup{Gal}(\overline{k}/k)-action on Λ~γ\widetilde{\Lambda}_{\gamma} gives it the structure of an étale group scheme over kk, just as Λγ\Lambda_{\gamma}. We have Λ~γ(k)=L0\widetilde{\Lambda}_{\gamma}(k)=L_{0}. There is an analog of Theorem 17 if we replace Λγ\Lambda_{\gamma} with Λ~γ\widetilde{\Lambda}_{\gamma}. In particular, Λ~γ\𝒳γ\widetilde{\Lambda}_{\gamma}\backslash\mathcal{X}_{\gamma} is a proper algebraic space over kk admitting a surjective map from a projective scheme.

34 Theorem.

Let G=GLnG=\textup{GL}_{n}. Let γ\gamma be a regular semisimple element in 𝔤(F)\mathfrak{g}(F). We fix the Haar measure on Gγ(F)G_{\gamma}(F) such that its maximal compact subgroup gets volume 11. Then we have

(35) Oγ(𝟏𝔤(𝒪F))=#(Λ~γ\𝒳γ)(k)=i(1)iTr(Frobk,Hi(Λ~γ\𝒳γ,k¯,)).O_{\gamma}(\mathbf{1}_{\mathfrak{g}(\mathcal{O}_{F})})=\#(\widetilde{\Lambda}_{\gamma}\backslash\mathcal{X}_{\gamma})(k)=\sum_{i}(-1)^{i}\textup{Tr}\left(\textup{Frob}_{k},\textup{H}^{i}({\widetilde{\Lambda}_{\gamma}\backslash\mathcal{X}_{\gamma,\overline{k}},\mathbb{Q}_{\ell}})\right).

The second equality in (35) follows from the Grothendieck-Lefschetz trace formula. Comparing (35) with (32), we only need to argue that L0\Xγ=Λ~γ(k)\𝒳γ(k)L_{0}\backslash X_{\gamma}=\widetilde{\Lambda}_{\gamma}(k)\backslash\mathcal{X}_{\gamma}(k) is the same as (Λ~γ\𝒳γ)(k)(\widetilde{\Lambda}_{\gamma}\backslash\mathcal{X}_{\gamma})(k). Let us first make some general remarks about kk-points on quotient stacks.

3.3.3. kk-points of a quotient

We consider a quotient stack 𝒴=[A\X]\mathcal{Y}=[A\backslash X] where XX is a scheme over kk and AA is an algebraic group over kk acting on XX. Then, by definition, 𝒴(k)\mathcal{Y}(k) is the groupoid of pairs (S,f)(S,f) where SSpeckS\to\textup{Spec}\ k is a left AA-torsor, and f:SXf:S\to X is an AA-equivariant morphism. The isomorphism class of the AA-torsor SS is classified by the Galois cohomology H1(k,A):=H1(Gal(k¯/k),A(k¯))\textup{H}^{1}({k,A}):=\textup{H}^{1}({\textup{Gal}(\overline{k}/k),A(\overline{k})}). For each class ξH1(k,A)\xi\in\textup{H}^{1}({k,A}), let SξS_{\xi} be an AA-torsor over kk with class ξ\xi. We may define a twisted form of XX over kk by Xξ:=A\(Sξ×X)X_{\xi}:=A\backslash(S_{\xi}\times X). We also have the inner form Aξ:=AutA(Sξ)A_{\xi}:=\textup{Aut}_{A}(S_{\xi}) of AA acting on the kk-scheme XξX_{\xi}. It is easy to see that AA-equivariant morphisms f:SξXf:S_{\xi}\to X are in bijection with Xξ(k)X_{\xi}(k). Therefore we get a decomposition of groupoids

(36) 𝒴(k)=[A\X](k)ξH1(k,A)Aξ(k)\Xξ(k).\mathcal{Y}(k)=[A\backslash X](k)\cong\bigsqcup_{\xi\in\textup{H}^{1}({k,A})}A_{\xi}(k)\backslash X_{\xi}(k).

3.3.4. Proof of Theorem 34

We have reduced to showing that (Λ~γ\𝒳γ)(k)=Λ~γ(k)\𝒳γ(k)(\widetilde{\Lambda}_{\gamma}\backslash\mathcal{X}_{\gamma})(k)=\widetilde{\Lambda}_{\gamma}(k)\backslash\mathcal{X}_{\gamma}(k). By (36), it suffices to show that H1(k,Λ~γ)=0\textup{H}^{1}({k,\widetilde{\Lambda}_{\gamma}})=0. We use the notation from §3.3.2. From the definition of Λ~γ\widetilde{\Lambda}_{\gamma} we see that, as a Gal(k¯/k)\textup{Gal}(\overline{k}/k)-module, it is of the form

Λ~γi=1mIndGal(k¯/ki)Gal(k¯/k)\widetilde{\Lambda}_{\gamma}\cong\bigoplus_{i=1}^{m}\textup{Ind}^{\textup{Gal}(\overline{k}/k)}_{\textup{Gal}(\overline{k}/k_{i})}\mathbb{Z}

where kik_{i} is the residue field of FiF_{i}. Therefore

H1(k,Λ~γ)i=1mH1(ki,)=0.\textup{H}^{1}({k,\widetilde{\Lambda}_{\gamma}})\cong\bigoplus_{i=1}^{m}\textup{H}^{1}({k_{i},\mathbb{Z}})=0.

3.4. Stable orbital integrals

The setup is the same as §3.3. For general GG, the generalization of the formula (35) is not straightforward. Namely, the orbital integral Oγ(𝟏𝔤(𝒪F))O_{\gamma}(\mathbf{1}_{\mathfrak{g}(\mathcal{O}_{F})}) by itself does not have a cohomological interpretation. The problem is that we may not be able to find an analog of Λ~γ\widetilde{\Lambda}_{\gamma} with vanishing first Galois cohomology so that (Λ~γ\𝒳γ)(k)(\widetilde{\Lambda}_{\gamma}\backslash\mathcal{X}_{\gamma})(k) is simply Λ~γ(k)\𝒳γ(k)\widetilde{\Lambda}_{\gamma}(k)\backslash\mathcal{X}_{\gamma}(k). In view of formula (36), a natural fix to this problem is to consider the twisted forms of 𝒳γ\mathcal{X}_{\gamma} altogether. This suggests taking not just the orbital integral Oγ(𝟏𝔤(𝒪F))O_{\gamma}(\mathbf{1}_{\mathfrak{g}(\mathcal{O}_{F})}) but a sum of several orbital integrals Oγ(𝟏𝔤(𝒪F))O_{\gamma^{\prime}}(\mathbf{1}_{\mathfrak{g}(\mathcal{O}_{F})}).

3.4.1. Stable conjugacy

Fix a regular semisimple element γ𝔤(F)\gamma\in\mathfrak{g}(F). An element γ𝔤(F)\gamma^{\prime}\in\mathfrak{g}(F) is called stably conjugate to γ\gamma if it is in the same G(F¯)G(\overline{F})-orbit of γ\gamma. Equivalently, γ\gamma^{\prime} is stably conjugate to γ\gamma if χ(γ)=χ(γ)𝔠(F)\chi(\gamma^{\prime})=\chi(\gamma)\in\mathfrak{c}(F). For γ\gamma^{\prime} stably conjugate to γ\gamma, one can attach a Galois cohomology class inv(γ,γ)H1(F,Gγ)\textup{inv}(\gamma,\gamma^{\prime})\in\textup{H}^{1}({F,G_{\gamma}}) which becomes trivial in H1(F,G)\textup{H}^{1}({F,G}). The assignment γinv(γ,γ)\gamma^{\prime}\mapsto\textup{inv}(\gamma,\gamma^{\prime}) gives a bijection of pointed sets

(37) {γ𝔤(F) stably conjugate to γ}/G(F)ker(H1(F,Gγ)H1(F,G)).\{\gamma^{\prime}\in\mathfrak{g}(F)\mbox{ stably conjugate to }\gamma\}/G(F)\cong\ker(\textup{H}^{1}({F,G_{\gamma}})\to\textup{H}^{1}({F,G})).

3.4.2. The case of GLn\textup{GL}_{n}

When G=GLnG=\textup{GL}_{n}, we use the notation from §2.3.4. We have

H1(F,Gγ)i=1mH1(Fi,𝔾m)=1.\textup{H}^{1}({F,G_{\gamma}})\cong\prod_{i=1}^{m}\textup{H}^{1}({F_{i},\mathbb{G}_{m}})=\langle{1}\rangle.

Therefore, by (37), all elements stably conjugate to γ\gamma are in fact G(F)G(F)-conjugate to γ\gamma. Of course this statement can also be proved directly using companion matrices.

3.4.3. The case of SL2\textup{SL}_{2}

We consider the case G=SL2G=\textup{SL}_{2}. Let ak×(k×)2a\in k^{\times}-(k^{\times})^{2}. Let

γ=(0att0),γ=(0at210)\gamma=\left(\begin{array}[]{cc}0&at\\ t&0\end{array}\right),\quad\gamma^{\prime}=\left(\begin{array}[]{cc}0&at^{2}\\ 1&0\end{array}\right)

be two regular semisimple elements in 𝔰𝔩2(F)\mathfrak{sl}_{2}(F). Since they have the same determinant at2-at^{2}, they are stably conjugate to each other. However, they are not conjugate to each other under SL2(F)\textup{SL}_{2}(F). One can show that the stable conjugacy class of γ\gamma consists of exactly two SL2(F)\textup{SL}_{2}(F)-orbits represented by γ\gamma and γ\gamma^{\prime}, see Exercise 3.7.3.

3.4.4. Definition of stable orbital integrals

Let φ𝒮(𝔤(F))\varphi\in\mathcal{S}(\mathfrak{g}(F)). We define the stable orbital integral of φ\varphi with respect to γ\gamma to be

(38) SOγ(φ)=γOγ(φ)SO_{\gamma}(\varphi)=\sum_{\gamma^{\prime}}O_{\gamma^{\prime}}(\varphi)

where γ\gamma^{\prime} runs over G(F)G(F)-orbits of elements γ𝔤(F)\gamma^{\prime}\in\mathfrak{g}(F) that are stably conjugate to γ\gamma. For γ\gamma^{\prime} stably conjugate to γ\gamma, we have a canonical isomorphism GγGγG_{\gamma}\cong G_{\gamma^{\prime}} as FF-groups. Therefore, once we fix a Haar measure on Gγ(F)G_{\gamma}(F), we get a canonical Haar measure on the other centralizers Gγ(F)G_{\gamma^{\prime}}(F). It is with this choice that we define Oγ(φ)O_{\gamma^{\prime}}(\varphi) in (38).

3.4.5. Stable part of the cohomology

The quotient group scheme LGγ/ΛγLG_{\gamma}/\Lambda_{\gamma} acts on Λγ\𝒳γ\Lambda_{\gamma}\backslash\mathcal{X}_{\gamma}. The component group π0(LGγ/Λγ)\pi_{0}(LG_{\gamma}/\Lambda_{\gamma}) is an étale group scheme over kk whose k¯\overline{k}-points acts on H(Lγ\𝒳γ,k¯,)\textup{H}^{*}({L_{\gamma}\backslash\mathcal{X}_{\gamma,\overline{k}},\mathbb{Q}_{\ell}}). We define the stable part of this cohomology group to be the invariants under this action

H(Λγ\𝒳γ,k¯)st=H(Λγ\𝒳γ,k¯)π0(LGγ/Λγ)(k¯).\textup{H}^{*}({\Lambda_{\gamma}\backslash\mathcal{X}_{\gamma,\overline{k}}})_{\textup{st}}=\textup{H}^{*}({\Lambda_{\gamma}\backslash\mathcal{X}_{\gamma,\overline{k}}})^{\pi_{0}(LG_{\gamma}/\Lambda_{\gamma})(\overline{k})}.

It turns out that if we replace Λγ\Lambda_{\gamma} with any other Gal(k¯/k)\textup{Gal}(\overline{k}/k)-stable free abelian subgroup ΛGγ(Fur)\Lambda\subset G_{\gamma}(F^{\textup{ur}}) commensurable with Λγ\Lambda_{\gamma}, the similarly defined stable part cohomology is canonically isomorphic to the above one.

39 Theorem (Special case of Goresky-Kottwitz-MacPherson [GKM, Th 15.8] and Ngô [NgoFL, Cor 8.2.10]).

We have

(40) SOγ(𝟏𝔤(𝒪F))=1vol(Kγ,μGγ)i(1)iTr(Frobk,Hi(Λγ\𝒳γ,k¯,)st).SO_{\gamma}(\mathbf{1}_{\mathfrak{g}(\mathcal{O}_{F})})=\frac{1}{\textup{vol}(K_{\gamma},\mu_{G_{\gamma}})}\sum_{i}(-1)^{i}\textup{Tr}\left(\textup{Frob}_{k},\textup{H}^{i}({\Lambda_{\gamma}\backslash\mathcal{X}_{\gamma,\overline{k}},\mathbb{Q}_{\ell}})_{\textup{st}}\right).

Here KγGγ(F)K_{\gamma}\subset G_{\gamma}(F) is the parahoric subgroup of the torus GγG_{\gamma}.

This cohomological interpretation of the stable orbital integral is the starting point of the proof of the Fundamental Lemma (see [GKM] and [NgoFL]).

3.4.6.

Let us briefly comment on the definition of the parahoric subgroup KγK_{\gamma}. Let K~γGγ(F)\widetilde{K}_{\gamma}\subset G_{\gamma}(F) be the maximal compact subgroup. Then there is a canonical smooth group scheme 𝒢γ\mathcal{G}_{\gamma} over 𝒪F\mathcal{O}_{F} whose FF-fiber is GγG_{\gamma} and whose 𝒪F\mathcal{O}_{F} points is K~γ\widetilde{K}_{\gamma}. This group scheme 𝒢γ\mathcal{G}_{\gamma} is the finite type Néron model for the torus GγG_{\gamma}. Now let 𝒢γ𝒢γ\mathcal{G}^{\circ}_{\gamma}\subset\mathcal{G}_{\gamma} be the open subgroup scheme obtained by removing the non-neutral component of the special fiber of 𝒢γ\mathcal{G}_{\gamma}. Then, by definition, Kγ=𝒢γ(𝒪F)K_{\gamma}=\mathcal{G}^{\circ}_{\gamma}(\mathcal{O}_{F}). The positive loop group L+𝒢γL^{+}\mathcal{G}^{\circ}_{\gamma} is, up to nilpotents, the neutral component of LGγLG_{\gamma}.

3.4.7. Sketch of proof of Theorem 39

We sketch an argument which is closer to that of [GKM] than that of [NgoFL]. One can show that there exists an étale kk-subgroup Λ~LGγ\widetilde{\Lambda}\subset LG_{\gamma} commensurable with Λγ\Lambda_{\gamma}, which maps onto the étale group scheme π0(LGγ)\pi_{0}(LG_{\gamma}) over kk (and the kernel is necessarily finite). The group Λ~\widetilde{\Lambda} is an analog of Λ~γ\widetilde{\Lambda}_{\gamma} defined in §3.3.2. We form the quotient stack [Λ~\𝒳γ][\widetilde{\Lambda}\backslash\mathcal{X}_{\gamma}]. Then we have an isomorphism

(41) H(Λγ\𝒳γ,k¯,)stH([Λ~\𝒳γ,k¯],).\textup{H}^{*}({\Lambda_{\gamma}\backslash\mathcal{X}_{\gamma,\overline{k}},\mathbb{Q}_{\ell}})_{\textup{st}}\cong\textup{H}^{*}({[\widetilde{\Lambda}\backslash\mathcal{X}_{\gamma,\overline{k}}],\mathbb{Q}_{\ell}}).

In fact, by the discussion in the end of §3.4.5, the left side above does not change if we replace Λγ\Lambda_{\gamma} by a commensurable lattice, so by shrinking Λγ\Lambda_{\gamma} we may assume ΛγΛ~\Lambda_{\gamma}\subset\widetilde{\Lambda}. On the other hand, the right side above can be computed by the Leray spectral sequence associated with the map Λγ\𝒳γ,k¯[Λ~\𝒳γ,k¯]\Lambda_{\gamma}\backslash\mathcal{X}_{\gamma,\overline{k}}\to[\widetilde{\Lambda}\backslash\mathcal{X}_{\gamma,\overline{k}}] which is a torsor under the finite discrete group (Λ~/Λγ)(k¯)(\widetilde{\Lambda}/\Lambda_{\gamma})(\overline{k}), therefore H([Λ~\𝒳γ,k¯],)=H(Λγ\𝒳γ,k¯,)(Λ~/Λγ)(k¯)\textup{H}^{*}({[\widetilde{\Lambda}\backslash\mathcal{X}_{\gamma,\overline{k}}],\mathbb{Q}_{\ell}})=\textup{H}^{*}({\Lambda_{\gamma}\backslash\mathcal{X}_{\gamma,\overline{k}},\mathbb{Q}_{\ell}})^{(\widetilde{\Lambda}/\Lambda_{\gamma})(\overline{k})}. Since (Λ~/Λγ)(k¯)(\widetilde{\Lambda}/\Lambda_{\gamma})(\overline{k}) surjects onto π0(LGγ/Λγ)(k¯)\pi_{0}(LG_{\gamma}/\Lambda_{\gamma})(\overline{k}) by the choice of Λ~\widetilde{\Lambda}, we see that

H(Λγ\𝒳γ,k¯,)π0(LGγ/Λγ)(k¯)=H(Λγ\𝒳γ,k¯,)(Λ~/Λγ)(k¯),\textup{H}^{*}({\Lambda_{\gamma}\backslash\mathcal{X}_{\gamma,\overline{k}},\mathbb{Q}_{\ell}})^{\pi_{0}(LG_{\gamma}/\Lambda_{\gamma})(\overline{k})}=\textup{H}^{*}({\Lambda_{\gamma}\backslash\mathcal{X}_{\gamma,\overline{k}},\mathbb{Q}_{\ell}})^{(\widetilde{\Lambda}/\Lambda_{\gamma})(\overline{k})},

from which (41) follows.

By the discussion in §3.3.3 and (36), we have

(42) [Λ~\𝒳γ,k¯](k)ξH1(k,Λ~)Λ~(k)\𝒳γ,ξ(k).[\widetilde{\Lambda}\backslash\mathcal{X}_{\gamma,\overline{k}}](k)\cong\bigsqcup_{\xi\in\textup{H}^{1}({k,\widetilde{\Lambda}})}\widetilde{\Lambda}(k)\backslash\mathcal{X}_{\gamma,\xi}(k).

The inclusion Λ~LGγ\widetilde{\Lambda}\subset LG_{\gamma} gives

θ:H1(k,Λ~)H1(k,LGγ)H1(F,Gγ).\theta:\textup{H}^{1}({k,\widetilde{\Lambda}})\twoheadrightarrow\textup{H}^{1}({k,LG_{\gamma}})\cong\textup{H}^{1}({F,G_{\gamma}}).

The first surjection follows from Lang’s theorem because the quotient LGγ/Λ~LG_{\gamma}/\widetilde{\Lambda} is connected; the second follows from another theorem of Lang which says that H1(Fur,Gγ)\textup{H}^{1}({F^{\textup{ur}},G_{\gamma}}) vanishes 777See [Serre, Ch.X, §7, p.170, Application and Example (b)]. Let KK be a complete discrete valuation field with perfect residue field, and KurK^{\textup{ur}} its maximal unramified extension. Then Lang’s theorem asserts that KurK^{\textup{ur}} is a C1C_{1}-field. Therefore H1(Kur,A)=0\textup{H}^{1}({K^{\textup{ur}},A})=0 for any torus AA over KK.. For each ξH1(k,Λ~)\xi\in\textup{H}^{1}({k,\widetilde{\Lambda}}) such that 𝒳γ,ξ(k)\mathcal{X}_{\gamma,\xi}(k)\neq\varnothing, one can show that the image of θ(ξ)\theta(\xi) in H1(F,G)\textup{H}^{1}({F,G}) is trivial. Therefore, by (37), to each ξH1(k,Λ~)\xi\in\textup{H}^{1}({k,\widetilde{\Lambda}}) we can attach an element γ\gamma^{\prime} stably conjugate to γ\gamma, unique up to G(F)G(F)-conjugacy, such that inv(γ,γ)=θ(ξ)\textup{inv}(\gamma,\gamma^{\prime})=\theta(\xi). One can show that 𝒳γ,ξ(k)\mathcal{X}_{\gamma,\xi}(k) is in bijection with the set XγX_{\gamma^{\prime}}. Therefore, (42) implies

(43) #[Λ~\𝒳γ](k)=#ker(θ)γ#(Λ~(k)\Xγ)\#[\widetilde{\Lambda}\backslash\mathcal{X}_{\gamma}](k)=\#\ker(\theta)\sum_{\gamma^{\prime}}\#(\widetilde{\Lambda}(k)\backslash X_{\gamma^{\prime}})

where the sum is over the G(F)G(F)-orbits of those γ\gamma^{\prime} stably conjugate to γ\gamma. Applying Lemma 30 to the discrete cocompact subgroup Λ~(k)\widetilde{\Lambda}(k), we have

Oγ(𝟏𝔤(𝒪F))=1vol(Gγ(F)/Λ~(k),μGγ)#(Λ~(k)\Xγ).O_{\gamma^{\prime}}(\mathbf{1}_{\mathfrak{g}(\mathcal{O}_{F})})=\frac{1}{\textup{vol}(G_{\gamma}(F)/\widetilde{\Lambda}(k),\mu_{G_{\gamma}})}\#(\widetilde{\Lambda}(k)\backslash X_{\gamma^{\prime}}).

Plugging this into the right side of (43), we get

(44) SOγ(𝟏𝔤(𝒪F))=1#ker(θ)vol(Gγ(F)/Λ~(k),μGγ)#[Λ~\𝒳γ](k).\textup{SO}_{\gamma}(\mathbf{1}_{\mathfrak{g}(\mathcal{O}_{F})})=\frac{1}{\#\ker(\theta)\textup{vol}(G_{\gamma}(F)/\widetilde{\Lambda}(k),\mu_{G_{\gamma}})}\#[\widetilde{\Lambda}\backslash\mathcal{X}_{\gamma}](k).

By the Grothendieck-Lefschetz trace formula, #[Λ~\𝒳γ,k¯](k)\#[\widetilde{\Lambda}\backslash\mathcal{X}_{\gamma,\overline{k}}](k) is equal to the alternating Frobenius trace on H([Λ~\𝒳γ,k¯],)\textup{H}^{*}({[\widetilde{\Lambda}\backslash\mathcal{X}_{\gamma,\overline{k}}],\mathbb{Q}_{\ell}}), which, by (41), can be identified with H(Λγ\𝒳γ,k¯,)st\textup{H}^{*}({\Lambda_{\gamma}\backslash\mathcal{X}_{\gamma,\overline{k}},\mathbb{Q}_{\ell}})_{\textup{st}}. Therefore the theorem follows from the identity (44) together with the volume identity

(45) #ker(θ)vol(Gγ(F)/Λ~(k),μGγ)=vol(Kγ,μGγ).\#\ker(\theta)\textup{vol}(G_{\gamma}(F)/\widetilde{\Lambda}(k),\mu_{G_{\gamma}})=\textup{vol}(K_{\gamma},\mu_{G_{\gamma}}).

To show this, let C=L+𝒢γΛ~C=L^{+}\mathcal{G}^{\circ}_{\gamma}\cap\widetilde{\Lambda} (where 𝒢γ\mathcal{G}^{\circ}_{\gamma} is the connected Néron model of GγG_{\gamma} whose 𝒪F\mathcal{O}_{F} points is KγK_{\gamma}, see §3.4.6). This is a finite étale group over kk. We have a short exact sequence of group ind-schemes over kk

1CL+𝒢γ×Λ~(LGγ)red11\to C\to L^{+}\mathcal{G}^{\circ}_{\gamma}\times\widetilde{\Lambda}\to(LG_{\gamma})^{\textup{red}}\to 1

The associated six term exact sequence for Gal(k¯/k)\textup{Gal}(\overline{k}/k)-cohomology gives

1C(k)Kγ×Λ~(k)Gγ(F)H1(k,C)ker(θ)11\to C(k)\to K_{\gamma}\times\widetilde{\Lambda}(k)\to G_{\gamma}(F)\to\textup{H}^{1}({k,C})\to\ker(\theta)\to 1

from which we get (45), using that #C(k)=#H1(k,C)\#C(k)=\#\textup{H}^{1}({k,C}). ∎

3.5. Examples in SL2\textup{SL}_{2}

By Theorem 39, in order to calculate orbital integrals, we need to know not just the geometry of the affine Springer fiber 𝒳γ\mathcal{X}_{\gamma}, but also the action of Frobenius on its cohomology. Having already seen many examples of affine Springer fibers over an algebraically closed field in §2.4, our emphasis here will be on the Frobenius action.

In this subsection we let G=SL2G=\textup{SL}_{2} and assume char(k)>2\textup{char}(k)>2. We will compute several orbital integrals in this case and verify Theorem 39 in these cases by explicit calculations.

3.5.1. Unramified case: γ\gamma

Let γ=(0att0)𝔰𝔩2(F)\gamma=\left(\begin{array}[]{cc}0&at\\ t&0\end{array}\right)\in\mathfrak{sl}_{2}(F) be a regular semisimple element with ak×(k×)2a\in k^{\times}-(k^{\times})^{2}.

Let E𝔤𝔩2(F)E\subset\mathfrak{gl}_{2}(F) be the centralizer of γ\gamma in 𝔤𝔩2(F)\mathfrak{gl}_{2}(F). Then E={u+vγ|u,vF}E=\{u+v\gamma|u,v\in F\} is an unramified quadratic extension of FF obtained by adjoining a\sqrt{a}. Therefore we have E=kE((t))E=k_{E}(\!(t)\!), with kE=k(a)k_{E}=k(\sqrt{a}). We have Gγ(F)=(E×)Nm=1=ker(Nm:E×F×)G_{\gamma}(F)=(E^{\times})^{\textup{Nm}=1}=\ker(\textup{Nm}:E^{\times}\to F^{\times}), which is compact. We fix a Haar measure on Gγ(F)G_{\gamma}(F) with total volume 1.

Let 𝒳γ\mathcal{X}_{\gamma} be the affine Springer fiber of γ\gamma, which is a scheme over kk. Lemma 30 implies that

Oγ(𝟏𝔤(𝒪F))=#Xγ=#𝒳γ(k).O_{\gamma}(\mathbf{1}_{\mathfrak{g}(\mathcal{O}_{F})})=\#X_{\gamma}=\#\mathcal{X}_{\gamma}(k).

In §2.2.4 we have shown that 𝒳γ,k¯\mathcal{X}_{\gamma,\overline{k}} is an infinite union of rational curves Cn1C_{n}\cong\mathbb{P}^{1} indexed by the integers nn\in\mathbb{Z}. Since γ\gamma is diagonalizable over EE, each component CnC_{n} is in fact defined over kE=k(a)k_{E}=k(\sqrt{a}). The lattice ΛγGγ(Fur)\Lambda_{\gamma}\subset G_{\gamma}(F^{\textup{ur}}) is contained in Gγ(E)E×G_{\gamma}(E)\cong E^{\times}, and is generated by the uniformizer tEt\in E. We label the components CnC_{n} so that tΛγt\in\Lambda_{\gamma} sends CnC_{n} to Cn+1C_{n+1}. Let xn+1/2:=CnCn+1x_{n+1/2}:=C_{n}\cap C_{n+1}, which is a kEk_{E}-point of 𝒳γ\mathcal{X}_{\gamma}.

The action of the nontrivial involution σGal(kE/k)\sigma\in\textup{Gal}(k_{E}/k) on Gγ(E)G_{\gamma}(E) is by inversion, hence it also acts on Λγ\Lambda_{\gamma} by inversion. The standard lattice 𝒪F2\mathcal{O}^{2}_{F} lies in both C0C_{0} and C1C_{1}, hence it is the point x1/2x_{1/2}. Therefore the point x1/2x_{1/2} is fixed by σ\sigma since it is defined over kk. Since the action of σ\sigma on 𝒳γ,kE\mathcal{X}_{\gamma,k_{E}} is compatible with its action on Λγ\Lambda_{\gamma} (by inversion), the only possibility is that

σ(Cn)=C1n,σ(xn+1/2)=xn+1/2,n.\sigma(C_{n})=C_{1-n},\quad\sigma(x_{n+1/2})=x_{-n+1/2},\quad\forall n\in\mathbb{Z}.

The action of σ\sigma can be represented by the picture

x3/2\textstyle{x_{-3/2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}σ\scriptstyle{\sigma}C1\scriptstyle{C_{-1}}C0\scriptstyle{C_{0}}C1\scriptstyle{C_{1}}C2\scriptstyle{C_{2}}x1/2\textstyle{x_{1/2}}x5/2\textstyle{x_{5/2}}x1/2\textstyle{x_{-1/2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}σ\scriptstyle{\sigma}x3/2\textstyle{x_{3/2}}

Here each double line represents a 1\mathbb{P}^{1}. From this we see that Xγ=𝒳γ(k)X_{\gamma}=\mathcal{X}_{\gamma}(k) consists of only one point x1/2x_{1/2}, namely the standard lattice 𝒪F2\mathcal{O}^{2}_{F}. This implies that

(46) Oγ(𝟏𝔤(𝒪F))=1.O_{\gamma}(\mathbf{1}_{\mathfrak{g}(\mathcal{O}_{F})})=1.

3.5.2. Unramified case: γ\gamma^{\prime}

Now consider the element γ=(0at210)𝔰𝔩2(F)\gamma^{\prime}=\left(\begin{array}[]{cc}0&at^{2}\\ 1&0\end{array}\right)\in\mathfrak{sl}_{2}(F). In Exercise 3.7.3 we see that γ\gamma^{\prime} is stably conjugate to γ\gamma but not conjugate to γ\gamma under SL2(F)\textup{SL}_{2}(F). However, γ\gamma^{\prime} is conjugate to γ\gamma under SL2(E)\textup{SL}_{2}(E). Therefore, the affine Springer fiber 𝒳γ\mathcal{X}_{\gamma^{\prime}} still looks the same as 𝒳γ\mathcal{X}_{\gamma} over kEk_{E}, but the action of σGal(kE/k)\sigma\in\textup{Gal}(k_{E}/k) is different.

Consider the component of 𝒳γ,kE\mathcal{X}_{\gamma^{\prime},k_{E}} whose kEk_{E}-points consist of γ\gamma^{\prime}-stable lattices ΛE2\Lambda\subset E^{2} such that t𝒪E𝒪EΛ𝒪Et1𝒪Et\mathcal{O}_{E}\oplus\mathcal{O}_{E}\subset\Lambda\subset\mathcal{O}_{E}\oplus t^{-1}\mathcal{O}_{E}. This component is cut out by conditions defined over kk, so it is stable under σ\sigma, and we call this component C0C^{\prime}_{0}. We label the other components of 𝒳γ,kE\mathcal{X}_{\gamma^{\prime},k_{E}} by CnC^{\prime}_{n} (nn\in\mathbb{Z}) so that the generator tΛγt\in\Lambda_{\gamma^{\prime}} sends CnC^{\prime}_{n} to Cn+1C^{\prime}_{n+1}. Let xn+1/2=CnCn+1𝒳γ(kE)x^{\prime}_{n+1/2}=C^{\prime}_{n}\cap C^{\prime}_{n+1}\in\mathcal{X}_{\gamma^{\prime}}(k_{E}). Since the action of σ\sigma on 𝒳γ,kE\mathcal{X}_{\gamma,k_{E}} is compatible with its action on Λγ\Lambda_{\gamma^{\prime}} (by inversion), the only possibility is that

σ(Cn)=Cn,σ(xn+1/2)=xn1/2,n.\sigma(C^{\prime}_{n})=C^{\prime}_{-n},\quad\sigma(x^{\prime}_{n+1/2})=x^{\prime}_{-n-1/2},\quad\forall n\in\mathbb{Z}.

The action of σ\sigma can be represented by the picture

x1/2\textstyle{x^{\prime}_{-1/2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}σ\scriptstyle{\sigma}x1/2\textstyle{x^{\prime}_{1/2}}x5/2\textstyle{x^{\prime}_{-5/2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}σ\scriptstyle{\sigma}C2\scriptstyle{C^{\prime}_{-2}}C1\scriptstyle{C^{\prime}_{-1}}C0\scriptstyle{C^{\prime}_{0}}C1\scriptstyle{C^{\prime}_{1}}C2\scriptstyle{C^{\prime}_{2}}x3/2\textstyle{x^{\prime}_{-3/2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}σ\scriptstyle{\sigma}x3/2\textstyle{x^{\prime}_{3/2}}x5/2\textstyle{x^{\prime}_{5/2}}

Therefore no point xn+1/2x^{\prime}_{n+1/2} is defined over kk. The component C0C^{\prime}_{0} is the only one that is defined over kk, and it has to be isomorphic to 1\mathbb{P}^{1} over kk because it is so over kEk_{E} (there are no nontrivial Brauer-Severi varieties over a finite field). We see that 𝒳γ(k)=C0(k)1(k)\mathcal{X}_{\gamma^{\prime}}(k)=C^{\prime}_{0}(k)\cong\mathbb{P}^{1}(k) has q+1q+1 elements. Therefore

(47) Oγ(𝟏𝔤(𝒪F))=q+1.O_{\gamma^{\prime}}(\mathbf{1}_{\mathfrak{g}(\mathcal{O}_{F})})=q+1.

Adding up (46) and (47) we get

SOγ(𝟏𝔤(𝒪F))=q+2.SO_{\gamma}(\mathbf{1}_{\mathfrak{g}(\mathcal{O}_{F})})=q+2.

3.5.3. Unramified case: cohomology

The quotient Λγ\𝒳γ\Lambda_{\gamma}\backslash\mathcal{X}_{\gamma} is a nodal rational curve obtained from 1\mathbb{P}^{1} by glueing two kk-points into a nodal point.

Now let us consider the quotient Λγ\𝒳γ\Lambda_{\gamma^{\prime}}\backslash\mathcal{X}_{\gamma^{\prime}}. Over kEk_{E} this is also a nodal rational curve consisting of a unique node yy which is image of all xjx_{j}. While yy is a kk-point of the quotient Λγ\𝒳γ\Lambda_{\gamma^{\prime}}\backslash\mathcal{X}_{\gamma^{\prime}}, none of its preimages xjx_{j} are defined over kk. Therefore, the q+1q+1 points in C0(k)C^{\prime}_{0}(k) still map injectively to the quotient, in addition to the point yy. We conclude that (Λγ\𝒳γ)(k)(\Lambda_{\gamma^{\prime}}\backslash\mathcal{X}_{\gamma^{\prime}})(k) consists of q+2q+2 points. Since LGγ/ΛγLG_{\gamma^{\prime}}/\Lambda_{\gamma^{\prime}} is connected, the stable part of the cohomology of Λγ\𝒳γ\Lambda_{\gamma^{\prime}}\backslash\mathcal{X}_{\gamma^{\prime}} is the whole H(Λγ\𝒳γ)\textup{H}^{*}({\Lambda_{\gamma^{\prime}}\backslash\mathcal{X}_{\gamma^{\prime}}}), and the alternating sum of Frobenius trace on it is the cardinality of (Λγ\𝒳γ)(k)(\Lambda_{\gamma^{\prime}}\backslash\mathcal{X}_{\gamma^{\prime}})(k). In this special case we have verified the formula (40). We remark that the action of σ\sigma on the 1-dimensional space H1(Λγ\𝒳γ,k¯,)\textup{H}^{1}({\Lambda_{\gamma^{\prime}}\backslash\mathcal{X}_{\gamma^{\prime},\overline{k}},\mathbb{Q}_{\ell}}) is by 1-1, and the Grothendieck-Lefschetz trace formula for Frobenius reads #(Λγ\𝒳γ)(k)=1+1+q\#(\Lambda_{\gamma^{\prime}}\backslash\mathcal{X}_{\gamma^{\prime}})(k)=1+1+q instead of 11+q=q1-1+q=q, the latter being the number of kk-points on a nodal rational curve obtained by identifying two kk-points on 1\mathbb{P}^{1}.

3.5.4. Ramified case: orbital integrals

Consider the elements γ=(0t2t0)\gamma=\left(\begin{array}[]{cc}0&t^{2}\\ t&0\end{array}\right) and γ=(0at2a1t0)\gamma^{\prime}=\left(\begin{array}[]{cc}0&at^{2}\\ a^{-1}t&0\end{array}\right) where ak×(k×)2a\in k^{\times}-(k^{\times})^{2}. Again γ\gamma and γ\gamma^{\prime} are stably conjugate but not conjugate in SL2(F)\textup{SL}_{2}(F).

Let EE be the centralizer of γ\gamma in 𝔤𝔩2(F)\mathfrak{gl}_{2}(F). Then EE is a ramified quadratic extension of FF, and Gγ(F)=(E×)Nm=1G_{\gamma}(F)=(E^{\times})^{\textup{Nm}=1}. Similarly, let EE^{\prime} be the centralizer of γ\gamma^{\prime} in 𝔤𝔩2(F)\mathfrak{gl}_{2}(F). Then EE^{\prime} is another ramified quadratic extension of FF, and Gγ(F)=(E×)Nm=1G_{\gamma^{\prime}}(F)=(E^{\prime\times})^{\textup{Nm}=1}. We choose Haar measures on compact groups Gγ(F)G_{\gamma}(F) and Gγ(F)G_{\gamma^{\prime}}(F) with total volume 11.

In both cases, 𝒳γ\mathcal{X}_{\gamma} and 𝒳γ\mathcal{X}_{\gamma^{\prime}} are isomorphic to 1\mathbb{P}^{1} as varieties over kk. In fact we have shown in §2.2.5 that these varieties are isomorphic to 1\mathbb{P}^{1} over k¯\overline{k}, hence they must be isomorphic to 1\mathbb{P}^{1} over kk as well. Both 𝒳γ(k)\mathcal{X}_{\gamma}(k) and 𝒳γ(k)\mathcal{X}_{\gamma^{\prime}}(k) consist of lattices t𝒪F𝒪FΛ𝒪Ft1𝒪Ft\mathcal{O}_{F}\oplus\mathcal{O}_{F}\subset\Lambda\subset\mathcal{O}_{F}\oplus t^{-1}\mathcal{O}_{F}, therefore 𝒳γ=𝒳γ\mathcal{X}_{\gamma}=\mathcal{X}_{\gamma^{\prime}} as subvarieties of GrG\textup{Gr}_{G}. By Lemma 30, we have

Oγ(𝟏𝔤(𝒪F))=Oγ(𝟏𝔤(𝒪F))=#𝒳γ(k)=q+1.O_{\gamma}(\mathbf{1}_{\mathfrak{g}(\mathcal{O}_{F})})=O_{\gamma^{\prime}}(\mathbf{1}_{\mathfrak{g}(\mathcal{O}_{F})})=\#\mathcal{X}_{\gamma}(k)=q+1.

Therefore

(48) SOγ(𝟏𝔤(𝒪F))=2(q+1).SO_{\gamma}(\mathbf{1}_{\mathfrak{g}(\mathcal{O}_{F})})=2(q+1).

3.5.5. Ramified case: cohomology

In the setup of §3.5.4, Λγ=0\Lambda_{\gamma}=0. The component group of LGγLG_{\gamma} is /2\mathbb{Z}/2\mathbb{Z}, but its action on 𝒳γ\mathcal{X}_{\gamma} is trivial. Therefore the stable part of the cohomology is the whole H(𝒳γ,)\textup{H}^{*}({\mathcal{X}_{\gamma},\mathbb{Q}_{\ell}}), on which the alternating sum of the Frobenius gives the cardinality of 𝒳γ(k)1(k)\mathcal{X}_{\gamma}(k)\cong\mathbb{P}^{1}(k). However, the parahoric subgroup of Gγ(F)=(E×)Nm=1G_{\gamma}(F)=(E^{\times})^{\textup{Nm}=1} has index 22 in it (KγK_{\gamma} consists of those e(E×)Nm=1e\in(E^{\times})^{\textup{Nm}=1} whose reduction in kk is 11). Therefore, the right side of formula (40) gets a factor 2=vol(Kγ,dγg)12=\textup{vol}(K_{\gamma},d_{\gamma}g)^{-1} in front of #𝒳γ(k)\#\mathcal{X}_{\gamma}(k). This is consistent with (48), and we have checked the formula (40) in our special case.

3.6. Remarks on the Fundamental Lemma

Let us go back to the situation in §3.5.1. What happens if we take the difference of Oγ(𝟏𝔤(𝒪F))O_{\gamma^{\prime}}(\mathbf{1}_{\mathfrak{g}(\mathcal{O}_{F})}) and Oγ(𝟏𝔤(𝒪F))O_{\gamma}(\mathbf{1}_{\mathfrak{g}(\mathcal{O}_{F})}) instead of their sum? Is there a geometric interpretation of this difference analogous to Theorem 39?

3.6.1. The κ\kappa-orbital integral

The linear combination Oγ(𝟏𝔤(𝒪F))Oγ(𝟏𝔤(𝒪F))O_{\gamma}(\mathbf{1}_{\mathfrak{g}(\mathcal{O}_{F})})-O_{\gamma^{\prime}}(\mathbf{1}_{\mathfrak{g}(\mathcal{O}_{F})}) is an example of κ\kappa-orbital integrals. More generally, let κ\kappa be a character of H1(F,Gγ)\textup{H}^{1}({F,G_{\gamma}}), then we define the κ\kappa-orbital integral of φ𝒮(𝔤(F))\varphi\in\mathcal{S}(\mathfrak{g}(F)) to be

Oγκ(φ)=γκ(inv(γ,γ))Oγ(φ)O^{\kappa}_{\gamma}(\varphi)=\sum_{\gamma^{\prime}}\kappa(\textup{inv}(\gamma,\gamma^{\prime}))O_{\gamma^{\prime}}(\varphi)

where the sum is over the G(F)G(F)-orbits in the stable conjugacy class of γ\gamma.

3.6.2. Statement of the Fundamental Lemma

The Langlands-Shelstad conjecture, also known as the Fundamental Lemma, states that the κ\kappa-orbital integral of 𝟏𝔤(𝒪F)\mathbf{1}_{\mathfrak{g}(\mathcal{O}_{F})} for γ𝔤(F)\gamma\in\mathfrak{g}(F) is equal to the stable orbital integral of an element γH𝔥(F)\gamma_{H}\in\mathfrak{h}(F) for a smaller group HH, up to a simple factor. In formula, the Fundamental Lemma is the identity

Oγκ(𝟏𝔤(𝒪F))=Δ(γ,γH)SOγH(𝟏𝔥(𝒪F)).O^{\kappa}_{\gamma}(\mathbf{1}_{\mathfrak{g}(\mathcal{O}_{F})})=\Delta(\gamma,\gamma_{H})\cdot SO_{\gamma_{H}}(\mathbf{1}_{\mathfrak{h}(\mathcal{O}_{F})}).

The smaller group HH depends on both GG and κ\kappa, and is called the endoscopic group of (G,κ)(G,\kappa). The number Δ(γ,γH)\Delta(\gamma,\gamma_{H}) is called the transfer factor, which turns out to be an integer power of qq (depending on γ\gamma and γH\gamma_{H}) if γ\gamma is chosen appropriately from its stable conjugacy class, and the measures on GγG_{\gamma} and HγHH_{\gamma_{H}} are chosen properly.

3.6.3. A simple case

In the situation of §3.5.1, take the nontrivial character κ\kappa on H1(F,Gγ)/2\textup{H}^{1}({F,G_{\gamma}})\cong\mathbb{Z}/2\mathbb{Z}, the corresponding endoscopic group HH is isomorphic to the torus GγG_{\gamma}; but in general it is not always isomorphic to a subgroup of GG. The Fundamental Lemma in this case is the identity

Oγ(𝟏𝔤(𝒪F))Oγ(𝟏𝔤(𝒪F))=q=qSOγH(𝟏𝔥(𝒪F))O_{\gamma^{\prime}}(\mathbf{1}_{\mathfrak{g}(\mathcal{O}_{F})})-O_{\gamma}(\mathbf{1}_{\mathfrak{g}(\mathcal{O}_{F})})=q=q\cdot SO_{\gamma_{H}}(\mathbf{1}_{\mathfrak{h}(\mathcal{O}_{F})})

where γH=γ\gamma_{H}=\gamma if we identify HH with GγG_{\gamma}.

On the other hand, in the ramified situation §3.5.4, the κ\kappa-orbital integral of γ\gamma for the nontrivial κ\kappa vanishes. This maybe explained without calculating the orbital integrals explicitly, for in general, κ\kappa must factor through a further quotient of H1(F,Gγ)\textup{H}^{1}({F,G_{\gamma}}) for Oγκ(𝟏𝔤(𝒪F))O^{\kappa}_{\gamma}(\mathbf{1}_{\mathfrak{g}(\mathcal{O}_{F})}) to be possibly nonzero. For the precise statement, see [NgoFL, Prop 8.2.7].

3.6.4. Comments on the proof

The Fundamental Lemma for general GG and function field FF was established by B-C. Ngô [NgoFL]. There is a generalization of Theorem 39 to κ\kappa-orbital integrals, in which we replace the stable part of the cohomology of Λγ\𝒳γ\Lambda_{\gamma}\backslash\mathcal{X}_{\gamma} by the κ\kappa-isotypic part. Using this generalization, Goresky, Kottwitz and MacPherson [GKM] reformulated the Fundamental Lemma as a relation between cohomology groups of the affine Springer fiber 𝒳γ\mathcal{X}_{\gamma} and its endoscopic cousin 𝒳γHH\mathcal{X}^{H}_{\gamma_{H}}. They were also able to prove the Fundamental Lemma in some special but highly nontrivial cases. Ngô’s proof builds on this cohomological reformulation, but also uses a new ingredient, namely Hitchin fibers, which can be viewed as a “global” analog of affine Springer fibers. This will be the topic of the next lecture.

3.7. Exercises

In these exercises, k=𝔽qk=\mathbb{F}_{q} denotes a finite field with char(k)2\textup{char}(k)\neq 2, and F=k((t))F=k(\!(t)\!).

3.7.1.

Let HH be an algebraic group over FF and let f𝒮(H(F))f\in\mathcal{S}(H(F)). Show that there exists a compact open subgroup KH(F)K\subset H(F) such that ff is both left and right invariant under KK.

3.7.2.

Prove Lemma 30.

3.7.3.

Let ak×(k×)2a\in k^{\times}-(k^{\times})^{2} and G=SL2G=\textup{SL}_{2}. Consider the matrices γ,γ\gamma,\gamma^{\prime} as in §3.4.3. Show that γ\gamma and γ\gamma^{\prime} are stably conjugate but not conjugate under SL2(F)\textup{SL}_{2}(F).

3.7.4.

Let G=GLnG=\textup{GL}_{n} over FF and let γ=diag(γ1,,γm)\gamma=\textup{diag}(\gamma_{1},\cdots,\gamma_{m}) be a block diagonal matrix in 𝔤(F)rs\mathfrak{g}(F)^{\textup{rs}}, where γi𝔤𝔩ni(F)rs\gamma_{i}\in\mathfrak{gl}_{n_{i}}(F)^{\textup{rs}}. Consider the asymptotic behavior of the orbital integral Oγ(𝟏𝔤(𝒪F))O_{\gamma}(\mathbf{1}_{\mathfrak{g}(\mathcal{O}_{F})}) as q=#kq=\#k tends to \infty. Find the smallest integer dd such that

Oγ(𝟏𝔤(𝒪F))=O(qd)O_{\gamma}(\mathbf{1}_{\mathfrak{g}(\mathcal{O}_{F})})=O(q^{d})

as qq\to\infty. Note that the O()O(\cdot) on the right side is analysts’ OO while on the left side it means orbital integral. Can you interpret dd in terms of the characteristic polynomials of the γi\gamma_{i}’s?

For an explicit estimate of Oγ(𝟏𝔤(𝒪F))O_{\gamma}(\mathbf{1}_{\mathfrak{g}(\mathcal{O}_{F})}), see [Y-Orb].

3.7.5.

Let G=SL3G=\textup{SL}_{3} and γ=(00t4100010)𝔤(F)\gamma=\left(\begin{array}[]{ccc}0&0&t^{4}\\ 1&0&0\\ 0&1&0\end{array}\right)\in\mathfrak{g}(F). Compute Oγ(𝟏𝔤(𝒪F))O_{\gamma}(\mathbf{1}_{\mathfrak{g}(\mathcal{O}_{F})}).

Hint: use the cell decomposition introduced in §2.4.2.

3.7.6.

Let G=SL2G=\textup{SL}_{2}. Let ff be the characteristic function of elements X𝔤(𝒪F)X\in\mathfrak{g}(\mathcal{O}_{F}) such that the reduction X¯\overline{X} in 𝔤(k)\mathfrak{g}(k) is regular nilpotent. Let γ𝔤(F)\gamma\in\mathfrak{g}(F) be a regular semisimple element.

  1. (1)

    Show that Oγ(f)=0O_{\gamma}(f)=0 unless det(γ)t𝒪F\det(\gamma)\in t\mathcal{O}_{F}.

  2. (2)

    When det(γ)t𝒪F\det(\gamma)\in t\mathcal{O}_{F}, show that

    Oγ(f)=Oγ(𝟏𝔤(𝒪F))Ot1γ(𝟏𝔤(𝒪F)).O_{\gamma}(f)=O_{\gamma}(\mathbf{1}_{\mathfrak{g}(\mathcal{O}_{F})})-O_{t^{-1}\gamma}(\mathbf{1}_{\mathfrak{g}(\mathcal{O}_{F})}).

3.7.7.

Let G=GL2G=\textup{GL}_{2} and γ=(0tn10)\gamma=\left(\begin{array}[]{cc}0&t^{n}\\ 1&0\end{array}\right) for n1n\geq 1 odd. Let G(F)dG(F)_{d} be the set of gG(F)g\in G(F) with valF(detg)=d\textup{val}_{F}(\det g)=d. Fix the Haar measure on G(F)G(F) such that G(𝒪F)G(\mathcal{O}_{F}) has volume 11. Show that, for any integer d0d\geq 0,

G(F)d𝟏𝔤(𝒪F)(g1γg)𝟏𝒪F2((0,1)g)𝑑g\int_{G(F)_{d}}\mathbf{1}_{\mathfrak{g}(\mathcal{O}_{F})}(g^{-1}\gamma g)\mathbf{1}_{\mathcal{O}^{2}_{F}}((0,1)g)dg

is the same as the number of closed subschemes ZZ of the plane curve y2tn=0y^{2}-t^{n}=0 satisfying: (1) the underlying topological space of ZZ is the point (y,t)=(0,0)(y,t)=(0,0); (2) dimk𝒪Z=d\dim_{k}\mathcal{O}_{Z}=d.

Note: this exercise relates orbital integrals to Hilbert schemes of points on curves. This relationship has been used in [Y-Orb] to provide an estimate for orbital integrals for GLn\textup{GL}_{n}. See also §4.2.6 and §4.3.5 for an global analog.

4. Lecture IV: Hitchin fibration

During the second half of 1980s, Hitchin introduced the famous integrable system, the moduli space of Higgs bundles, in his study of gauge theory. Around the same time, Kazhdan and Lusztig introduced affine Springer fibers as natural analogs of Springer fibers. For more than 15 years these two objects stayed unrelated until B-C.Ngô saw a connection between the two. Ngô’s fundamental insight can be summarized as saying that Hitchin fibers are global analogs of affine Springer fibers, while affine Springer fibers are local versions of Hitchin fibers. Here “global” refers to objects involving a global function field, or an algebraic curve, rather than just a local function field, or a formal disk. This observation, along with ingenious technical work, allowed Ngô to prove the Fundamental Lemma for orbital integrals conjectured by Langlands and Shelstad. We will review Hitchin’s integrable system in a slightly more general setting, and make precise its connection to affine Springer fibers.

4.1. The Hitchin moduli stack

4.1.1. The setting

We are back to the setting in §1.1. In addition, we fix an algebraic curve XX over kk (assumed algebraically closed) which is smooth, projective and connected.

4.1.2. The moduli stack of bundles

There is a moduli stack Bunn\textup{Bun}_{n} classifying vector bundles of rank nn over XX. For any kk-algebra RR, Bunn(R)\textup{Bun}_{n}(R) is the groupoid of rank nn vector bundles (locally free coherent sheaves) on XR:=X×SpeckSpecRX_{R}:=X\times_{\textup{Spec}\ k}\textup{Spec}\ R. The stack Bunn\textup{Bun}_{n} is algebraic, see [LMB, Th 4.6.2.1]. Moreover, it is smooth and locally of finite type over kk.

4.1.3. GG-torsors

Recall a (right) GG-torsor over XX is a scheme \mathcal{E} over XX with a fiberwise action of GG, such that locally for the étale topology of XX, \mathcal{E} becomes the G×XG\times X and the GG-action becomes the right translation action of GG on the first factor.

For general reductive GG, We have the moduli stack BunG\textup{Bun}_{G} of GG-torsors over XX. For a kk-algebra RR, the RR-points of BunG\textup{Bun}_{G} is the groupoid of GG-torsors over XRX_{R}. Then BunG\textup{Bun}_{G} is also a smooth algebraic stack locally of finite type over kk.

4.1.4. Associated bundles

Let (V,ρ)(V,\rho) be a kk-representation of GG. Let \mathcal{E} be a GG-torsor over XX. Then there is a vector bundle ρ()\rho(\mathcal{E}) of XX whose total space is

Tot(ρ())=G\(×V)\textup{Tot}(\rho(\mathcal{E}))=G\backslash(\mathcal{E}\times V)

where GG acts on (e,v)×V(e,v)\in\mathcal{E}\times V by g(e,v)=(eg1,ρ(g)v)g\cdot(e,v)=(eg^{-1},\rho(g)v). The vector bundle ρ()\rho(\mathcal{E}) is said to be associated to \mathcal{E} and ρ\rho.

When G=GLnG=\textup{GL}_{n}, there is an equivalence of groupoids

(49) {vector bundles 𝒱 of rank n over X}{GLn-torsors  over X}.\{\mbox{vector bundles $\mathcal{V}$ of rank $n$ over $X$}\}\cong\{\mbox{$\textup{GL}_{n}$-torsors $\mathcal{E}$ over $X$}\}.

The direction 𝒱\mathcal{V}\mapsto\mathcal{E} sends a vector bundle 𝒱\mathcal{V} to the GLn\textup{GL}_{n}-torsor of framings of 𝒱\mathcal{V}, namely take =Isom¯X(𝒪Xn,𝒱)\mathcal{E}=\underline{\textup{Isom}}_{X}(\mathcal{O}^{n}_{X},\mathcal{V}), with the natural action of GLn\textup{GL}_{n} on the trivial bundle 𝒪Xn\mathcal{O}^{n}_{X}. The other direction 𝒱\mathcal{E}\mapsto\mathcal{V} sends a GLn\textup{GL}_{n}-torsor \mathcal{E} over XX to the vector bundle St()\textup{St}(\mathcal{E}) associated to \mathcal{E} and the standard representation St of GLn\textup{GL}_{n}. The equivalence (49) gives a canonical isomorphism of stacks BunnBunGLn\textup{Bun}_{n}\cong\textup{Bun}_{\textup{GL}_{n}}.

4.1.5.

For other classical groups GG, GG-torsors have more explicit descriptions in terms of vector bundles. For example, when G=SLnG=\textup{SL}_{n}, a GG-torsor over XX amounts to the same thing as a pair (𝒱,ι)(\mathcal{V},\iota) where 𝒱\mathcal{V} is a vector bundle over XX of rank nn and ι\iota is an isomorphism of line bundles n𝒱𝒪X\wedge^{n}\mathcal{V}\cong\mathcal{O}_{X}.

When G=Sp2nG=\textup{Sp}_{2n}, the groupoid of GG-torsors \mathcal{E} on XX is equivalent to the groupoid of pairs (𝒱,ω)(\mathcal{V},\omega) where 𝒱\mathcal{V} is a vector bundle of rank 2n2n over XX and ω:2𝒱𝒪X\omega:\wedge^{2}\mathcal{V}\to\mathcal{O}_{X} is an 𝒪X\mathcal{O}_{X}-linear map of coherent sheaves that gives a symplectic form on geometric fibers. The map in one direction sends a Sp2n\textup{Sp}_{2n}-torsor \mathcal{E} to the pair (𝒱,ω)(\mathcal{V},\omega), where 𝒱=St()\mathcal{V}=\textup{St}(\mathcal{E}) is the vector bundle associated to \mathcal{E} and the standard representation St of Sp2n\textup{Sp}_{2n}, and the symplectic form ω\omega on 𝒱\mathcal{V} comes from the canonical map of Sp2n\textup{Sp}_{2n}-representations 2(St)𝟏\wedge^{2}(\textup{St})\to\mathbf{1} (where 𝟏\mathbf{1} is the trivial representation).

4.1.6. Higgs bundles

Fix a line bundle \mathcal{L} over XX. An \mathcal{L}-twisted Higgs bundle of rank nn over XX is a pair (𝒱,φ)(\mathcal{V},\varphi) where 𝒱\mathcal{V} is a vector bundle over XX of rank nn, and φ:𝒱𝒱\varphi:\mathcal{V}\to\mathcal{V}\otimes\mathcal{L} is an 𝒪X\mathcal{O}_{X}-linear map. The endomorphism φ\varphi is called a Higgs field on 𝒱\mathcal{V}.

There is a moduli stack n,\mathcal{M}_{n,\mathcal{L}} classifying \mathcal{L}-twisted Higgs bundles of rank nn over XX. The morphism n,Bunn\mathcal{M}_{n,\mathcal{L}}\to\textup{Bun}_{n} forgetting the Higgs field is representable. Therefore n,\mathcal{M}_{n,\mathcal{L}} is also an algebraic stack over kk.

4.1.7. GG-Higgs bundles

An \mathcal{L}-twisted GG-Higgs bundle is a pair (,φ)(\mathcal{E},\varphi) where \mathcal{E} is a GG-torsor over XX and φ\varphi is a global section of the vector bundle Ad()\textup{Ad}(\mathcal{E})\otimes\mathcal{L} over XX. Here, Ad()\textup{Ad}(\mathcal{E}) is the vector bundle associated to \mathcal{E} and the adjoint representation (𝔤,Ad)(\mathfrak{g},\textup{Ad}) of GG, in the sense of §4.1.4. We call φ\varphi an \mathcal{L}-twisted Higgs field on \mathcal{E}.

When G=GLnG=\textup{GL}_{n}, the notion of \mathcal{L}-twisted GG-Higgs bundle is equivalent to that of an \mathcal{L}-twisted Higgs bundle of rank nn. In fact, to each \mathcal{L}-twisted GG-Higgs bundle (,φ)(\mathcal{E},\varphi), we get a Higgs bundle (𝒱=St(),ϕ)(\mathcal{V}=\textup{St}(\mathcal{E}),\phi), where ϕ:𝒱𝒱\phi:\mathcal{V}\to\mathcal{V}\otimes\mathcal{L} viewed as a global section of End¯(𝒱)\underline{\textup{End}}(\mathcal{V})\otimes\mathcal{L} corresponds to φ\varphi under the canonical isomorphism End¯(𝒱)Ad()\underline{\textup{End}}(\mathcal{V})\cong\textup{Ad}(\mathcal{E}).

We also have the moduli stack G,\mathcal{M}_{G,\mathcal{L}} of \mathcal{L}-twisted Higgs GG-torsors over XX. The RR-points of G,\mathcal{M}_{G,\mathcal{L}} is the groupoid of R\mathcal{L}_{R}-twisted GG-Higgs bundles on XRX_{R}, where R\mathcal{L}_{R} denotes the pullback of \mathcal{L} to XRX_{R}. The forgetful morphism G,BunG\mathcal{M}_{G,\mathcal{L}}\to\textup{Bun}_{G} is representable, hence G,\mathcal{M}_{G,\mathcal{L}} is an algebraic stack over kk. When G=GLnG=\textup{GL}_{n}, we have a canonical isomorphism of stacks n,GLn,\mathcal{M}_{n,\mathcal{L}}\cong\mathcal{M}_{\textup{GL}_{n},\mathcal{L}}.

4.1.8. Examples

When G=SLnG=\textup{SL}_{n}, an \mathcal{L}-twisted GG-Higgs bundle over XX amounts to the same thing as a triple (𝒱,ι,φ)(\mathcal{V},\iota,\varphi) where 𝒱\mathcal{V} is a vector bundle over XX of rank nn, ι:n𝒱𝒪X\iota:\wedge^{n}\mathcal{V}\stackrel{{\scriptstyle\sim}}{{\to}}\mathcal{O}_{X} and φ:𝒱𝒱\varphi:\mathcal{V}\to\mathcal{V}\otimes\mathcal{L} satisfies Tr(φ)=0\textup{Tr}(\varphi)=0.

When G=Sp2nG=\textup{Sp}_{2n}, an \mathcal{L}-twisted GG-Higgs bundle over XX amounts to the same thing as a triple (𝒱,ω,φ)(\mathcal{V},\omega,\varphi) where 𝒱\mathcal{V} is a vector bundle over XX of rank 2n2n, ω:2𝒱𝒪X\omega:\wedge^{2}\mathcal{V}\to\mathcal{O}_{X} is nondegenerate fiberwise, and φ:𝒱𝒱\varphi:\mathcal{V}\to\mathcal{V}\otimes\mathcal{L} such that for all local sections uu and vv of 𝒱\mathcal{V}, ω(φ(u),v)+ω(u,φ(v))=0\omega(\varphi(u),v)+\omega(u,\varphi(v))=0 as a local section of \mathcal{L}.

4.1.9. Hitchin moduli stack as a cotangent bundle

In Hitchin’s original paper [Hi], he considered the case where =ωX\mathcal{L}=\omega_{X} is the sheaf of 11-forms on XX. This case is particularly important because G,ωX\mathcal{M}_{G,\omega_{X}} is closely related to the cotangent bundle of BunG\textup{Bun}_{G}. For a point BunG(R)\mathcal{E}\in\textup{Bun}_{G}(R) which is a GG-torsor over XRX_{R}, the cotangent complex of BunG\textup{Bun}_{G} at \mathcal{E} is given by the derived global sections of the complex Ad()ωX\textup{Ad}(\mathcal{E})^{\vee}\otimes\omega_{X} over XRX_{R}. Using a Killing form on 𝔤\mathfrak{g} we may identify Ad()\textup{Ad}(\mathcal{E})^{\vee} with Ad()\textup{Ad}(\mathcal{E}), therefore the cotangent complex of BunG\textup{Bun}_{G} at \mathcal{E} is 𝐑Γ(XR,Ad()ωX)\mathbf{R}\Gamma(X_{R},\textup{Ad}(\mathcal{E})\otimes\omega_{X}), which lives in degrees 0 and 11. In particular, when BunG(k)\mathcal{E}\in\textup{Bun}_{G}(k) has finite automorphism group (e.g., \mathcal{E} is stable), the Zariski cotangent space at \mathcal{E} is H0(X,Ad()ωX)\textup{H}^{0}({X,\textup{Ad}(\mathcal{E})\otimes\omega_{X}}), i.e., a cotangent vector of BunG\textup{Bun}_{G} at \mathcal{E} is the same thing as a ωX\omega_{X}-twisted Higgs field on \mathcal{E}. Therefore, TBunGT^{*}\textup{Bun}_{G} (properly defined) and G,ωX\mathcal{M}_{G,\omega_{X}} share an open substack TBunGsT^{*}\textup{Bun}^{s}_{G}, where BunGs\textup{Bun}^{s}_{G} is the open substack of stable GG-bundles.

4.2. Hitchin fibration

4.2.1. Hitchin fibration for GLn\textup{GL}_{n}

For an \mathcal{L}-twisted Higgs bundle (𝒱,φ)(\mathcal{V},\varphi) on XX, locally on XX we may view φ\varphi as a matrix with entries which are local sections of \mathcal{L}, and we may take the characteristic polynomial of this matrix. The coefficients of this polynomial are well-defined global sections of i\mathcal{L}^{\otimes i}, 1in1\leq i\leq n. More intrinsically, φ\varphi induces a map iφ:i𝒱i𝒱i\wedge^{i}\varphi:\wedge^{i}\mathcal{V}\to\wedge^{i}\mathcal{V}\otimes\mathcal{L}^{\otimes i}, and we may take

ai(φ):=Tr(iφ)H0(X,i).a_{i}(\varphi):=\textup{Tr}(\wedge^{i}\varphi)\in\textup{H}^{0}({X,\mathcal{L}^{\otimes i}}).

This way we have defined a morphism

f:n,𝒜n,:=i=1nH0(X,i)f:\mathcal{M}_{n,\mathcal{L}}\to\mathcal{A}_{n,\mathcal{L}}:=\prod_{i=1}^{n}\textup{H}^{0}({X,\mathcal{L}^{i}})

sending (𝒱,φ)(\mathcal{V},\varphi) to (a1(φ),,an(φ))(a_{1}(\varphi),\cdots,a_{n}(\varphi)). We view 𝒜n,\mathcal{A}_{n,\mathcal{L}} as an affine space over kk. The morphism ff is called the Hitchin fibration in the case G=GLnG=\textup{GL}_{n}.

4.2.2. Hitchin fibration in general

For general connected reductive GG as in §1.1, the coefficients of the characteristic polynomial in the case of GLn\textup{GL}_{n} are replaced with the fundamental GG-invariant polynomials on 𝔤\mathfrak{g}. Recall that 𝔠=𝔤G=SpecSym(𝔤)G\mathfrak{c}=\mathfrak{g}\sslash G=\textup{Spec}\ \textup{Sym}(\mathfrak{g}^{*})^{G}. Chevalley’s theorem says that Sym(𝔤)GSym(𝔱)W\textup{Sym}(\mathfrak{g}^{*})^{G}\cong\textup{Sym}(\mathfrak{t}^{*})^{W}, and the latter is a polynomial ring in rr variables. We fix homogeneous generators f1,,frf_{1},\cdots,f_{r} of Sym(𝔤)G\textup{Sym}(\mathfrak{g}^{*})^{G} as a kk-algebra, whose degrees d1drd_{1}\leq\cdots\leq d_{r} are canonically defined although fif_{i} are not canonical. When GG is almost simple, the numbers ei=di1e_{i}=d_{i}-1 are the exponents of GG. Viewing fif_{i} as a symmetric multilinear function 𝔤dik\mathfrak{g}^{\otimes d_{i}}\to k invariant under GG, for any GG-torsor \mathcal{E} over XX, fif_{i} induces a map of the associated bundles

fi:Ad()di𝒪X.f_{i}:\textup{Ad}(\mathcal{E})^{\otimes d_{i}}\to\mathcal{O}_{X}.

This further induces

fi:(Ad())didi.f^{\mathcal{L}}_{i}:(\textup{Ad}(\mathcal{E})\otimes\mathcal{L})^{\otimes d_{i}}\to\mathcal{L}^{\otimes d_{i}}.

If φ\varphi is an \mathcal{L}-twisted Higgs field on \mathcal{E}, we may evaluate fif^{\mathcal{L}}_{i} on the section φdi\varphi^{\otimes d_{i}} of (Ad())di(\textup{Ad}(\mathcal{E})\otimes\mathcal{L})^{\otimes d_{i}} to get

ai(φ):=fi(φdi)H0(X,di),i=1,2,,r.a_{i}(\varphi):=f^{\mathcal{L}}_{i}(\varphi^{\otimes d_{i}})\in\textup{H}^{0}({X,\mathcal{L}^{\otimes d_{i}}}),\quad i=1,2,\cdots,r.

The assignment (,φ)(ai(φ))1ir(\mathcal{E},\varphi)\mapsto(a_{i}(\varphi))_{1\leq i\leq r} defines the Hitchin fibration for G,\mathcal{M}_{G,\mathcal{L}}

f=fG,:G,𝒜G,:=i=1rH0(X,di).f=f_{G,\mathcal{L}}:\mathcal{M}_{G,\mathcal{L}}\to\mathcal{A}_{G,\mathcal{L}}:=\prod_{i=1}^{r}\textup{H}^{0}({X,\mathcal{L}^{\otimes d_{i}}}).

The target 𝒜G,\mathcal{A}_{G,\mathcal{L}} is again viewed as an affine space over kk, and is called the Hitchin base.

A more intrinsic way to define the Hitchin base 𝒜G,\mathcal{A}_{G,\mathcal{L}} is the following. The affine scheme 𝔠=SpecSym(𝔤)G\mathfrak{c}=\textup{Spec}\ \textup{Sym}(\mathfrak{g}^{*})^{G} is equipped with a 𝔾m\mathbb{G}_{m}-action inducing the grading on its coordinate ring. Let Tot×()X\textup{Tot}^{\times}(\mathcal{L})\to X be the complement of the zero section in the total space of \mathcal{L}. Consider the \mathcal{L}-twisted version of 𝔠\mathfrak{c} over XX:

𝔠X,:=(𝔠×Tot×())/𝔾m\mathfrak{c}_{X,\mathcal{L}}:=(\mathfrak{c}\times\textup{Tot}^{\times}(\mathcal{L}))/\mathbb{G}_{m}

where λ𝔾m\lambda\in\mathbb{G}_{m} acts by λ:(c,x~)(λc,λ1x~)\lambda:(c,\widetilde{x})\mapsto(\lambda c,\lambda^{-1}\widetilde{x}) on the two coordinates. This is an affine space bundle over XX whose fibers are isomorphic to 𝔠\mathfrak{c}. Then 𝒜G,\mathcal{A}_{G,\mathcal{L}} can be canonically identified with the moduli space of sections of the map 𝔠X,X\mathfrak{c}_{X,\mathcal{L}}\to X. In particular, every point a𝒜G,a\in\mathcal{A}_{G,\mathcal{L}} gives a map a:X[𝔠/𝔾m]a:X\to[\mathfrak{c}/\mathbb{G}_{m}].

4.2.3. The generically regular semisimple locus

Trivializing \mathcal{L} at the generic point η\eta of XX and restricting aia_{i} to η\eta, we have a polynomial Pa(y)=yna1yn1+a2yn2++(1)nanF[y]P_{a}(y)=y^{n}-a_{1}y^{n-1}+a_{2}y^{n-2}+\cdots+(-1)^{n}a_{n}\in F[y], where F=k(X)F=k(X) is the function field of XX. When Pa(y)P_{a}(y) is a separable polynomial in F[y]F[y], we call such an aa generically regular semisimple. The generic regular semisimplicity of aa is equivalent to the nonvanishing of the discriminant Δ(Pa)H0(X,n(n1))\Delta(P_{a})\in\textup{H}^{0}({X,\mathcal{L}^{n(n-1)}}) and therefore it defines an open subscheme 𝒜n,𝒜n,\mathcal{A}^{\heartsuit}_{n,\mathcal{L}}\subset\mathcal{A}_{n,\mathcal{L}}.

For general GG, viewing a𝒜G,a\in\mathcal{A}_{G,\mathcal{L}} as a map a:X[𝔠/𝔾m]a:X\to[\mathfrak{c}/\mathbb{G}_{m}] (see the end of §4.2.2), we call aa generically regular semisimple if aa sends the generic point of XX into the open substack [𝔠rs/𝔾m][\mathfrak{c}^{\textup{rs}}/\mathbb{G}_{m}]. This defines an open subscheme 𝒜G,𝒜G,\mathcal{A}^{\heartsuit}_{G,\mathcal{L}}\subset\mathcal{A}_{G,\mathcal{L}} generalizing the construction of 𝒜n,\mathcal{A}^{\heartsuit}_{n,\mathcal{L}} above.

4.2.4. Geometric properties

When deg>2g2\deg\mathcal{L}>2g-2, the stack |𝒜G,\mathcal{M}|_{\mathcal{A}^{\heartsuit}_{G,\mathcal{L}}} is smooth, see [NgoFL, Th 4.14.1]. In this situation, the morphism fG,f_{G,\mathcal{L}} is flat over 𝒜G,\mathcal{A}^{\heartsuit}_{G,\mathcal{L}}, see [NgoFL, Cor 4.16.4]. When GG is semisimple, there is a further open dense subset 𝒜G,ani𝒜G,\mathcal{A}^{\textup{ani}}_{G,\mathcal{L}}\subset\mathcal{A}^{\heartsuit}_{G,\mathcal{L}} over which \mathcal{M} is a Deligne-Mumford stack and the map fG,f_{G,\mathcal{L}} is proper, see [NgoFL, Prop 6.1.3]. Comparing to the infinite-dimensionality involved in the geometry of affine Springer fibers, the Hitchin fibration has much nicer geometric properties, and yet it is closely related to the affine Springer fibers, as we shall see in §4.4.

4.2.5. Generalization

Let HH be a reductive group over kk that fits into an exact sequence of reductive groups

1H1HA1.1\to H_{1}\to H\to A\to 1.

Let (V,ρ)(V,\rho) be a representation of HH. Then we may consider pairs (,ψ)(\mathcal{E},\psi) where \mathcal{E} is an HH-torsor over XX and ψ\psi is a section of the associated bundle ρ()\rho(\mathcal{E}). Alternatively, such a pair is the same as a morphism X[V/H]X\to[V/H]. One can prove that there is an algebraic stack H,ρ\mathcal{M}_{H,\rho} classifying such pairs. Every HH-torsor induces an AA-torsor, hence we have a morphism α:H,ρBunA\alpha:\mathcal{M}_{H,\rho}\to\textup{Bun}_{A}. Fix an AA-torsor A\mathcal{E}_{A} over XX, we denote the preimage α1(A)\alpha^{-1}(\mathcal{E}_{A}) by H,ρ,A\mathcal{M}_{H,\rho,\mathcal{E}_{A}}.

To recover the Hitchin moduli stack, we consider the case H=G×𝔾mH=G\times\mathbb{G}_{m} with H1=GH_{1}=G and A=𝔾mA=\mathbb{G}_{m}. Let V=𝔤V=\mathfrak{g} with the action ρ\rho of HH defined as follows: GG acts by the adjoint representation and 𝔾m\mathbb{G}_{m} acts by scaling on VV. An HH-torsor is a pair consisting of a GG-torsor and a line bundle \mathcal{L} on XX. Fixing the line bundle \mathcal{L} (which is equivalent to fixing an A=𝔾mA=\mathbb{G}_{m}-torsor), we get an isomorphism

H,ρ,G,.\mathcal{M}_{H,\rho,\mathcal{L}}\cong\mathcal{M}_{G,\mathcal{L}}.

For general (1H1HA1,V,ρ,A)(1\to H_{1}\to H\to A\to 1,V,\rho,\mathcal{E}_{A}) as above, we may define the analog of the Hitchin base as follows. Let 𝔠H,ρ=VH1\mathfrak{c}_{H,\rho}=V\sslash H_{1}. This is the analog of 𝔠\mathfrak{c}, and it carries an action of AA. Then we form the twisted version of 𝔠H,ρ\mathfrak{c}_{H,\rho} over XX

𝔠H,ρ,A:=A×A𝔠H,ρ\mathfrak{c}_{H,\rho,\mathcal{E}_{A}}:=\mathcal{E}_{A}\stackrel{{\scriptstyle A}}{{\times}}\mathfrak{c}_{H,\rho}

Then we define 𝒜H,ρ,A\mathcal{A}_{H,\rho,\mathcal{E}_{A}} to be the moduli space of sections to the map 𝔠H,ρ,AX\mathfrak{c}_{H,\rho,\mathcal{E}_{A}}\to X. The morphism [V/H][𝔠H,ρ/A][V/H]\to[\mathfrak{c}_{H,\rho}/A] then induces the analog of the Hitchin fibration

fH,ρ,A:H,ρ,A𝒜H,ρ,Af_{H,\rho,\mathcal{E}_{A}}:\mathcal{M}_{H,\rho,\mathcal{E}_{A}}\to\mathcal{A}_{H,\rho,\mathcal{E}_{A}}

4.2.6. Example

Consider H=GL(U)×𝔾m×𝔾mH=\textup{GL}(U)\times\mathbb{G}_{m}\times\mathbb{G}_{m}, and V=End(U)UV=\textup{End}(U)\oplus U^{*}. The action ρ(g,s1,s2)\rho(g,s_{1},s_{2}) on VV is given by (A,u)(s1gAg1,s2gu)(A,u^{*})\mapsto(s_{1}gAg^{-1},s_{2}gu^{*}). The moduli stack H,ρ\mathcal{M}_{H,\rho} then maps to Pic(X)×Pic(X)\textup{Pic}(X)\times\textup{Pic}(X) by remembering only the two 𝔾m\mathbb{G}_{m}-torsors. Fixing (1,2)Pic(X)×Pic(X)(\mathcal{L}_{1},\mathcal{L}_{2})\in\textup{Pic}(X)\times\textup{Pic}(X), its preimage in H,ρ\mathcal{M}_{H,\rho} classifies triples (𝒰,φ,β)(\mathcal{U},\varphi,\beta) where (𝒰,φ)(\mathcal{U},\varphi) is a Higgs bundle over XX of rank n=dimVn=\dim V, β\beta is an 𝒪X\mathcal{O}_{X}-linear map 𝒰2\mathcal{U}\to\mathcal{L}_{2}. The Hitchin base in this case is the same as the classical Hitchin base 𝒜n,1\mathcal{A}_{n,\mathcal{L}_{1}}. Later in §4.3.5 we will relate this moduli space to Hilbert schemes of curves.

4.3. Hitchin fibers

An important observation made by Hitchin is that the fibers of the Hitchin fibration ff can be described in “abelian” terms, namely by line bundles on certain finite coverings of XX. We elaborate on this observation for G=GLnG=\textup{GL}_{n} and G=Sp2nG=\textup{Sp}_{2n}.

4.3.1. The case of GLn\textup{GL}_{n} and the spectral curve

For a=(a1,,an)𝒜n,a=(a_{1},\cdots,a_{n})\in\mathcal{A}_{n,\mathcal{L}}, one can define a curve YaY_{a} equipped with a degree nn morphism πa:YaX\pi_{a}:Y_{a}\to X. The construction is as follows. The total space of \mathcal{L} can be written as a relative spectrum over XX

Σ:=Tot()=Spec(𝒪12)=SpecSym(𝒪1)\Sigma:=\textup{Tot}(\mathcal{L})=\textup{Spec}\ (\mathcal{O}\oplus\mathcal{L}^{\otimes-1}\oplus\mathcal{L}^{\otimes-2}\oplus\cdots)=\textup{Spec}\ \textup{Sym}(\mathcal{O}\oplus\mathcal{L}^{\otimes-1})

Let π:ΣX\pi:\Sigma\to X be the projection. Consider the map of coherent sheaves on XX

ιa:nπ𝒪Σ=𝒪12\iota_{a}:\mathcal{L}^{\otimes-n}\to\pi_{*}\mathcal{O}_{\Sigma}=\mathcal{O}\oplus\mathcal{L}^{\otimes-1}\oplus\mathcal{L}^{\otimes-2}\oplus\cdots

given in coordinates by ((1)nan,(1)n1an1,,a1,1,0,)((-1)^{n}a_{n},(-1)^{n-1}a_{n-1},\cdots,-a_{1},1,0,\cdots). By adjunction ιa\iota_{a} induces a map ιa:πn𝒪Σ\iota^{\prime}_{a}:\pi^{*}\mathcal{L}^{\otimes-n}\to\mathcal{O}_{\Sigma}, whose image we denote by a\mathcal{I}_{a}. Then a\mathcal{I}_{a} is an ideal sheaf on Σ\Sigma. Define the spectral curve YaY_{a} to be the closed subscheme of Σ\Sigma defined by a\mathcal{I}_{a}

Ya=Spec(𝒪12)/aY_{a}=\textup{Spec}\ (\mathcal{O}\oplus\mathcal{L}^{\otimes-1}\oplus\mathcal{L}^{\otimes-2}\oplus\cdots)/\mathcal{I}_{a}

If we trivialize \mathcal{L} on some open subset UXU\subset X, and view aia_{i} as functions on UU, then Ya|UY_{a}|_{U} is the subscheme of U×𝔸1U\times\mathbb{A}^{1} defined by one equation yna1yn1+a2yn2++(1)nan=0y^{n}-a_{1}y^{n-1}+a_{2}y^{n-2}+\cdots+(-1)^{n}a_{n}=0 (where yy is the coordinate on 𝔸1\mathbb{A}^{1}). The projection πa:YaX\pi_{a}:Y_{a}\to X is finite flat of degree nn. The curve YaY_{a} is called the spectral curve of aa since the fibers of πa\pi_{a} are the roots of the characteristic polynomial yna1yn1+a2yn2++(1)nany^{n}-a_{1}y^{n-1}+a_{2}y^{n-2}+\cdots+(-1)^{n}a_{n}.

When a𝒜n,a\in\mathcal{A}^{\heartsuit}_{n,\mathcal{L}}, the curve YaY_{a} is reduced and therefore smooth on a Zariski dense open subset, there is a moduli stack Pic¯(Ya)\overline{\textup{Pic}}(Y_{a}) classifying torsion-free coherent 𝒪Ya\mathcal{O}_{Y_{a}}-modules that are generically of rank 11, see [AK]. The usual Picard stack Pic(Ya)\textup{Pic}(Y_{a}) classifying line bundles on YaY_{a} is an open substack of Pic¯(Ya)\overline{\textup{Pic}}(Y_{a}), and it acts on Pic¯(Ya)\overline{\textup{Pic}}(Y_{a}) by tensoring.

50 Proposition.

Suppose a𝒜n,(k)a\in\mathcal{A}^{\heartsuit}_{n,\mathcal{L}}(k). Let a\mathcal{M}_{a} be the fiber of f:n,𝒜n,f:\mathcal{M}_{n,\mathcal{L}}\to\mathcal{A}_{n,\mathcal{L}} over aa. Then there is a canonical isomorphism of stacks

Pic¯(Ya)a.\overline{\textup{Pic}}(Y_{a})\cong\mathcal{M}_{a}.

4.3.2. Sketch of proof

We give the morphism Pic¯(Ya)a\overline{\textup{Pic}}(Y_{a})\to\mathcal{M}_{a}. For any coherent sheaf \mathcal{F} on YaY_{a}, the direct image πa,\pi_{a,*}\mathcal{F} is a coherent sheaf on XX equipped with a map φ:πa,1πa,\varphi_{\mathcal{F}}:\pi_{a,*}\mathcal{F}\otimes\mathcal{L}^{-1}\to\pi_{a,*}\mathcal{F} because \mathcal{F} is an 𝒪Ya\mathcal{O}_{Y_{a}}-module and 𝒪Ya\mathcal{O}_{Y_{a}} contains 1\mathcal{L}^{-1} as the second direct summand. When \mathcal{F} is torsion-free and generically rank 11, 𝒱:=πa,\mathcal{V}:=\pi_{a,*}\mathcal{F} is torsion-free over XX (hence a vector bundle) of rank nn, and the map φ\varphi_{\mathcal{F}} induces a Higgs field φ:𝒱𝒱\varphi:\mathcal{V}\to\mathcal{V}\otimes\mathcal{L}. The assignment (𝒱,φ)\mathcal{F}\mapsto(\mathcal{V},\varphi) defines the morphism Pic¯(Ya)a\overline{\textup{Pic}}(Y_{a})\to\mathcal{M}_{a}, which can be shown to be an isomorphism. ∎

4.3.3. The case G=Sp2nG=\textup{Sp}_{2n}

In this case a𝒜G,a\in\mathcal{A}_{G,\mathcal{L}} is a tuple (a1,,an)(a_{1},\cdots,a_{n}) with aiH0(X,2i)a_{i}\in\textup{H}^{0}({X,\mathcal{L}^{\otimes 2i}}). For a𝒜G,a\in\mathcal{A}_{G,\mathcal{L}}, we can similarly define a spectral curve YaY_{a} as the closed subscheme of the total space of \mathcal{L} cut out by the ideal locally generated by

Pa(y)=y2n+a1y2n2++an.P_{a}(y)=y^{2n}+a_{1}y^{2n-2}+\cdots+a_{n}.

Note that YaY_{a} carries an involution σ(y)=y\sigma(y)=-y under which the projection πa:YaX\pi_{a}:Y_{a}\to X is invariant. Now suppose Pa(y)P_{a}(y) is separable when restricted to the generic point of XX, so that YaY_{a} is reduced. The involution σ\sigma induces an involution σ\sigma on the compactified Picard Pic¯(Ya)\overline{\textup{Pic}}(Y_{a}). Let ()(-)^{\vee} be the relative Serre duality functor on coherent sheaves on YaY_{a}, i.e., =Hom¯𝒪X(,𝒪X)\mathcal{F}^{\vee}=\underline{\textup{Hom}}_{\mathcal{O}_{X}}(\mathcal{F},\mathcal{O}_{X}) viewed as an 𝒪Ya\mathcal{O}_{Y_{a}}-module in a natural way. Let Prym¯(Ya;σ)\overline{\textup{Prym}}(Y_{a};\sigma) be the moduli stack of pairs (,ι)(\mathcal{F},\iota) where Pic¯(Ya)\mathcal{F}\in\overline{\textup{Pic}}(Y_{a}) and ι\iota is an isomorphism ι:σ\iota:\sigma^{*}\mathcal{F}\cong\mathcal{F}^{\vee} satisfying that (σι)=ι(\sigma^{*}\iota)^{\vee}=-\iota. We called Prym¯(Ya;σ)\overline{\textup{Prym}}(Y_{a};\sigma) the compactified Prym stack of YaY_{a} with respect to the involution σ\sigma. Similar to the case of GLn\textup{GL}_{n}, we have the following description of a\mathcal{M}_{a}.

51 Proposition.

Suppose a𝒜G,a\in\mathcal{A}_{G,\mathcal{L}} such that YaY_{a} is reduced. Then there is a canonical isomorphism of stacks

Prym¯(Ya)a.\overline{\textup{Prym}}(Y_{a})\cong\mathcal{M}_{a}.

4.3.4. Sketch of proof

For a𝒜G,a\in\mathcal{A}_{G,\mathcal{L}}, points in a\mathcal{M}_{a} are triples (𝒱,ω,φ)(\mathcal{V},\omega,\varphi) where 𝒱\mathcal{V} is a vector bundle of rank 2n2n on XX, ω:2𝒱𝒪X\omega:\wedge^{2}\mathcal{V}\to\mathcal{O}_{X} is a symplectic form on 𝒱\mathcal{V} as in §4.1.5, and φ:𝒱𝒱\varphi:\mathcal{V}\to\mathcal{V}\otimes\mathcal{L} satisfies ω(φu,v)+ω(u,φv)=0\omega(\varphi u,v)+\omega(u,\varphi v)=0 for local sections u,vu,v of 𝒱\mathcal{V}, and the characteristic polynomial of φ\varphi is Pa(y)P_{a}(y). By Proposition 50, the Higgs bundle (𝒱,φ)(\mathcal{V},\varphi) gives a point Pic¯(Ya)\mathcal{F}\in\overline{\textup{Pic}}(Y_{a}). The symplectic form can be viewed as an isomorphism ȷ:(𝒱,φ)(𝒱,φ)\jmath:(\mathcal{V},\varphi)\stackrel{{\scriptstyle\sim}}{{\to}}(\mathcal{V}^{\vee},-\varphi^{\vee}) such that ȷ=ȷ\jmath^{\vee}=-\jmath. Note that the Higgs bundle (𝒱,φ)(\mathcal{V}^{\vee},-\varphi^{\vee}) corresponds to σ\sigma^{*}\mathcal{F}^{\vee} under the isomorphism in Proposition 50. The isomorphism ȷ\jmath then turns into ι:σ\iota:\sigma^{*}\mathcal{F}\stackrel{{\scriptstyle\sim}}{{\to}}\mathcal{F}^{\vee} satisfying (σι)=ι(\sigma^{*}\iota)^{\vee}=-\iota. ∎

4.3.5. Example 4.2.6 continued

Fix two line bundles 1\mathcal{L}_{1} and 2\mathcal{L}_{2} on XX, and let 1,2\mathcal{H}_{\mathcal{L}_{1},\mathcal{L}_{2}} be the fiber of the moduli stack H,ρ\mathcal{M}_{H,\rho} in §4.2.6 over (1,2)(\mathcal{L}_{1},\mathcal{L}_{2}). Then 1,2\mathcal{H}_{\mathcal{L}_{1},\mathcal{L}_{2}} classifies (𝒱,φ,β)(\mathcal{V},\varphi,\beta) where 𝒱\mathcal{V} is a vector bundle of rank nn over XX, φ:𝒰𝒰1\varphi:\mathcal{U}\to\mathcal{U}\otimes\mathcal{L}_{1} is a Higgs field and β:𝒰2\beta:\mathcal{U}\to\mathcal{L}_{2}. We have a Hitchin-type map 1,2𝒜n,1\mathcal{H}_{\mathcal{L}_{1},\mathcal{L}_{2}}\to\mathcal{A}_{n,\mathcal{L}_{1}} sending (𝒱,φ,β)(\mathcal{V},\varphi,\beta) to the (ai(φ))1in(a_{i}(\varphi))_{1\leq i\leq n}. Let a𝒜n,1a\in\mathcal{A}^{\heartsuit}_{n,\mathcal{L}_{1}} and let 1,2,a\mathcal{H}_{\mathcal{L}_{1},\mathcal{L}_{2},a} be the fiber of 1,2\mathcal{H}_{\mathcal{L}_{1},\mathcal{L}_{2}} over aa. Consider the same spectral curve YaY_{a} as in §4.3.1. Using Proposition 50 we may identify (𝒰,φ)(\mathcal{U},\varphi) with a point Pic¯(Ya)\mathcal{F}\in\overline{\textup{Pic}}(Y_{a}). The map β:πa,=𝒰2\beta:\pi_{a,*}\mathcal{F}=\mathcal{U}\to\mathcal{L}_{2} gives a map b:πa!2b:\mathcal{F}\to\pi^{!}_{a}\mathcal{L}_{2} by adjunction. Here πa!2πa2ωYa/X\pi^{!}_{a}\mathcal{L}_{2}\cong\pi^{*}_{a}\mathcal{L}_{2}\otimes\omega_{Y_{a}/X}, and the relative dualizing complex ωYa/X=ωYaπaωX1\omega_{Y_{a}/X}=\omega_{Y_{a}}\otimes\pi_{a}^{*}\omega_{X}^{-1} is a line bundle on YaY_{a} because YaY_{a} is a planar curve hence Gorenstein. Since \mathcal{F} is torsion-free and generically a line bundle, bb is an injective map of coherent sheaves. Hence the data (𝒰,φ,β)(\mathcal{U},\varphi,\beta) turns into the data of a coherent subsheaf \mathcal{F} of the line bundle πa!2\pi^{!}_{a}\mathcal{L}_{2}. Since πa!2\pi^{!}_{a}\mathcal{L}_{2} is a line bundle, the subsheaf \mathcal{F} is determined by the support of the quotient (πa!2)/(\pi^{!}_{a}\mathcal{L}_{2})/\mathcal{F}, which is a zero-dimensional subscheme of YaY_{a}. We conclude that there is a canonical isomorphism

1,2,aHilb(Ya)\mathcal{H}_{\mathcal{L}_{1},\mathcal{L}_{2},a}\cong\textup{Hilb}(Y_{a})

where Hilb(Ya)\textup{Hilb}(Y_{a}) is the disjoint union of Hilbert schemes of zero-dimensional subschemes of YaY_{a} of various lengths.

To conclude, the Hilbert scheme of points of a family of planar curves can be realized as a Hitchin-type moduli space in the framework of §4.2.5.

4.3.6. Symmetry on Hitchin fibers

In the case G=GLnG=\textup{GL}_{n} and a𝒜n,a\in\mathcal{A}^{\heartsuit}_{n,\mathcal{L}}, we have identified the Hitchin fiber a\mathcal{M}_{a} with the compactified Picard stack Pic¯(Ya)\overline{\textup{Pic}}(Y_{a}) of the spectral curve YaY_{a}. The usual Picard stack Pic(Ya)\textup{Pic}(Y_{a}) acts on Pic¯(Ya)\overline{\textup{Pic}}(Y_{a}) by tensor product. Therefore we have an action of Pic(Ya)\textup{Pic}(Y_{a}) on a\mathcal{M}_{a}. This action is simply transitive if YaY_{a} is smooth.

In the case G=Sp2nG=\textup{Sp}_{2n} and a𝒜G,a\in\mathcal{A}^{\heartsuit}_{G,\mathcal{L}}, the Prym stack Prym(Ya;σ)\textup{Prym}(Y_{a};\sigma) is defined as the moduli stack of pairs (,ι)(\mathcal{L},\iota) where Pic(Ya)\mathcal{L}\in\textup{Pic}(Y_{a}) and ι:σ1\iota:\sigma^{*}\mathcal{L}\cong\mathcal{L}^{-1} satisfying (σι)=ι(\sigma^{*}\iota)^{\vee}=\iota. Then Prym(Ya;σ)\textup{Prym}(Y_{a};\sigma) acts on aPrym¯(Ya;σ)\mathcal{M}_{a}\cong\overline{\textup{Prym}}(Y_{a};\sigma) by tensoring.

For general GG and a𝒜G,a\in\mathcal{A}^{\heartsuit}_{G,\mathcal{L}}, one can similarly define a commutative group stack 𝒫a\mathcal{P}_{a} that acts on the corresponding Hitchin fiber a\mathcal{M}_{a}. In fact, via the map a:X[𝔠/𝔾m]a:X\to[\mathfrak{c}/\mathbb{G}_{m}], the regular centralizer group scheme JJ in §2.3.6, which descends to [𝔠/𝔾m][\mathfrak{c}/\mathbb{G}_{m}], pulls back to a smooth group scheme JaJ_{a} over XX which is generically a torus. The stack 𝒫a\mathcal{P}_{a} then classifies JaJ_{a}-torsors over XX.

4.4. Relation with affine Springer fibers

In this subsection we will state a precise relationship between Hitchin fibers and affine Springer fibers, observed by Ngô.

4.4.1.

Fix a point a𝒜G,(k)a\in\mathcal{A}^{\heartsuit}_{G,\mathcal{L}}(k). Viewing aa as a map X[𝔠/𝔾m]X\to[\mathfrak{c}/\mathbb{G}_{m}], let UaXU_{a}\subset X be the preimage of [𝔠rs/𝔾m][\mathfrak{c}^{\textup{rs}}/\mathbb{G}_{m}] under aa. Since a𝒜G,a\in\mathcal{A}^{\heartsuit}_{G,\mathcal{L}}, UaU_{a} is non-empty hence the complement XUaX-U_{a} consists of finitely many kk-points of XX. For each point xXUax\in X-U_{a}, let 𝒪x\mathcal{O}_{x} be the completed local ring of XX at xx, with fraction field FxF_{x} and residue field k(x)=kk(x)=k. We fix a trivialization of \mathcal{L} near xx and we may identify ax=a|Spec𝒪xa_{x}=a|_{\textup{Spec}\ \mathcal{O}_{x}} as an element in 𝔠(𝒪x)𝔠rs(Fx)\mathfrak{c}(\mathcal{O}_{x})\cap\mathfrak{c}^{\textup{rs}}(F_{x}). Let ϵ(ax)𝔤rs(Fx)\epsilon(a_{x})\in\mathfrak{g}^{\textup{rs}}(F_{x}) be the corresponding point in the Kostant section, then we have the affine Springer fiber 𝒳ax:=𝒳ϵ(ax)\mathcal{X}_{a_{x}}:=\mathcal{X}_{\epsilon(a_{x})} in the affine Grassmannian GrG,x=LxG/Lx+G\textup{Gr}_{G,x}=L_{x}G/L^{+}_{x}G (we put xx in the subscript to emphasize that the definition of the loop groups uses the field FxF_{x}, which is isomorphic to k((t))k(\!(t)\!)). The loop group of the centralizer LGϵ(ax)LG_{\epsilon(a_{x})} acts on 𝒳ax\mathcal{X}_{a_{x}}, and the action factors through the local Picard group PaxP_{a_{x}} as in §2.3.6. On the other hand, we have the action of the global Picard stack 𝒫a\mathcal{P}_{a} on the Hitchin fiber a\mathcal{M}_{a} mentioned in §4.3.6. The product formula of Ngô roughly says that, modulo the actions of the local and global Picard stacks, a\mathcal{M}_{a} is the product of the affine Springer fibers 𝒳ax\mathcal{X}_{a_{x}} for all points xXUax\in X-U_{a}.

52 Theorem (Product Formula, Ngô [NgoHit, Th 4.6] and [NgoFL, Prop 4.15.1]).

For a𝒜(k)a\in\mathcal{A}^{\heartsuit}(k), there is a canonical morphism

𝒫a×xXUaPax(xXUa𝒳ax)a\mathcal{P}_{a}\stackrel{{\scriptstyle\prod_{x\in X-U_{a}}P_{a_{x}}}}{{\times}}\left(\prod_{x\in X-U_{a}}\mathcal{X}_{a_{x}}\right)\to\mathcal{M}_{a}

which is a homeomorphism of stacks. Here the notation P×HYP\stackrel{{\scriptstyle H}}{{\times}}Y (where HH acts on PP on the right and acts on YY on the left) means the quotient of P×YP\times Y by the action of HH given by h(p,y)=(ph1,hy)h\cdot(p,y)=(ph^{-1},hy), pP,yYp\in P,y\in Y and hHh\in H.

In the case G=GLnG=\textup{GL}_{n}, the product formula can be reinterpreted in more familiar terms using the compactified Picard stack of spectral curves, which in fact makes sense for all reduced curves. Let CC be a reduced and projective curve over kk. For each point xC(k)x\in C(k), one can define a local Picard group Px(C)P_{x}(C) whose kk points are Fx×/𝒪x×F^{\times}_{x}/\mathcal{O}^{\times}_{x}, where 𝒪x\mathcal{O}_{x} is the completed local ring of CC at xx and FxF_{x} its ring of total fractions (which is a product of fields in general). There is also a local analog P¯x(C)\overline{P}_{x}(C) of Pic¯(C)\overline{\textup{Pic}}(C) whose kk-points are the fractional 𝒪x\mathcal{O}_{x}-ideals of FxF_{x}, compare §3.2.6. Then the following variant of the product formula holds, whose proof is similar to that of Theorem 52.

53 Proposition.

Let CC be a reduced and projective curve over kk. Let Z=CCsmZ=C-C^{sm} be the singular locus of CC. Then there is a canonical morphism

Pic(C)×xZPx(C)(xZP¯x(C))Pic¯(C)\textup{Pic}(C)\stackrel{{\scriptstyle\prod_{x\in Z}P_{x}(C)}}{{\times}}\left(\prod_{x\in Z}\overline{P}_{x}(C)\right)\to\overline{\textup{Pic}}(C)

which is a homeomorphism of stacks.

The product formula provides a link between the geometry of Hitchin fibers and that of affine Springer fibers, the latter is closely related to orbital integrals as we have already seen. This link makes it possible to approach the Fundamental Lemma by studying the cohomology of Hitchin fibers. The advantage of using Hitchin fibers instead of affine Springer fibers is that the Hitchin fibration has nicer geometric properties, as we have seen in §4.2.4.

4.5. A global version of the Springer action

The product formula in Theorem 52 suggests that there should be a global analog of affine Springer theory where affine Springer fibers are replaced by Hitchin fibers. Such a theory was developed in a series of papers of the author starting with [GS].

4.5.1. Iwahori level structure

We now define a Hitchin-type analog of the affine Springer fibers 𝒴γ\mathcal{Y}_{\gamma}. We fix the curve XX and a line bundle \mathcal{L} on it as before. Let G,par\mathcal{M}^{\textup{par}}_{G,\mathcal{L}} be the moduli stack classifying (x,,φ,xB)(x,\mathcal{E},\varphi,\mathcal{E}^{B}_{x}) where xXx\in X, (,φ)(\mathcal{E},\varphi) is an \mathcal{L}-twisted Higgs GG-bundle, and xB\mathcal{E}^{B}_{x} is a reduction of the fiber x\mathcal{E}_{x} at xx to a BB-torsor xB\mathcal{E}^{B}_{x} (here BGB\subset G is a Borel subgroup of GG) compatible with φ\varphi.

When G=GLnG=\textup{GL}_{n}, G,par\mathcal{M}^{\textup{par}}_{G,\mathcal{L}} classifies, in addition to an \mathcal{L}-twisted Higgs bundle (𝒱,φ)(\mathcal{V},\varphi), a point xXx\in X and a full flag of the fiber 𝒱x\mathcal{V}_{x}. Such a full flag is the same data as a chain of coherent subsheaves 𝒱i𝒱\mathcal{V}_{i}\subset\mathcal{V}

𝒱(x)=𝒱0𝒱1𝒱n1𝒱n=𝒱\mathcal{V}(-x)=\mathcal{V}_{0}\subset\mathcal{V}_{1}\subset\cdots\subset\mathcal{V}_{n-1}\subset\mathcal{V}_{n}=\mathcal{V}

such that 𝒱i/𝒱i1\mathcal{V}_{i}/\mathcal{V}_{i-1} has length 11 supported at xx. The compatibility condition between φ\varphi and the full flag requires that φ\varphi restrict to a map 𝒱i𝒱i\mathcal{V}_{i}\to\mathcal{V}_{i}\otimes\mathcal{L} for each 0in0\leq i\leq n.

We have an analog of the Hitchin fibration

fG,par:G,par𝒜G,×Xf^{\textup{par}}_{G,\mathcal{L}}:\mathcal{M}^{\textup{par}}_{G,\mathcal{L}}\to\mathcal{A}_{G,\mathcal{L}}\times X

that records also the point xXx\in X in the data, in addition to a=fG,(,φ)𝒜G,a=f_{G,\mathcal{L}}(\mathcal{E},\varphi)\in\mathcal{A}_{G,\mathcal{L}}.

54 Theorem (See [GS, Th 3.3.3]).

Suppose deg>2g1\deg\mathcal{L}>2g-1, then there is a natural action of W~=𝕏(T)W\widetilde{W}=\mathbb{X}_{*}(T)\rtimes W on the restriction of the complex 𝐑f!par\mathbf{R}f^{\textup{par}}_{!}\mathbb{Q}_{\ell} to a certain open subset (𝒜G,×X)(\mathcal{A}^{\heartsuit}_{G,\mathcal{L}}\times X)^{\prime} of 𝒜G,×X\mathcal{A}^{\heartsuit}_{G,\mathcal{L}}\times X. 888When char(k)=0\textup{char}(k)=0, one can take (𝒜G,×X)=𝒜G,×X(\mathcal{A}^{\heartsuit}_{G,\mathcal{L}}\times X)^{\prime}=\mathcal{A}^{\heartsuit}_{G,\mathcal{L}}\times X; in general, the restriction to the open subset (𝒜G,×X)(\mathcal{A}^{\heartsuit}_{G,\mathcal{L}}\times X)^{\prime} does not limit applications to questions about affine Springer fibers.

4.5.2. Extended symmetry

Just as in the case of affine Springer fibers, the W~\widetilde{W}-action in Theorem 54 may be extended to an action of the graded version of the double affine Hecke algebra, see [GS, Th 6.1.6]. Also, there is a product formula relating the fiber a,xpar\mathcal{M}^{\textup{par}}_{a,x} over (a,x)𝒜G,×X(a,x)\in\mathcal{A}^{\heartsuit}_{G,\mathcal{L}}\times X and the product of the affine Springer fiber 𝒴ax\mathcal{Y}_{a_{x}} with affine Springer fibers 𝒳ay\mathcal{X}_{a_{y}} for yxy\neq x. The induced W~\widetilde{W}-action on the stalk (𝐑f!par)a,xH(a,xpar)(\mathbf{R}f^{\textup{par}}_{!}\mathbb{Q}_{\ell})_{a,x}\cong\textup{H}^{*}({\mathcal{M}^{\textup{par}}_{a,x}}) is compatible with the affine Springer action on H(𝒴ax)\textup{H}^{*}({\mathcal{Y}_{a_{x}}}) in Theorem 24. This connection between W~\widetilde{W}-actions on the cohomology of Hitchin fibers and affine Springer fibers can be used to prove results about affine Springer actions. See [YSph] for such an application.

4.6. Exercises

4.6.1.

Suppose G=SLnG=\textup{SL}_{n}. Describe the Hitchin fibers over 𝒜G,\mathcal{A}^{\heartsuit}_{G,\mathcal{L}} in terms of spectral curves.

4.6.2.

Suppose G=SLnG=\textup{SL}_{n}. Compute the dimension of 𝒜G,\mathcal{A}_{G,\mathcal{L}} and of a Hitchin fiber a\mathcal{M}_{a}. When =ωX\mathcal{L}=\omega_{X}, check that they have the same dimension. This is a numerical evidence that the Hitchin fibration in this case is a Lagrangian fibration.

4.6.3.

Describe the Hitchin base for G=SOnG=\textup{SO}_{n}.

4.6.4.

For G=SOnG=\textup{SO}_{n}, describe Hitchin fibers in terms of spectral curves.

4.6.5.

Let CC be a rational curve over kk with a unique singularity x0x_{0} which is unibranched (i.e., the preimage of x0x_{0} in the normalization C~1\widetilde{C}\cong\mathbb{P}^{1} is a single point). Let P¯x0\overline{P}_{x_{0}} be the moduli space of fractional ideals for the completed local ring 𝒪^C,x0\widehat{\mathcal{O}}_{C,x_{0}}. Show that there is a canonical homeomorphism

P¯x0Pic¯(C).\overline{P}_{x_{0}}\to\overline{\textup{Pic}}(C).

Explain why this is a special case of Proposition 53.

References

  • \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry \ProcessBibTeXEntry