Lectures on Springer theories and orbital integrals
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.
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 | local field | global field | |
| symmetry | |||
| extended sym | graded AHA | graded DAHA | graded DAHA |
| rep theory | characters of | orbital integrals | trace formula |
| when | for | for over |
Here stands for the Affine Hecke Algebra, while DAHA stands for the Double Affine Hecke Algebra; denotes an algebraic curve over ; and 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
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 and 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 instead of , we will be using the étale cohomology with coefficients in -adic sheaves (usually the constant sheaf ) on these varieties. We denote the étale cohomology of a scheme over with -coefficients simply by . Readers not familiar with étale cohomology are encouraged to specialize to the case and understand as the singular cohomology of with -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 be an algebraically closed field. Let be a connected reductive group over whose adjoint group is simple (so the adjoint group is determined by one of the seven series of Dynkin diagrams). Assume that is large compared to . Let be the rank of .
Sometimes it will be convenient to fix a Cartan subalgebra of , or equivalently a maximal torus . Once we have done this, we may talk about the roots of the -action on .
Let be the flag variety of : this is the -homogeneous projective variety parametrizing Borel subgroups of . Choosing a Borel subgroup , we may identify with .
Let be the Lie algebra of . For , let denote the centralizer of in , i.e., the stabilizer at of the adjoint action of on .
Let be the subvariety of nilpotent elements. This is a cone: it is stable under the action of on by scaling. It is known that there are finitely many -orbits on under the adjoint action.
1.2. Springer fibers
1.2.1. The Springer resolution
The cotangent bundle classifies pairs where and is a Borel subgroup of such that , where is the nilpotent radical of . The Springer resolution is the forgetful map
sending to . This map is projective.
For , the fiber is called the Springer fiber of . By definition, is the closed subscheme of consisting of those Borel subgroups such that is contained in the nilpotent radical of .
1.2.2. The Grothendieck alteration
Consider the variety of pairs where and such that . The forgetful map
is called the Grothendieck alteration 111The term “alteration” refers to a proper, generically finite map whose source is smooth over ., also known as the Grothendieck simultaneous resolution.
Let be the categorical quotient of by the adjoint action of . Then we have a commutative diagram
| (1) |
Here is the natural quotient map and sends to the image of in (where with nilpotent radical ), which can be canonically identified with (upon choosing a Borel containing ). The diagram (1) is Cartesian when restricting the left column to regular elements 222An element is called regular if its centralizer in has dimension , the rank of . in . In particular, if we restrict the diagram (1) to the regular semisimple locus 333An element is regular semisimple if for any root . Denote by the open subscheme of regular semisimple elements in . Since is stable under , it is the full preimage of an open subset . The open subset is by definition the locus of regular semisimple elements of ., we see that is a -torsor; i.e., there is an action of on preserving the projection to making the fibers of into principal homogeneous spaces for . The map becomes branched but still finite over the regular locus .
For , we denote the fiber simply by . Restricting to , the fibers for a nilpotent element is the closed subvariety of consisting of those such that . Clearly is a subscheme of , and the two schemes have the same reduced structure. However, as schemes, and are different in general. See §1.3.2 below and Exercise 1.7.2.
1.3. Examples of Springer fibers
1.3.1.
When , .
1.3.2. Regular nilpotent elements
The unique dense open -orbit consists of regular nilpotent elements, i.e., those such that . When is a regular nilpotent element, is a single point: there is a unique Borel subgroup of whose Lie algebra contains . What is this Borel subgroup?
By the Jacobson-Morosov theorem, we may extend to an -triple in . The adjoint action of on has integer weights, and it decomposes into weight spaces , . Let . This is a Borel subalgebra of , and the corresponding Borel subgroup is the unique point in .
When is regular, the fiber of the Grothendieck alteration is a non-reduced scheme whose underlying reduced scheme is a point. The coordinate ring of is isomorphic to the coinvariant algebra , which is, interestingly, also isomorphic to the cohomology ring with -coefficients of the complex flag variety .
1.3.3.
When for some vector space of dimension , is the moduli space of full flags . The Springer fiber consists of those flags such that .
1.3.4.
Consider the case and under the standard basis of . Then is the union of two s: the first consisting of flags with varying inside the fixed plane ; the second consisting of flags with a varying containing .
1.3.5.
Consider the case and the nilpotent element under the standard basis . If a flag is in , then , and . We denote . There are two cases:
-
(1)
. Then we may choose to be any line in and to be any hyperplane containing . We get a closed subvariety of isomorphic to . We denote this closed subvariety of by .
-
(2)
. This defines an open subscheme of . Suppose the line is spanned by for some , then the image of in is spanned by . Fixing , the choices of are given by . Once is fixed, is also fixed. Therefore is isomorphic to the total space of the line bundle over .
From the above discussion we see that has dimension , is an irreducible component of and so is the closure of , which we denote by . We have and is the diagonal inside .
1.3.6. Components of type Springer fibers
When , Spaltenstein [Spa76] and Steinberg [St] gave a description of the irreducible components of using standard Young tableaux of size . 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 whose Jordan type is a partition of . This means, if the partition is , has Jordan blocks of sizes . We shall construct a (non-algebraic) map , where is the discrete set of standard Young tableau for the partition . For each full flag such that , induces a nilpotent endomorphism of . Let be the Jordan type of the on , then is a partition of . The increasing sequence of partitions satisfies that is obtained from by increasing one part of by 1 (including creating a part of size ). This gives an increasing sequence of subdiagrams of the Young diagram of . We label the unique box in by 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 is an irreducible component. Moreover, all irreducible components of arise in this way and they all have the same dimension , where is the conjugate partition of . In particular, the top dimensional cohomology has dimension equal to , which is also the dimension of an irreducible representation of the symmetric group . This statement is a numerical shadow of the Springer correspondence, which says that is naturally an irreducible representation of .
Spaltenstein [Spaltenstein, Ch II, Prop 5.9] also showed that there exists a stratification of into affine spaces such that is a union of strata. This implies that the restriction map on cohomology is surjective.
1.3.7.
Consider the case for some symplectic vector space of dimension , then is the moduli space of full flags
such that for . The Springer fiber consists of those flags such that for all .
Consider the case where . We choose a basis for such that the symplectic form on satisfies if and , and for . Let be the nilpotent element in given by . Then a flag in must satisfy , and this is the only condition for it to lie in (Exercise 1.7.3). Such a corresponds to a line , hence a point in . Over this we have a tautological rank two bundle whose fiber at is the two-dimensional vector space . The further choice of gives a point in the projectivization of . The exact sequence gives an exact sequence of vector bundles over . Therefore is isomorphic to , and 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 of codimension 2 in , which is called the subregular nilpotent orbit. For , it is known that is a union of ’s whose configuration we now describe. We may form the dual graph to whose vertices are the irreducible components of 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 is of adjoint type. Let be another adjoint simple group whose type is defined as follows. When is simply-laced, take . When is of type , , and , take to be of type , , and respectively. One can show that is always a union of whose dual graph is the Dynkin diagram of . The rule in the non-simply-laced case is that each long simple root corresponds to 2 or 3 ’s while each short simple root corresponds to a unique . Such a configuration of ’s is called a Dynkin curve, see [StConj, §3.10, Definition and Prop 2] and [Slodowy, §6.3].
For example, when is of type , then is a chain consisting of ’s: with 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 . Assume is simply-laced. One can construct a transversal slice through of dimension two such that consists of regular elements except . Then is a normal surface with a Kleinian singularity at of the same type as the Dynkin diagram of . 444A Kleinian singularity is a surface singularity analytically isomorphic to the singularity at of the quotient of by a finite subgroup of .. The preimage turns out to be the minimal resolution of , and hence is the union of exceptional divisors. Upon identifying the components of with simple roots of , the intersection matrix of the exceptional divisors is exactly the negative of the Cartan matrix of . 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 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 have the same dimension.
1.4.3. The dimension formula
Let be the dimension of . Steinberg [St, Thm 4.6] and Springer [Springer] showed that
| (2) |
1.4.4. Centralizer action
The group acts on (where by dilation). Let be the stabilizer. Then acts on . Note that always surjective onto with kernel .
The action of on factors through the finite group . Note that depends not only on the isogeny class of , but on the isomorphism class of . For example, for regular, where is the center of . The action of on further factors through the image of in the adjoint group .
1.4.5. Purity
Springer [SpringerPure] proved that the cohomology of is always pure (in the sense of Hodge theory when , or in the sense of Frobenius weights when is a finite field).
1.4.6.
Let . Consider the restriction map induced by the inclusion . Then the image of is exactly the invariants of under . This is a theorem of Hotta and Springer [HottaSpringer, Theorem 1.1]. In particular, when is of type , such restriction maps are always surjective.
1.4.7. Parity vanishing
De Concini, Lusztig and Procesi [DLP] proved that vanishes for all odd and any . When , they prove a stronger statement: vanishes for odd and is torsion-free for even .
1.5. The Springer correspondence
Let be the Weyl group of . In 1976, Springer [Springer] made the fundamental observation that there is natural -action on , even though does not act on as automorphisms of varieties.
3 Theorem (Springer [Springer, Thm 6.10]).
-
(1)
For each nilpotent element , there is a natural graded action of on that commutes with the action of .
-
(2)
For each nilpotent element and each irreducible representation of , the multiplicity space is either zero or an irreducible representation of under the action in part (1).
-
(3)
Each irreducible representation of appears as for a unique pair up to -conjugacy. The assignment thus gives an injection
(4)
1.5.1. Convention
In fact there are two natural actions of on that differ by tensoring with the sign representation of . In these notes we use the action that is normalized by the following properties. The trivial representation of corresponds to regular nilpotent and the trivial . The sign representation of corresponds to . Note however that Springer’s original paper [Springer] uses the other action.
1.5.2. The case
Taking , Springer’s theorem gives a graded action of on . What is this action? First, this action can be seen geometrically by considering instead of . In fact, since , the right action of on induces an action of on , which then induces an action of on . Since the projection is an affine space bundle, it follows that . It can be shown that under this isomorphism, the action of on corresponds exactly to Springer’s action on .
Let be the graded symmetric algebra where has degree . The reflection representation of on then induces a graded action of on . Recall Borel’s presentation of the cohomology ring of the flag variety
| (5) |
where is the ideal spanned by elements of positive degree, and denotes the ideal of generated by -invariants on . Then (5) is in fact an isomorphism of -modules (see [Springer, Prop 7.2]). By a theorem of Chevalley, is isomorphic to the regular representation of , therefore, as a -module, is also isomorphic to the regular representation of .
6 Remark.
The target set in (4) can be canonically identified with the set of isomorphism classes of irreducible -equivariant local systems on nilpotent orbits. In fact, for an irreducible -equivariant local systems on a nilpotent orbit , its stalk at gives an irreducible representation of the centralizer which factors through . Note that the notion of -equivariance changes when varies in a fixed isogeny class. It is possible to extend the above injection (4) into a bijection by supplementing with 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 and, when 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 of large characteristics or characteristic zero. To state it, we need to use the language of constructible (complexes of) -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 is a perverse sheaf on whose endomorphism ring is canonically isomorphic to the group algebra . In particular, acts on the stalks of , i.e., acts on for all .
We sketch three constructions of the -action on .
1.5.3. Construction via middle extension
This construction (or rather the version where is replaced by ) is due to Lusztig [LuGreen, §3]. The dimension formula for Springer fibers (2) imply that
-
•
The map is semismall. 555A proper surjective map of irreducible varieties is called semismall (resp. small) if for any , has codimension at least (resp. ) in .
There is an extension of the dimension formula (2) for the dimension of valid for all elements . Using this formula one can show that
-
•
The map is small.
As a well-known fact in the theory of perverse sheaves, the smallness of (together with the fact that is smooth) implies that is a perverse sheaf which is the middle extension of its restriction to any open dense subset of . Over the regular semisimple locus , is a -torsor, therefore is a local system shifted in degree that admits an action of . By the functoriality of middle extension, admits an action of . Taking stalks of , we get an action of on for all .
In particular, for a nilpotent element , we get an action of on , because and 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 .
1.5.4. Construction via Fourier transform
By the semismallness of , the complex is also a perverse sheaf. However, it is not the middle extension from an open subset of . There is a notion of Fourier transform for -equivariant sheaves on affine spaces [Laumon]. One can show that is isomorphic to the Fourier transform of and vice versa. The -action on then induces an action of on by the functoriality of Fourier transform. Taking the stalk of at we get an action of on . This action is normalized according to our convention in §1.5.1.
1.5.5. Construction via correspondences
Consider the Steinberg variety which classifies triples , where are Borel subgroups of and . The irreducible components of are indexed by elements in the Weyl group: for , letting be the closure of the graph of the -action on , then 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 from each . It is nontrivial to show that these endomorphisms together form an action of on . The key ingredient in the argument is still the smallness of the map . After the -action on is defined, one then define the Springer action on by either twisting the action of on the stalk 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 in the nilpotent Steinberg variety .
Note that the above three constructions all allow one to show that , hence giving a proof of Theorem 7.
1.5.6. Construction via monodromy
We sketch a construction of Slodowy [Slo4, §4] which works for . This construction was conjectured to give the same action of on 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 -action as in Theorem 3.
Let and let be a transversal slice to the orbit of . Upon choosing an -triple containing , there is a canonical choice of such a transversal slice , where is the centralizer of in . Now consider the following diagram where the squares are Cartesian except for the rightmost one
| (8) |
Here the rightmost square is (1), and and are the preimages of and under . Let be a small ball around and let be an even smaller ball around . Let . Then the diagram (8) restricts to a diagram
| (9) |
Here , and and are the preimages of and under the vertical maps in (8). The key topological facts here are
-
•
The inclusion admits a deformation retract, hence it is a homotopy equivalence;
-
•
The map is a trivializable fiber bundle (in the sense of differential topology).
Now a general fiber of admits a homotopy action of by the second point above because the rightmost square in (9) is Cartesian over and the map is a -torsor over . By the first point above, has the same homotopy type with , hence also has the same homotopy type as a general fiber of because is a fiber bundle. Combining these facts, we get an action of on the homotopy type of , which is a stronger structure than an action of on the cohomology of . A consequence of this construction is that the -action on in Theorem 3 preserves the ring structure.
1.5.7. Proof of Theorem 3 assuming Theorem 7
We decompose the perverse sheaf into isotypical components under the -action
where is the space on which acts via the irreducible representation , and is a perverse sheaf on . Since , we conclude that each is nonzero and that
| (10) |
The decomposition theorem [BBD, Th 6.2.5] implies that each is a semisimple perverse sheaf. Therefore (10) implies that is simple. Hence is of the form where is a nilpotent orbit and is an irreducible -equivariant local system on . Moreover, since for , the simple perverse sheaves 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 -equivariant local systems on nilpotent orbits. If and , the semismallness of allows us to identify the stalk with an -isotypic subspace of . This proves part (2) of Theorem 3. ∎
We give some further examples of the Springer correspondence.
1.5.8. Type
When , all are trivial. The Springer correspondence sets a bijection between irreducible representations of and nilpotent orbits of , both parametrized by partitions of . In §1.3.6 we have seen that if has Jordan type , the top dimensional cohomology has a basis indexed by the standard Young tableaux of , the latter also indexing a basis of the irreducible representation of corresponding to the partition . For example, for , the Springer correspondences reads
-
•
trivial representation regular orbit, partition ;
-
•
two-dimensional representation subregular orbit, partition ;
-
•
sign representation {0}, partition .
1.5.9. The subregular orbit and the reflection representation
Consider the case is a subregular nilpotent element. In this case, the component group can be identified with the automorphism group of the Dynkin diagram of introduced in §1.3.8 (see [Slodowy, §7.5, Proposition]). After identifying the irreducible components of with the vertices of the Dynkin diagram of , the action of on is by permuting the basis given by irreducible components in the same way as its action on the Dynkin diagram of . For example, when , we may write with , , each intersecting in a point and otherwise disjoint. The group is isomorphic to , and its action on fixes the fundamental class of and permutes the fundamental classes of and .
Note that always has dimension , the rank of . In fact, as a -module, is isomorphic to the reflection representation of on . In other words, under the Springer correspondence, the pair corresponds to the reflection representation of .
1.6. Comments and generalizations
1.6.1. Extended symmetry
The -action on can be extended to an action of a larger algebra in various ways, if we use more sophisticated cohomology theories. On the equivariant cohomology , there is an action of the graded affine Hecke algebra (see Lusztig [LuGrHk]). On the -equivariant -group of , 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 and are replaced with the unipotent variety and itself. When is large, the exponential map identifies with in a -equivariant manner, hence the theories of Springer fibers for nilpotent elements and unipotent elements are identical. The group version of the perverse sheaf 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 .
1.6.3. Partial Springer resolutions
We may define analogs of in partial flag varieties. Let be a partial flag variety of classifying parabolic subgroups of of a given type. There are two analogs of the map one may consider.
First, instead of considering , we may consider , which classifies pairs such that , where is the nilpotent radical of . Let be the first projection. In general this map is not surjective, its image is the closure of a nilpotent orbit . The orbit is characterized by the property that its intersection with is dense in , for any . This is called the Richardson class associated to parabolic subgroups of type . When is of type , each nilpotent class is the Richardson class associated to parabolic subgroups of some type (not unique in general). The map is a resolution of singularities. For general , not every nilpotent orbit is Richardson.
Second, we may consider the subscheme classifying pairs such that , where is the nilpotent cone of . The projection is now surjective, and is a partial resolution of singularities. The Springer resolution can be factored as
We have an embedding . We may consider the fibers of either or as parabolic analogs of Springer fibers. We call them partial Springer fibers. The Springer action of on the cohomology of has an analog for partial Springer fibers. For more information, we refer the readers to [BM].
1.6.4. Hessenberg varieties
The Grothendieck alteration admits a generalization where is replaced with another linear representation of .
Fix a Borel subgroup of . Let be a representation of and be a subspace which is stable under . Now we use the pair instead of the pair , we get a generalization of the Grothendieck alteration. More precisely, let be the subscheme consisting of pairs such that . Let be the first projection. The fibers of 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 , determine the sizes of the Jordan blocks of a regular and subregular nilpotent element of .
1.7.2.
For and , calculate the coordinate ring of the non-reduced Springer fiber for a regular nilpotent element . Show also that the Springer fiber is indeed a reduced point.
Hint: if you write as an upper triangular matrix, then lies in the big Bruhat cell of the flag variety , from which you get coordinates for your calculation.
1.7.3.
Verify the statement in §1.3.7: consider , with the symplectic form given by if and for . Let . Then any flag in must satisfy
Moreover, this is the only condition for a flag to lie in .
1.7.4.
Let . Let be a Borel subgroup.
-
(1)
Let be a simple root. Let be a parabolic subgroup whose Levi factor has semisimple rank one with roots . Let be the partial flag variety of classifying parabolics conjugate to . Restricting the projection to , we get a map
Show that the pullback on cohomology induces an isomorphism
(11) where is the simple reflection associated with , which acts on via Springer’s action.
-
(2)
Can you generalize (11) to other partial flag varieties?
1.7.5.
Let and let be a subregular element. Calculate the action of on the two dimensional 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 . Calculate the Springer correspondence for explicitly.
1.7.7.
Using the dimension formula for , verify that the Springer resolution is semismall.
1.7.8.
Let and let be the Jordan decomposition of . Let . This is a Levi subgroup of . Let be the flag variety of and let be the Springer fiber of viewed as a nilpotent element in . Show that .
1.7.9.
1.7.10.
Show that is rationally smooth; i.e., its intersection cohomology complex is isomorphic to the shifted constant sheaf .
Hint: the largest direct summand in is the shifted IC sheaf of , and it is also the restriction of a direct summand of .
1.7.11.
Show that the Springer fibers are connected.
Hint: the of the Springer fibers are packed in some sheaf.
1.7.12.
Denote the simple roots of by . A parabolic subgroup is called of type if the roots of its Levi quotient are .
-
(1)
Let and let be a parabolic subgroup of type . Let be the nilpotent radical of . Show that is dense in .
-
(2)
Let . Show that for each , there are finitely many parabolics of type such that . For each such , the subvariety of is isomorphic to , and is called a curve of type .
-
(3)
For parabolics of type and , show that is either empty or a point.
-
(4)
Let be a parabolic subgroup of type such that . For any , intersects exactly curves of type .
-
(5)
Show that has the configuration described in §1.3.8.
-
(6)
Use Exercise 1.7.4 to calculate the Springer action of on .
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 -vector space fixed by a -linear transformation, affine Springer fibers classify (chains of) lattices in an -vector space fixed by an -linear transformation, where . 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 be the field of formal Laurent series in one variable . Then has a discrete valuation such that and its valuation ring is .
2.1.1. The loop group
The positive loop group is a group-valued functor on -algebras. For any -algebra , we define . It turns out that is representable by a scheme over which is not of finite type.
For example, when , an element in is given by formal Laurent series (), with , subject to one condition that (which is a polynomial in the of degree ) is invertible in . Therefore in this case is an open subscheme in the infinite dimensional affine space with coordinates , and .
Similarly we may define the loop group to be the functor on -algebras . The functor is no longer representable by a scheme, but rather by an ind-scheme. An ind-scheme is an inductive limit in the category of schemes, i.e., form an inductive system of schemes over , and is the functor . When , we may define to be the subfunctor of such that consists of -by- invertible matrices with entries in . Then the same argument as in the case shows that is representable by a scheme over . For , we have a natural closed embedding , and in this case is the inductive limit . For general , see [BL, §1] and [Faltings, §2].
2.1.2. The affine Grassmannian
The affine Grassmannian of is defined as the sheafification of the presheaf in the category of -algebras under the fpqc topology. In particular, we have .
When , the affine Grassmannian can be identified with the moduli space of projective -submodules such that
| (12) |
for some . Such an -module is called a lattice in . For fixed , let be the subfunctor of classifying those such that (12) holds, then is representable by a projective scheme over . The natural closed embeddings make into an inductive system of projective schemes, and is representable by the ind-scheme .
Let us elaborate on the bijection between and the lattices in . Let be the standard lattice. Let be the set of lattices in (in the case a lattice is simply an -submodules of of rank ). The group acts on by . This action is transitive and the stabilizer of the standard lattice is . Therefore this action induces a -equivariant bijection
| (13) |
For general , is always representable by an ind-scheme of the form where are projective schemes over , and the transition maps are closed embeddings. We have a canonical exhaustion of by projective schemes given by the affine Schubert stratification, which we now recall. The action of on by left translation has orbits indexed by dominant cocharacters . We denote by the -orbit through . Let be the closure of . Then is a projective scheme and is the union of . 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 of is an example of a class of subgroups called parahoric subgroups. Fix a Borel subgroup and let be the preimage of under the map given by reduction modulo . Then is an example of an Iwahori subgroup of . General Iwahori subgroups are conjugates of in . Like , Iwahori subgroups are group subschemes of of infinite type. Parahoric subgroups are connected group subschemes of 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 are in bijection with subsets of the Dynkin diagram of , the -conjugacy classes of parahoric subgroups of are in bijection with proper subsets of the vertices of the extended Dynkin diagram of , which has one more vertex than the Dynkin diagram of . 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 admits a canonical exact sequence of group schemes
where is the pro-unipotent radical of and is a reductive group over , called the Levi quotient of . If corresponds to a subset of the vertices of , then the Dynkin diagram of the Levi quotient is the sub-diagram of spanned by .
2.1.4. Affine flag varieties
For each parahoric subgroup we may define the corresponding affine partial flag variety as the fpqc sheafification of the functor on the category of -algebras. This functor is also representable by an ind-scheme where each is a projective scheme over and the transition maps are closed embeddings. The affine Grassmannian is a special case of for .
Consider the special case is an Iwahori subgroup of . When is simply-connected, is its own normalizer, and we may identify as the moduli space of Iwahori subgroups of , hence giving an intrinsic definition of the affine flag variety. We usually denote by Fl or and call it the affine flag variety of , with the caveat that is canonically independent of the choice of only when is simply-connected.
Let be two parahoric subgroups of . Then we have a natural projection . The fibers of this projection are isomorphic to the partial flag variety of corresponding to its parabolic subgroup given by the image of . In particular, there is a natural projection whose fibers are isomorphic to the flag variety .
2.1.5. The case of
We have seen in §2.1.2 that the affine Grassmannian of has an interpretation as the moduli space of lattices. In fact, parahoric subgroups of and the associated affine partial flag varieties can also be described using lattices. Here we consider the case .
Recall that the set of lattices in is denoted by . For any two lattices we may define their relative length to be the integer
Let be a non-empty subset. Let be the preimage of under the projection . A periodic -chain of lattices is a function sending each to a lattice such that
-
•
for all ;
-
•
for in ;
-
•
for all .
Let be the set of periodic -chains of lattices. For each , let to be the simultaneous stabilizers of all ’s. Then is a parahoric subgroup of . We call such a parahoric subgroup of type . In fact all parahoric subgroups of arise from a unique periodic -chain of lattices, for a unique non-empty . Therefore we get a bijection between and the set of parahoric subgroups of . In particular, is the parahoric subgroup corresponding to the periodic -chain of lattices given by , where .
The extended Dynkin diagram of is a loop with nodes which we index cyclically by the set , such that corresponds to the extra node compared to the usual Dynkin diagram. Parahoric subgroups of type corresponds to the proper subset of the nodes of the extended Dynkin diagram.
One can find the moduli space of periodic -chains of lattices such that . Fixing any parahoric subgroup of type , can be identified with the affine partial flag variety . In particular, the affine flag variety Fl for can be identified with the moduli space of periodic full chains of lattices, i.e., a sequence of lattices in with and for all .
2.1.6. The case of
Now consider , where is equipped with a symplectic form. We extended the symplectic form on -linearly to a symplectic form on . For a lattice , define its symplectic dual to be the set . Then is again a lattice in . The operation defines an involution on .
Let be a non-empty subset stable under multiplication by . Let be the preimage of under the natural projection . A periodic self-dual -chain of lattices in is a periodic -chain of lattices (i.e., an element in in the notation of §2.1.5) satisfying the extra condition that
Denote the set of periodic self-dual -chains of lattices in by . This is a set with an action of . For any , the simultaneous stabilizer of the ’s is a parahoric subgroup of , and every parahoric subgroup of arises this way. For a parahoric subgroup of type , the corresponding affine partial flag variety can be identified with the moduli space of periodic self-dual -chains of lattices so that as -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 -algebra , we denote by . In particular, is the Lie algebra of the loop group . For , let denote its adjoint action on .
Let be a regular semisimple element 666Here we are dealing with a Lie algebra over the non-algebraically-closed field . An element is regular semisimple if it is regular semisimple as an element in , see the footnote in §1.2.2. Equivalently, is regular semisimple if its image in lies in .. We consider the subfunctor of whose value on a -algebra is given by
| (14) |
Then is a closed sub-ind-scheme of . Let be the underlying reduced ind-scheme of . We call the affine Springer fiber of in the affine Grassmannian .
2.2.2. Alternative definition in terms of lattices
We consider the case . Let be a regular semisimple matrix. As in 2.1.2 we identify with the moduli space of lattices in , or more precisely is the set of lattices in . Then can be identified with those lattices such that , i.e., those stable under the endomorphism of given by .
When , classifies lattices in such that . The affine Springer fiber in this case is cut out by the same condition .
When , classifies lattices in such that , see §2.1.6. The affine Springer fiber in this case is also cut out by the same condition .
We give the simplest examples of affine Springer fibers.
2.2.3.
Let such that the reduction is regular semisimple. For each cocharacter , the element gives a point which lies in since . The reduced ind-scheme is in fact the discrete set which is in bijection with . More canonically, there is an action of the loop group on given by its left translation action on . This action factors through the quotient and realizes as a -torsor.
2.2.4.
Consider the case and . Then is an infinite chain of ’s. More precisely, for each , we consider the subscheme of classifying lattices such that . Then . We have is an infinite chain of ’s. The components and intersect at one point and otherwise the components are disjoint.
Here is a way to calculate the -points of . We use the Iwasawa decomposition for :
According to this decomposition, any point in can be represented by
| (15) |
for some and a unique , and has a well-defined image in . Since
the condition is the same as requiring . Therefore where consists of elements of the form (15) with . Therefore each can be identified with . It is easy to check that .
2.2.5.
Consider the case and . Then consists exactly of those lattices such that . Therefore . Details of these calculations are left to the reader, see Exercises 2.8.2.
2.2.6. Invariance under conjugation
If are related by for some , then the left multiplication by on restricts to an isomorphism , hence also . Therefore the isomorphism type of is invariant under -conjugation on . Recall we have map . For a regular semisimple point , the fiber is a single -conjugacy class (here we are using that the residue field of is algebraically closed). Therefore, the isomorphism type of depends only on .
Unlike Springer fibers, can be empty for certain . The affine Springer fiber is nonempty if and only if . In fact, if , then hence . Conversely, we have a Kostant section of which identifies with , where is a regular -triple in . Therefore, for any , , and contains the unit coset in hence nonempty; since is isomorphic to , it is also nonempty.
For , we also write for . The above discussion shows that all Springer fibers are isomorphic to for .
2.2.7. Parahoric versions
For each parahoric subgroup , we may similarly define the closed sub-indscheme using the analog of the condition (14) with replaced by . The reduced ind-scheme is called the affine Springer fiber of of type . In particular, when is an Iwahori subgroup of , we denote by .
For two parahoric subgroups of , the natural projection induces a map . In particular we have a map .
2.3. Symmetry on affine Springer fibers
For the Springer fiber , the centralizer acts on it. In this subsection we investigate a similar structure for affine Springer fibers.
2.3.1. Centralizer action
Let be the centralizer of in (the algebraic group over obtained from by base change). Then is an algebraic group over . Since is regular semisimple, is a torus over . One can define the loop group of as the functor on -algebras.
We claim that acts on the ind-scheme . This can be seen on the level of -points. Suppose and . Then the coset still satisfies
using that is in the centralizer of . Therefore . The assignment for and defines an action of on . It induces an action of on the reduced structure .
2.3.2. The split case
We consider the case where . In this case , and
where is the loop group of the multiplicative group . For any -algebra , . It is easy to see that an element is invertible if and only if starts with finitely many nilpotent coefficients and the first non-nilpotent coefficient is invertible in . When is reduced, the leading coefficient of must be invertible in , which implies , and . We see that the reduced ind-scheme , and that as schemes, where . Therefore, when , we have , and is an affine scheme of infinite type.
2.3.3. The lattice
For a general regular semisimple , let be the -rational cocharacter lattice of the torus . For each viewed as a homomorphism defined over , we may consider the element . The assignment defines an injective homomorphism
whose image is denoted by . It can be shown that the quotient is an affine scheme that is a finite disjoint union of for some integer .
2.3.4. The case
We continue with the setup of §2.2.2. We assume that . Then the characteristic polynomial of is separable. The -algebra is then a product of fields , with . Each field extension is obtained by adjoining a root of an irreducible factor of , and is necessarily of the form since . Then the centralizer is isomorphic to the product of induced tori
We have , and the map is given by
The quotient is isomorphic to . Since each is isomorphic to , we have an exact sequence , and hence the quotient contains the group scheme with finite index. Here is isomorphic to as a scheme, except that we are renaming the uniformizer .
Alternatively, we may fix a uniformizer (for example take ) and let . The lattice will be useful in calculating orbital integrals, see §3.3.2.
2.3.5. The case
Let and where is odd. Then consists of matrices with and . Note that this equation forces , hence . The torus is non-split and splits over the quadratic extension . The lattice . Writing and , we see that , and once and are fixed, the higher coefficients of can be solved uniquely using the Taylor expansion of . Therefore is isomorphic to , given by .
2.3.6. Symmetry on affine Springer fibers
Ngô has found a more precise statement about the action of on , namely the action factors through a canonical finite-dimensional quotient. We sketch the story following [NgoFL, §3.3]. Let be the image of under . We assume for otherwise is empty.
There is a smooth affine group scheme over called the regular centralizer group scheme. It is characterized by the property that its pullback to via , denoted , maps into the universal centralizer group scheme over , and this map is an isomorphism over the regular locus . Let be pullback of under the map . Then is a smooth affine group scheme over whose -fiber is the torus (i.e., is an integral model of over ). We may form the positive loop group of as well as its affine Grassmannian (also called the local Picard group). The reduced group scheme is finite-dimensional and locally of finite type. Ngô showed that the action of on (and hence on ) factors through the local Picard group , 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 is equipped an extra -action coming from the loop rotation on by dilation on the uniformizer . For more information on homogeneous affine Springer fibers and their application to representation theory, see [OY].
2.4.1.
When and . When is even, consists exactly of those lattices such that
In this case, coincides with the closure of the -orbit in corresponding to the coweight .
When is odd, consists exactly of those lattices such that
In this case, consider instead the affine Grassmannian of , which contains as a component. Then can be identified with the closure of the -orbit in corresponding to the coweight .
In either case, we have .
2.4.2. The Lusztig-Smelt examples
Let and with characteristic polynomial , where . If a lattice is stable under , it carries an action of the ring . Let , then with and . Then the integral closure of in is . The action of on makes it a one-dimensional -vector space. We fix a -linear isomorphism , under which a lattice in stable under is simply a fractional -ideal, i.e., a finitely generated -submodule of . Then can be identified with the moduli space of fractional -ideals.
The centralizer is simply , which acts on the set of fractional -ideals by multiplication. This action clearly factors through , which is the group of -points of the local Picard group scheme .
There is an action of on (by field automorphisms) given by scaling (so gets weight under this action). This induces an action of on . The fixed points of on correspond to fractional ideals generated by monomials of . More precisely, if a fractional -ideal is fixed by , define which is a subset of stable under adding and , because is an -module. Therefore is a finitely generated module for the monoid . The assignment gives a bijection
Any -submodule of contains all sufficiently large integers. Therefore any two such -module and differ by finitely many elements, and we can define . Fox any fixed , we have a total of fixed points with . For a fixed point corresponding to an -module , consider the subvariety . Then is isomorphic to an affine space whose dimension can be expressed combinatorially in terms of . The cells give a stratification of . This gives a way to compute the Poincaré polynomial of connected components of . For more details, and the similar picture for , see [LS].
2.4.3.
We look at the geometry of in a special case of §2.4.2. We consider the case and . Introducing the variable with and as before, then with fraction field . The affine Springer fiber classifies a chain of fractional -ideals . We consider a component of , classifying chains as above with , .
We first study the -fixed points on . For each -fixed -submodule of , we denote its associated module for the monoid (see §2.4.2) by a sequence of integers. For example, stands for the standard lattice . Note that the sequence for any -fixed -fractional ideal is either consecutive or has at most one gap at the second place . We have the following four fixed points in :
-
(1)
;
-
(2)
;
-
(3)
;
-
(4)
.
Our indexing scheme is that is obtained from by changing the lattice .
Then is the union of three irreducible components , and each component is isomorphic to . They all contain and that is the only intersection between any two of them. We have for .
There is a natural way to index affine partial flag varieties of by subsets , as we saw in §2.1.5. Let be the affine Springer fiber of in . Under the projection , the curves for collapse to a point, and the other curves map isomorphically onto their images.
2.4.4.
Let where is a symplectic space of dimension over , and assume . Fix a decomposition into Lagrangian subspaces of , such that the symplectic form restricts to the natural pairing on . Consider where (viewed as a self-adjoint map ) and (viewed as a self-adjoint map ). The condition that is regular semisimple is equivalent to that: (1) both and are isomorphisms; (2) is regular semisimple, or equivalently is regular semisimple.
The affine Springer fiber classifies self-dual lattices which are stable under (see §2.2.2). Consider the action of on such that has weight and has weight . This induces a -action on and on . We first consider the fixed points . A lattice is fixed under if and only if it is the -adic completion of
where is a filtration of such that for and for , and is a similar filtration of , such that
| (16) |
under the duality pairing between and . The last condition reflects the fact that is self-dual under the symplectic form.
A lattice is then determined by two filtrations of and of , dual in the sense (16), with the extra condition that
We summarize the data into the following diagram
For example, when , hence , there are two possibilities. The first possibility is , which corresponds to the standard lattice . The second possibility is , and is a line satisfying , i.e., is an isotropic line under the quadratic form . There are two such lines , giving two other -fixed points. Therefore, consists of 3 points. We have where , the two components intersect at the standard lattice, and each contains one of the remaining -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 be the partial affine flag variety of classifying pairs of lattices such that and . Let be as in §2.4.4. Then the same acts on , and the fixed points can be described by two pairs of filtrations and , where is the kind of filtration of and as described in §2.4.4, and is similar except that (16) is replaced by . Moreover, the inclusion is equivalent to and , for all .
Consider for example and we fix the dimension of the filtrations:
Such filtrations are determined by the complete flag satisfying and that . In other words, is an isotropic line in under the quadratic form , and is an isotropic line in under the quadratic form . The pair determines a point in , the product of conics defined by and . The incidence relation defines a curve of bidegree in , which is then a curve of genus one. Therefore, a connected component of is a curve of genus one. Consider the points in that contract to this curve, and take its closure . One can show that . Hence contains an irreducible component which is irrational. We refer to the appendix of [KL] for more details.
2.4.6. “Subregular” affine Springer fibers
When or is one-dimensional, we may call them subregular affine Springer fibers, by analog with subregular Springer fibers discussed in §1.3.8. If , it is a union of ’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 (see Example 2.2.4).
2.5. Geometric Properties of affine Springer fibers
2.5.1. Non-reducedness
The ind-scheme is never reduced if is nontrivial and is regular semisimple in . For example, in the case considered in §2.2.3, is isomorphic to . We have seen in §2.3.1 that for non-reduced rings , elements in can have nilpotent leading coefficients. Therefore is not just , which is . This shows that is non-reduced, hence is non-reduced.
The next theorem is the fundamental finiteness statement about .
17 Theorem (Essentially Kazhdan and Lusztig [KL, Prop 2.1]).
Let be a regular semisimple element. Then
-
(1)
There exists a closed subscheme which is projective over , such that .
-
(2)
The ind-scheme is a scheme locally of finite type over .
-
(3)
The action of on is free, and the quotient (as an fppf sheaf on -algebras) is representable by a proper algebraic space over .
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 be a projective subscheme as in (1). To show (2), we would like to show that any has an open neighborhood which is a scheme of finite type. By the -action we may assume . Since is of finite type, the set is finite. Let , then is an open neighborhood of , hence an open neighborhood of . Moreover, is contained in the finite union , hence contained in some Schubert variety . Hence is an open subset of the projective scheme , therefore is itself a scheme of finite type. To show (3), note that the fppf sheaf quotient is a separated algebraic space because it is the quotient of by the étale equivalence relation (given by the action and projection maps). By (1), there is a surjection from a projective scheme , which implies that is proper.
2.5.3. Proof of (1) when lies in a split torus
We first consider the case where lies in a split torus. By -conjugation, we may assume . In this case . Fix a Borel subgroup containing and let be the unipotent radical of . The Iwasawa decomposition of gives
Let . It is enough to show that lies in some affine Schubert variety , for then its closure in satisfies the condition in (1). For later use, we note that the translations for are disjoint and cover .
Fixing an ordering of the positive roots of with respect to , we may write an element uniquely as
| (18) |
where and is the root group corresponding to . To show that is contained in an affine Schubert variety, it suffices to give a lower bound for the valuations of appearing in (18) for any such that . We may expand in terms of the root decomposition
where is a fixed basis for each root space and is an -valued polynomial function in and linear in . Induction on the height of shows that takes the following form
| (19) |
where is a polynomial involving only , and homogeneous of degree (we define ).
Let . This is finite because is regular semisimple. If , i.e., , then induction on the height of shows that
which gives the desired lower bound and shows that 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 over which can be conjugated into a split torus. Let be the affine Grassmannian of defined using the field in place of (so that ), and let be the corresponding affine Springer fiber. Then both and carry an action of induced from its action on , and we have a closed embedding .
Let be the lattice constructed using the field . There is an action of on with fixed points . This action induces an action of on , but it does not respect the embedding . The fixed points may not lie in , however it always contains with finite index.
From the proof in the split case in §2.5.3 we have a finite type locally closed subscheme coming from the Iwasawa decomposition, such that can be decomposed as the disjoint union (not as schemes but as constructible sets). We identify with an element in then makes sense for . One checks that for even though does not respect the embedding . Therefore . Choosing representatives for the finite coset space , we see that is contained in . Since is of finite type, is a projective subscheme of whose -translations cover . Finally the projective subscheme of 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 be the -orbit of , for . This is called a semi-infinite orbit, because it has infinite dimension and also has infinite codimension in . Let , then in the notation of §2.5.3. The formula (19) implies that (or equivalently , or ) if and only if for all roots , and if so, is isomorphic to an almost affine space (namely an iterated -bundle) of dimension
Here is the determinant of the adjoint action of on . Therefore can be decomposed into almost affine spaces of the same dimension indexed by . However, this decomposition is not a stratification: the closure of will intersect other but certainly not a union of such ’s.
The decomposition in the split case has a generalization. Suppose is a parabolic subgroup of with unipotent radical and a Levi subgroup . Let and suppose is regular semisimple as an element in . Using the generalized Iwasawa decomposition , there is a well-defined map sending to , for and . 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 we get
where is the affine Springer fiber for and the group . The fibers of , if non-empty, are almost affine spaces of dimension , where
| (20) |
Assume , then the connected components of are indexed by where is the coroot lattice of . If we decompose into connected components for and taking their preimages under , we get a decomposition into locally closed sub-ind-schemes indexed also by . One can show that , if non-empty, is an almost affine space bundle over with fiber dimension .
2.5.6. Connectivity and equidimensionality
When is simply-connected, Fl is connected, and in this case the affine Springer fiber is also connected. See [KL, §4, Lemma 2]. As a consequence, when is simply-connected, is connected for all parahoric because the natural projection is surjective.
In [KL], it is also shown that 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 is well-defined, and is the dimension of as an algebraic space. To state a formula for , we need some more notation.
Consider the adjoint action . The kernel of this map is , and the induced endomorphism on is invertible. Let be the determinant of . This is consistent with our earlier definition of in the case lies in a split torus .
On the other hand, recall is the group of -rational cocharacters of , which is also the rank of the maximal -split subtorus of . Let
Then is also the rank of the maximal -anisotropic subtorus .
21 Theorem (Bezrukavnikov [Bez], conjectured by Kazhdan-Lusztig [KL]).
Let be a regular semisimple element. Then we have
| (22) |
2.5.8. Sketch of proof
A key role in the proof is played by the notion of regular points of . We have an evaluation map sending to the reduction of modulo , which is well-defined up to the adjoint action by . We say is a regular point if lies in the open substack of . Let be the open sub-ind-scheme of consisting of regular points. It can be shown that is non-empty. Denote the preimage of in by , then the projection map is an isomorphism. Since is equidimensional as mentioned in §2.5.6, we see that . Of course we have , therefore we must have . It remains to calculate the dimension of .
Recall we have defined the local Picard group in §2.3.6. The action of on preserves , and in fact is a torsor under . Therefore it suffices to compute the dimension of .
Consider the projection . Pullback along we get a finite morphism , for a finite flat -algebra . There is a close relationship between the group scheme and , where the Weyl group acts diagonally on both and . In fact and are equal up to connected components in their special fibers. In particular, , where . It is not hard to see that
where is the normalization of . 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 is in fact dense in . In particular, each irreducible component of 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 , or in the sense of Hodge structures if ). 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 . We say if and have the same image under the map . In [NgoFL, Prop 3.5.1], it is shown that for fixed , there exists some (depending on ) such that whenever , we have isomorphisms
in a way compatible with the actions. Therefore, we may say that varies locally constantly with under the -adic topology on .
For example, consider the case and let correspond to a characteristic polynomial whose roots are in and are distinct modulo . Then for any , the characteristic polynomial of also has distinct roots modulo . In this case, and are both torsors under .
2.6. Affine Springer representations
In this subsection we introduce an analog of Springer’s -action on in the affine situation.
2.6.1. The affine Weyl group
We view as a group of automorphisms of the cocharacter lattice , where is a fixed maximal torus of . The extended affine Weyl group is the semidirect product
When is simply-connected, so that is spanned by coroots, is a Coxeter group with simple reflections in bijection with the nodes of the extended Dynkin diagram of . In general, is a semidirect product of the affine Weyl group (where is the coroot lattice), which is a Coxeter group, and an abelian group . The group naturally acts on the affine space by affine transformations, where acts by translations.
24 Theorem (Lusztig [L96], Sage [Sage]).
There is a canonical action of on .
Since is not of finite type, the -adic homology is understood as the inductive limit , whenever we present as a union of projective subschemes .
2.6.2. Sketch of the construction of the -action
We consider only the case is simply-connected so that is generated by affine simple reflections . For each parahoric subgroup we have a corresponding affine Springer fiber . For containing a fixed Iwahori subgroup , we have a projection .
Let be the Levi quotient of and . We have an evaluation map defined as follows. For such that , we send the coset to the image of under the projection . This is well-defined up to the adjoint action of . We have a Cartesian diagram
where is the Grothendieck alteration for the reductive group . By the Springer theory for , we have a -action on the direct image complex (where stands for the dualizing complex for ). By proper base change, we get an action of on , and hence on . Taking a standard parahoric corresponding to the -th node in the extended Dynkin diagram, then , and we get an involution acting on . To check the braid relation between and for neighboring nodes and , we may choose a standard parahoric such that and the braid relation holds because acts on . This shows that acts on . ∎
Despite the simplicity of the construction of the -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 on may not be semisimple, as we shall see in the next example.
2.6.3. An example in
Consider the case and the element . This is a subregular case. The affine Springer fiber has two irreducible components and both isomorphic to . Here parametrizes chains of lattices where is the standard lattice and is varying. The other component parametrizes chains of lattices where and is varying. The fundamental classes give a natural basis for the top homology group . One can show that the action of the affine Weyl group on takes the following form under the basis and :
We see that is a nontrivial extension of the sign representation of by the trivial representation spanned by . One can also canonically identify the -module with the affine coroot lattice of the loop group .
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 -adic groups .
2.7.2. Extended symmetry
The -action on can be extended to an action of the wreath product . For homogeneous affine Springer fibers (those admitting a torus action coming from loop rotation), the equivariant cohomology group admits an action of the graded double affine Hecke algebra, which is a deformation of . For details we refer to [OY]. Vasserot and Varagnolo [Vass] [VV] constructed an action of the double affine Hecke algebra on the -groups of affine Springer fibers.
2.7.3. The group version
Taking instead of in , one can similarly define the group version of affine Springer fibers, which we still denote by with reduced structure . For a -algebras , we have
| (25) |
However, in the group version, the definition above admits an interesting generalization. Recall the -double cosets in are indexed by dominant cocharacters . For we denote the corresponding double coset by , which is the preimage of the Schubert stratum under the projection . Similarly we may define to be the preimage of the closure of . One can replace the condition in (25) by or , and take reduced structures to obtain reduced generalized affine Springer fibers and . We have an open embedding , whose complement is the union of for dominant coweights . The motivation for introducing is to give geometric interpretation of orbital integrals of spherical Hecke functions on .
A.Bouthier has established the fundamental geometric properties of , parallel to Theorem 17 and Theorem 21.
26 Theorem (Bouthier [Bouthier]).
Let be regular semisimple, and let .
-
(1)
The generalized affine Springer fiber is non-empty if and only if , where is the Newton point of , see [KV, §2].
-
(2)
The ind-scheme is locally of finite type.
-
(3)
We have
where is half the sum of positive roots, and and 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 .
2.7.4.
In [KV], Kottwitz and Viehmann defined two generalizations of affine Springer fibers for elements in the Lie algebra .
2.7.5.
As an analog of Hessenberg varieties, one can also consider the following situation. Let be a linear representation of a reductive group over . Let be an -lattice stable under . For we may define a sub-ind-scheme of
Let be the reduced structure of . The cohomology of these ind-schemes are related to orbital integrals that appear in relative trace formulae.
2.8. Exercises
2.8.1.
Let for some vector space over equipped with a quadratic form . Give an interpretation of the parahoric subgroups and affine partial flag varieties of in terms of self-dual lattice chains in , in the same style as in §2.1.6.
2.8.2.
2.8.3.
Let and . Describe .
2.8.4.
Let and . Construct a nontrivial -action on involving loop rotations (i.e., the action scales ) and determine its fixed points.
2.8.5.
Let and . Let be the diagonal torus, then . What is the regular locus (see §2.5.8)? Study the -orbits on .
2.8.6.
In the setup of §2.3.3, show that the action of on is free, which implies that its action on is free. Show also that the permutation action of on the set of irreducible components of is free.
2.8.7.
For , let be the Levi subgroup consisting of block diagonal matrices with sizes of blocks , . Let be regular semisimple as an element in . What is the invariant (see (20)) in terms of familiar invariants of the characteristic polynomials of the ?
2.8.8.
Let and , with pairwise distinct and . Describe the affine Springer fibers and .
Note: this is a good exercise if you have a whole day to kill.
2.8.9.
For and as in Example 2.4.4, describe the -fixed points on and .
2.8.10.
Verify the calculations in §2.6.3.
2.8.11.
Let and let . Describe the affine Springer fiber . What is the action of on in terms of the basis given by the irreducible components of ?
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 -adic groups. Just as conjugacy classes of a finite group are fundamental to understanding its representations, orbital integrals are fundamental to understanding representations of -adic groups. In certain cases, orbital integrals can be interpreted as counting points on affine Springer fibers.
3.1. Integration on a -adic group
3.1.1. The setup
Let be a local non-archimedean field, i.e., is either a finite extension of or a finite extension of . Then has a discrete valuation which we normalize to be surjective. Let be the valuation ring of and be the residue field. Therefore, unlike in the previous sections, is a finite field. We assume that is large with respect to the groups in question.
3.1.2. Haar measure and integration
Let be an algebraic group over . The topological group is locally compact and totally disconnected. It has a right invariant Haar measure which is unique up to a scalar. For a measurable subset , we denote its volume under by . Fixing a compact open subgroup , we may normalize the Haar measure so that has volume . For example, if we choose an integral model of over , we may take .
With the Haar measure one can integrate smooth (i.e., locally constant) compactly supported functions on with complex values. We denote this function space by (where stands for Schwarz). For , the integral
can be calculated as follows. One can find a subgroup of finite index such that is right -invariant, i.e., for all and (see Exercise 3.7.1). Then the integral above becomes a weighted counting in the coset :
It is easy to check that the right side above is independent of the choice of as long as is right -invariant and has finite index in .
3.1.3. Variant
Let be a subgroup defined over together with a Haar measure on it. Consider a function , i.e., is a left -invariant, locally constant function on whose support is compact modulo , we may define the integral
| (27) |
This integral is calculated in the following way. Again we choose a finite index subgroup such that is right -invariant. Then the integral (27) can be written as a weighted sum over double cosets :
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 by to emphasize that it is a vector space over . Let and . Consider the map given by . Then the composition is a smooth function on . Then is locally constant, left invariant under the centralizer of in , and has compact support modulo .
Fix Haar measures on and on . The following integral is then a special case of (27) (except that we write the integration variable explicit below while not in (27))
Such integrals are called orbital integrals on the Lie algebra . We may similarly define orbital integrals on the group by replacing with an element in and with an element in .
3.2.2. Specific situation
For the rest of the section we will restrict to the following situation. Let be a split reductive group over . We may fix an integral model of by base changing the corresponding Chevalley group scheme from to . In the following we will regard as a reductive group scheme over . We normalize the Haar measure on by requiring that have volume .
The Lie algebra contains a canonical lattice coming from the integral model over . We will be most interested in the orbital integral of the characteristic function of the lattice .
3.2.3. The centralizer of
Suppose is regular semisimple so that its centralizer is a torus over . Let , which is a complete discrete valuation field whose residue field is algebraically closed. We continue to let denote a uniformizer of , which is also a uniformizer in . Using , the construction in §2.3.3 gives an embedding whose image we still denote by . This embedding being -equivariant, carries an action of . We think of as an étale group scheme over , then the notation makes sense, and it is just the -invariants in if we regard the latter as a plain group. Then is a discrete and cocompact subgroup.
3.2.4. Centralizers in
Let and let be a regular semisimple element in which is not necessarily diagonalizable over . Assume either or . As in §2.3.4, the characteristic polynomial of is separable, hence the -algebra is isomorphic to a product of fields . We have
In this case, the lattice consists of elements of the form for . The quotient is isomorphic to . Each fits into an exact sequence
| (28) |
where is the ramification degree of the extension , therefore the quotient is compact.
3.2.5. Orbital integrals in terms of counting
Consider the following subset of
This is a set-theoretic version of the affine Springer fiber.
The group acts on by the rule . For any free abelian group , its action on by left translation is free (because the stabilizers are necessarily finite), hence it acts freely on .
More generally, for any discrete cocompact subgroup , the quotient groupoid is finitary, i.e., it has finitely many isomorphism classes and the automorphism group of each object is finite. For a finitary groupoid , we define the cardinality of to be
| (29) |
The next lemma follows directly from the definitions, whose proof is left to the reader as Exercise 3.7.2.
30 Lemma.
Let be a regular semisimple element in . Let be any discrete cocompact subgroup. We have
with the cardinality on the right side interpreted as in (29).
3.2.6. The case and fractional ideals
We continue with the situation in §3.2.4. Under the identification of with the set of -lattices in (see (13)), we have
The bijection sends to the lattice .
We give another interpretation of . Let be the characteristic polynomial of . Let
be the commutative -subalgebra of generated by . The ring of total fractions is a finite étale -algebra of degree , and is an order in it. The canonical action of on realizes as a free -module of rank . Recall a fractional -ideal is a finitely generated -submodule . If we choose an element as a basis for the -module structure, we get a bijection
| (31) |
which sends to the -lattice .
Using the algebra , we have a canonical isomorphism
This isomorphism intertwines the action of on by left translation and the action of on the set of fractional -ideals by multiplication.
When happens to be a product of Dedekind domains (i.e., is the maximal order in ), all fractional -ideals are principal, which is the same as saying that the action of on is transitive. In general, principal fractional ideals form a homogeneous space under ; the difficulty in counting in general is caused by the singularity of the ring .
3.3. Relation with affine Springer fibers
From this subsection we restrict to the case is a local function field, i.e., for a finite field . Let be regular semisimple. The definitions of the affine Grassmannian and the affine Springer fiber we gave in §2 make sense when the base field is a finite field, so we have a sub-ind-scheme of , both defined over .
The following lemma is clear from the definitions.
33 Lemma.
The set of -rational points is the same as the set defined in §3.2.5, both as subsets of .
3.3.1.
If we base change from to , by Theorem 17 we know that is a proper algebraic space over . The proof there actually shows that this algebraic space is defined over , which we denote by . We emphasize here that is viewed as an étale group scheme over whose -points is the plain group used to be denoted .
3.3.2. The case of
In the situation of §3.2.6, upon choosing uniformizers , we defined the lattice . Base change from to , we may similarly define a lattice as in §2.3.4, using the same choice of uniformizers (note that may split into a product of fields, but will project to a uniformizer in each factor). The -action on gives it the structure of an étale group scheme over , just as . We have . There is an analog of Theorem 17 if we replace with . In particular, is a proper algebraic space over admitting a surjective map from a projective scheme.
34 Theorem.
Let . Let be a regular semisimple element in . We fix the Haar measure on such that its maximal compact subgroup gets volume . Then we have
| (35) |
3.3.3. -points of a quotient
We consider a quotient stack where is a scheme over and is an algebraic group over acting on . Then, by definition, is the groupoid of pairs where is a left -torsor, and is an -equivariant morphism. The isomorphism class of the -torsor is classified by the Galois cohomology . For each class , let be an -torsor over with class . We may define a twisted form of over by . We also have the inner form of acting on the -scheme . It is easy to see that -equivariant morphisms are in bijection with . Therefore we get a decomposition of groupoids
| (36) |
3.3.4. Proof of Theorem 34
3.4. Stable orbital integrals
The setup is the same as §3.3. For general , the generalization of the formula (35) is not straightforward. Namely, the orbital integral by itself does not have a cohomological interpretation. The problem is that we may not be able to find an analog of with vanishing first Galois cohomology so that is simply . In view of formula (36), a natural fix to this problem is to consider the twisted forms of altogether. This suggests taking not just the orbital integral but a sum of several orbital integrals .
3.4.1. Stable conjugacy
Fix a regular semisimple element . An element is called stably conjugate to if it is in the same -orbit of . Equivalently, is stably conjugate to if . For stably conjugate to , one can attach a Galois cohomology class which becomes trivial in . The assignment gives a bijection of pointed sets
| (37) |
3.4.2. The case of
3.4.3. The case of
We consider the case . Let . Let
be two regular semisimple elements in . Since they have the same determinant , they are stably conjugate to each other. However, they are not conjugate to each other under . One can show that the stable conjugacy class of consists of exactly two -orbits represented by and , see Exercise 3.7.3.
3.4.4. Definition of stable orbital integrals
Let . We define the stable orbital integral of with respect to to be
| (38) |
where runs over -orbits of elements that are stably conjugate to . For stably conjugate to , we have a canonical isomorphism as -groups. Therefore, once we fix a Haar measure on , we get a canonical Haar measure on the other centralizers . It is with this choice that we define in (38).
3.4.5. Stable part of the cohomology
The quotient group scheme acts on . The component group is an étale group scheme over whose -points acts on . We define the stable part of this cohomology group to be the invariants under this action
It turns out that if we replace with any other -stable free abelian subgroup commensurable with , 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) |
Here is the parahoric subgroup of the torus .
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 . Let be the maximal compact subgroup. Then there is a canonical smooth group scheme over whose -fiber is and whose points is . This group scheme is the finite type Néron model for the torus . Now let be the open subgroup scheme obtained by removing the non-neutral component of the special fiber of . Then, by definition, . The positive loop group is, up to nilpotents, the neutral component of .
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 -subgroup commensurable with , which maps onto the étale group scheme over (and the kernel is necessarily finite). The group is an analog of defined in §3.3.2. We form the quotient stack . Then we have an isomorphism
| (41) |
In fact, by the discussion in the end of §3.4.5, the left side above does not change if we replace by a commensurable lattice, so by shrinking we may assume . On the other hand, the right side above can be computed by the Leray spectral sequence associated with the map which is a torsor under the finite discrete group , therefore . Since surjects onto by the choice of , we see that
from which (41) follows.
By the discussion in §3.3.3 and (36), we have
| (42) |
The inclusion gives
The first surjection follows from Lang’s theorem because the quotient is connected; the second follows from another theorem of Lang which says that vanishes 777See [Serre, Ch.X, §7, p.170, Application and Example (b)]. Let be a complete discrete valuation field with perfect residue field, and its maximal unramified extension. Then Lang’s theorem asserts that is a -field. Therefore for any torus over .. For each such that , one can show that the image of in is trivial. Therefore, by (37), to each we can attach an element stably conjugate to , unique up to -conjugacy, such that . One can show that is in bijection with the set . Therefore, (42) implies
| (43) |
where the sum is over the -orbits of those stably conjugate to . Applying Lemma 30 to the discrete cocompact subgroup , we have
Plugging this into the right side of (43), we get
| (44) |
By the Grothendieck-Lefschetz trace formula, is equal to the alternating Frobenius trace on , which, by (41), can be identified with . Therefore the theorem follows from the identity (44) together with the volume identity
| (45) |
To show this, let (where is the connected Néron model of whose points is , see §3.4.6). This is a finite étale group over . We have a short exact sequence of group ind-schemes over
The associated six term exact sequence for -cohomology gives
from which we get (45), using that . ∎
3.5. Examples in
By Theorem 39, in order to calculate orbital integrals, we need to know not just the geometry of the affine Springer fiber , 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 and assume . We will compute several orbital integrals in this case and verify Theorem 39 in these cases by explicit calculations.
3.5.1. Unramified case:
Let be a regular semisimple element with .
Let be the centralizer of in . Then is an unramified quadratic extension of obtained by adjoining . Therefore we have , with . We have , which is compact. We fix a Haar measure on with total volume 1.
Let be the affine Springer fiber of , which is a scheme over . Lemma 30 implies that
In §2.2.4 we have shown that is an infinite union of rational curves indexed by the integers . Since is diagonalizable over , each component is in fact defined over . The lattice is contained in , and is generated by the uniformizer . We label the components so that sends to . Let , which is a -point of .
The action of the nontrivial involution on is by inversion, hence it also acts on by inversion. The standard lattice lies in both and , hence it is the point . Therefore the point is fixed by since it is defined over . Since the action of on is compatible with its action on (by inversion), the only possibility is that
The action of can be represented by the picture
Here each double line represents a . From this we see that consists of only one point , namely the standard lattice . This implies that
| (46) |
3.5.2. Unramified case:
Now consider the element . In Exercise 3.7.3 we see that is stably conjugate to but not conjugate to under . However, is conjugate to under . Therefore, the affine Springer fiber still looks the same as over , but the action of is different.
Consider the component of whose -points consist of -stable lattices such that . This component is cut out by conditions defined over , so it is stable under , and we call this component . We label the other components of by () so that the generator sends to . Let . Since the action of on is compatible with its action on (by inversion), the only possibility is that
The action of can be represented by the picture
Therefore no point is defined over . The component is the only one that is defined over , and it has to be isomorphic to over because it is so over (there are no nontrivial Brauer-Severi varieties over a finite field). We see that has elements. Therefore
| (47) |
Adding up (46) and (47) we get
3.5.3. Unramified case: cohomology
The quotient is a nodal rational curve obtained from by glueing two -points into a nodal point.
Now let us consider the quotient . Over this is also a nodal rational curve consisting of a unique node which is image of all . While is a -point of the quotient , none of its preimages are defined over . Therefore, the points in still map injectively to the quotient, in addition to the point . We conclude that consists of points. Since is connected, the stable part of the cohomology of is the whole , and the alternating sum of Frobenius trace on it is the cardinality of . In this special case we have verified the formula (40). We remark that the action of on the 1-dimensional space is by , and the Grothendieck-Lefschetz trace formula for Frobenius reads instead of , the latter being the number of -points on a nodal rational curve obtained by identifying two -points on .
3.5.4. Ramified case: orbital integrals
Consider the elements and where . Again and are stably conjugate but not conjugate in .
Let be the centralizer of in . Then is a ramified quadratic extension of , and . Similarly, let be the centralizer of in . Then is another ramified quadratic extension of , and . We choose Haar measures on compact groups and with total volume .
3.5.5. Ramified case: cohomology
In the setup of §3.5.4, . The component group of is , but its action on is trivial. Therefore the stable part of the cohomology is the whole , on which the alternating sum of the Frobenius gives the cardinality of . However, the parahoric subgroup of has index in it ( consists of those whose reduction in is ). Therefore, the right side of formula (40) gets a factor in front of . 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 and instead of their sum? Is there a geometric interpretation of this difference analogous to Theorem 39?
3.6.1. The -orbital integral
The linear combination is an example of -orbital integrals. More generally, let be a character of , then we define the -orbital integral of to be
where the sum is over the -orbits in the stable conjugacy class of .
3.6.2. Statement of the Fundamental Lemma
The Langlands-Shelstad conjecture, also known as the Fundamental Lemma, states that the -orbital integral of for is equal to the stable orbital integral of an element for a smaller group , up to a simple factor. In formula, the Fundamental Lemma is the identity
The smaller group depends on both and , and is called the endoscopic group of . The number is called the transfer factor, which turns out to be an integer power of (depending on and ) if is chosen appropriately from its stable conjugacy class, and the measures on and are chosen properly.
3.6.3. A simple case
In the situation of §3.5.1, take the nontrivial character on , the corresponding endoscopic group is isomorphic to the torus ; but in general it is not always isomorphic to a subgroup of . The Fundamental Lemma in this case is the identity
where if we identify with .
On the other hand, in the ramified situation §3.5.4, the -orbital integral of for the nontrivial vanishes. This maybe explained without calculating the orbital integrals explicitly, for in general, must factor through a further quotient of for 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 and function field was established by B-C. Ngô [NgoFL]. There is a generalization of Theorem 39 to -orbital integrals, in which we replace the stable part of the cohomology of by the -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 and its endoscopic cousin . 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, denotes a finite field with , and .
3.7.1.
Let be an algebraic group over and let . Show that there exists a compact open subgroup such that is both left and right invariant under .
3.7.2.
Prove Lemma 30.
3.7.3.
Let and . Consider the matrices as in §3.4.3. Show that and are stably conjugate but not conjugate under .
3.7.4.
Let over and let be a block diagonal matrix in , where . Consider the asymptotic behavior of the orbital integral as tends to . Find the smallest integer such that
as . Note that the on the right side is analysts’ while on the left side it means orbital integral. Can you interpret in terms of the characteristic polynomials of the ’s?
For an explicit estimate of , see [Y-Orb].
3.7.5.
Let and . Compute .
Hint: use the cell decomposition introduced in §2.4.2.
3.7.6.
Let . Let be the characteristic function of elements such that the reduction in is regular nilpotent. Let be a regular semisimple element.
-
(1)
Show that unless .
-
(2)
When , show that
3.7.7.
Let and for odd. Let be the set of with . Fix the Haar measure on such that has volume . Show that, for any integer ,
is the same as the number of closed subschemes of the plane curve satisfying: (1) the underlying topological space of is the point ; (2) .
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 over (assumed algebraically closed) which is smooth, projective and connected.
4.1.2. The moduli stack of bundles
There is a moduli stack classifying vector bundles of rank over . For any -algebra , is the groupoid of rank vector bundles (locally free coherent sheaves) on . The stack is algebraic, see [LMB, Th 4.6.2.1]. Moreover, it is smooth and locally of finite type over .
4.1.3. -torsors
Recall a (right) -torsor over is a scheme over with a fiberwise action of , such that locally for the étale topology of , becomes the and the -action becomes the right translation action of on the first factor.
For general reductive , We have the moduli stack of -torsors over . For a -algebra , the -points of is the groupoid of -torsors over . Then is also a smooth algebraic stack locally of finite type over .
4.1.4. Associated bundles
Let be a -representation of . Let be a -torsor over . Then there is a vector bundle of whose total space is
where acts on by . The vector bundle is said to be associated to and .
When , there is an equivalence of groupoids
| (49) |
The direction sends a vector bundle to the -torsor of framings of , namely take , with the natural action of on the trivial bundle . The other direction sends a -torsor over to the vector bundle associated to and the standard representation St of . The equivalence (49) gives a canonical isomorphism of stacks .
4.1.5.
For other classical groups , -torsors have more explicit descriptions in terms of vector bundles. For example, when , a -torsor over amounts to the same thing as a pair where is a vector bundle over of rank and is an isomorphism of line bundles .
When , the groupoid of -torsors on is equivalent to the groupoid of pairs where is a vector bundle of rank over and is an -linear map of coherent sheaves that gives a symplectic form on geometric fibers. The map in one direction sends a -torsor to the pair , where is the vector bundle associated to and the standard representation St of , and the symplectic form on comes from the canonical map of -representations (where is the trivial representation).
4.1.6. Higgs bundles
Fix a line bundle over . An -twisted Higgs bundle of rank over is a pair where is a vector bundle over of rank , and is an -linear map. The endomorphism is called a Higgs field on .
There is a moduli stack classifying -twisted Higgs bundles of rank over . The morphism forgetting the Higgs field is representable. Therefore is also an algebraic stack over .
4.1.7. -Higgs bundles
An -twisted -Higgs bundle is a pair where is a -torsor over and is a global section of the vector bundle over . Here, is the vector bundle associated to and the adjoint representation of , in the sense of §4.1.4. We call an -twisted Higgs field on .
When , the notion of -twisted -Higgs bundle is equivalent to that of an -twisted Higgs bundle of rank . In fact, to each -twisted -Higgs bundle , we get a Higgs bundle , where viewed as a global section of corresponds to under the canonical isomorphism .
We also have the moduli stack of -twisted Higgs -torsors over . The -points of is the groupoid of -twisted -Higgs bundles on , where denotes the pullback of to . The forgetful morphism is representable, hence is an algebraic stack over . When , we have a canonical isomorphism of stacks .
4.1.8. Examples
When , an -twisted -Higgs bundle over amounts to the same thing as a triple where is a vector bundle over of rank , and satisfies .
When , an -twisted -Higgs bundle over amounts to the same thing as a triple where is a vector bundle over of rank , is nondegenerate fiberwise, and such that for all local sections and of , as a local section of .
4.1.9. Hitchin moduli stack as a cotangent bundle
In Hitchin’s original paper [Hi], he considered the case where is the sheaf of -forms on . This case is particularly important because is closely related to the cotangent bundle of . For a point which is a -torsor over , the cotangent complex of at is given by the derived global sections of the complex over . Using a Killing form on we may identify with , therefore the cotangent complex of at is , which lives in degrees and . In particular, when has finite automorphism group (e.g., is stable), the Zariski cotangent space at is , i.e., a cotangent vector of at is the same thing as a -twisted Higgs field on . Therefore, (properly defined) and share an open substack , where is the open substack of stable -bundles.
4.2. Hitchin fibration
4.2.1. Hitchin fibration for
For an -twisted Higgs bundle on , locally on we may view as a matrix with entries which are local sections of , and we may take the characteristic polynomial of this matrix. The coefficients of this polynomial are well-defined global sections of , . More intrinsically, induces a map , and we may take
This way we have defined a morphism
sending to . We view as an affine space over . The morphism is called the Hitchin fibration in the case .
4.2.2. Hitchin fibration in general
For general connected reductive as in §1.1, the coefficients of the characteristic polynomial in the case of are replaced with the fundamental -invariant polynomials on . Recall that . Chevalley’s theorem says that , and the latter is a polynomial ring in variables. We fix homogeneous generators of as a -algebra, whose degrees are canonically defined although are not canonical. When is almost simple, the numbers are the exponents of . Viewing as a symmetric multilinear function invariant under , for any -torsor over , induces a map of the associated bundles
This further induces
If is an -twisted Higgs field on , we may evaluate on the section of to get
The assignment defines the Hitchin fibration for
The target is again viewed as an affine space over , and is called the Hitchin base.
A more intrinsic way to define the Hitchin base is the following. The affine scheme is equipped with a -action inducing the grading on its coordinate ring. Let be the complement of the zero section in the total space of . Consider the -twisted version of over :
where acts by on the two coordinates. This is an affine space bundle over whose fibers are isomorphic to . Then can be canonically identified with the moduli space of sections of the map . In particular, every point gives a map .
4.2.3. The generically regular semisimple locus
Trivializing at the generic point of and restricting to , we have a polynomial , where is the function field of . When is a separable polynomial in , we call such an generically regular semisimple. The generic regular semisimplicity of is equivalent to the nonvanishing of the discriminant and therefore it defines an open subscheme .
For general , viewing as a map (see the end of §4.2.2), we call generically regular semisimple if sends the generic point of into the open substack . This defines an open subscheme generalizing the construction of above.
4.2.4. Geometric properties
When , the stack is smooth, see [NgoFL, Th 4.14.1]. In this situation, the morphism is flat over , see [NgoFL, Cor 4.16.4]. When is semisimple, there is a further open dense subset over which is a Deligne-Mumford stack and the map 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 be a reductive group over that fits into an exact sequence of reductive groups
Let be a representation of . Then we may consider pairs where is an -torsor over and is a section of the associated bundle . Alternatively, such a pair is the same as a morphism . One can prove that there is an algebraic stack classifying such pairs. Every -torsor induces an -torsor, hence we have a morphism . Fix an -torsor over , we denote the preimage by .
To recover the Hitchin moduli stack, we consider the case with and . Let with the action of defined as follows: acts by the adjoint representation and acts by scaling on . An -torsor is a pair consisting of a -torsor and a line bundle on . Fixing the line bundle (which is equivalent to fixing an -torsor), we get an isomorphism
For general as above, we may define the analog of the Hitchin base as follows. Let . This is the analog of , and it carries an action of . Then we form the twisted version of over
Then we define to be the moduli space of sections to the map . The morphism then induces the analog of the Hitchin fibration
4.2.6. Example
Consider , and . The action on is given by . The moduli stack then maps to by remembering only the two -torsors. Fixing , its preimage in classifies triples where is a Higgs bundle over of rank , is an -linear map . The Hitchin base in this case is the same as the classical Hitchin base . 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 can be described in “abelian” terms, namely by line bundles on certain finite coverings of . We elaborate on this observation for and .
4.3.1. The case of and the spectral curve
For , one can define a curve equipped with a degree morphism . The construction is as follows. The total space of can be written as a relative spectrum over
Let be the projection. Consider the map of coherent sheaves on
given in coordinates by . By adjunction induces a map , whose image we denote by . Then is an ideal sheaf on . Define the spectral curve to be the closed subscheme of defined by
If we trivialize on some open subset , and view as functions on , then is the subscheme of defined by one equation (where is the coordinate on ). The projection is finite flat of degree . The curve is called the spectral curve of since the fibers of are the roots of the characteristic polynomial .
When , the curve is reduced and therefore smooth on a Zariski dense open subset, there is a moduli stack classifying torsion-free coherent -modules that are generically of rank , see [AK]. The usual Picard stack classifying line bundles on is an open substack of , and it acts on by tensoring.
50 Proposition.
Suppose . Let be the fiber of over . Then there is a canonical isomorphism of stacks
4.3.2. Sketch of proof
We give the morphism . For any coherent sheaf on , the direct image is a coherent sheaf on equipped with a map because is an -module and contains as the second direct summand. When is torsion-free and generically rank , is torsion-free over (hence a vector bundle) of rank , and the map induces a Higgs field . The assignment defines the morphism , which can be shown to be an isomorphism. ∎
4.3.3. The case
In this case is a tuple with . For , we can similarly define a spectral curve as the closed subscheme of the total space of cut out by the ideal locally generated by
Note that carries an involution under which the projection is invariant. Now suppose is separable when restricted to the generic point of , so that is reduced. The involution induces an involution on the compactified Picard . Let be the relative Serre duality functor on coherent sheaves on , i.e., viewed as an -module in a natural way. Let be the moduli stack of pairs where and is an isomorphism satisfying that . We called the compactified Prym stack of with respect to the involution . Similar to the case of , we have the following description of .
51 Proposition.
Suppose such that is reduced. Then there is a canonical isomorphism of stacks
4.3.4. Sketch of proof
For , points in are triples where is a vector bundle of rank on , is a symplectic form on as in §4.1.5, and satisfies for local sections of , and the characteristic polynomial of is . By Proposition 50, the Higgs bundle gives a point . The symplectic form can be viewed as an isomorphism such that . Note that the Higgs bundle corresponds to under the isomorphism in Proposition 50. The isomorphism then turns into satisfying . ∎
4.3.5. Example 4.2.6 continued
Fix two line bundles and on , and let be the fiber of the moduli stack in §4.2.6 over . Then classifies where is a vector bundle of rank over , is a Higgs field and . We have a Hitchin-type map sending to the . Let and let be the fiber of over . Consider the same spectral curve as in §4.3.1. Using Proposition 50 we may identify with a point . The map gives a map by adjunction. Here , and the relative dualizing complex is a line bundle on because is a planar curve hence Gorenstein. Since is torsion-free and generically a line bundle, is an injective map of coherent sheaves. Hence the data turns into the data of a coherent subsheaf of the line bundle . Since is a line bundle, the subsheaf is determined by the support of the quotient , which is a zero-dimensional subscheme of . We conclude that there is a canonical isomorphism
where is the disjoint union of Hilbert schemes of zero-dimensional subschemes of 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 and , we have identified the Hitchin fiber with the compactified Picard stack of the spectral curve . The usual Picard stack acts on by tensor product. Therefore we have an action of on . This action is simply transitive if is smooth.
In the case and , the Prym stack is defined as the moduli stack of pairs where and satisfying . Then acts on by tensoring.
For general and , one can similarly define a commutative group stack that acts on the corresponding Hitchin fiber . In fact, via the map , the regular centralizer group scheme in §2.3.6, which descends to , pulls back to a smooth group scheme over which is generically a torus. The stack then classifies -torsors over .
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 . Viewing as a map , let be the preimage of under . Since , is non-empty hence the complement consists of finitely many -points of . For each point , let be the completed local ring of at , with fraction field and residue field . We fix a trivialization of near and we may identify as an element in . Let be the corresponding point in the Kostant section, then we have the affine Springer fiber in the affine Grassmannian (we put in the subscript to emphasize that the definition of the loop groups uses the field , which is isomorphic to ). The loop group of the centralizer acts on , and the action factors through the local Picard group as in §2.3.6. On the other hand, we have the action of the global Picard stack on the Hitchin fiber mentioned in §4.3.6. The product formula of Ngô roughly says that, modulo the actions of the local and global Picard stacks, is the product of the affine Springer fibers for all points .
52 Theorem (Product Formula, Ngô [NgoHit, Th 4.6] and [NgoFL, Prop 4.15.1]).
For , there is a canonical morphism
which is a homeomorphism of stacks. Here the notation (where acts on on the right and acts on on the left) means the quotient of by the action of given by , and .
In the case , 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 be a reduced and projective curve over . For each point , one can define a local Picard group whose points are , where is the completed local ring of at and its ring of total fractions (which is a product of fields in general). There is also a local analog of whose -points are the fractional -ideals of , 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 be a reduced and projective curve over . Let be the singular locus of . Then there is a canonical morphism
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 . We fix the curve and a line bundle on it as before. Let be the moduli stack classifying where , is an -twisted Higgs -bundle, and is a reduction of the fiber at to a -torsor (here is a Borel subgroup of ) compatible with .
When , classifies, in addition to an -twisted Higgs bundle , a point and a full flag of the fiber . Such a full flag is the same data as a chain of coherent subsheaves
such that has length supported at . The compatibility condition between and the full flag requires that restrict to a map for each .
We have an analog of the Hitchin fibration
that records also the point in the data, in addition to .
54 Theorem (See [GS, Th 3.3.3]).
Suppose , then there is a natural action of on the restriction of the complex to a certain open subset of . 888When , one can take ; in general, the restriction to the open subset 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 -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 over and the product of the affine Springer fiber with affine Springer fibers for . The induced -action on the stalk is compatible with the affine Springer action on in Theorem 24. This connection between -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 . Describe the Hitchin fibers over in terms of spectral curves.
4.6.2.
Suppose . Compute the dimension of and of a Hitchin fiber . When , 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 .
4.6.4.
For , describe Hitchin fibers in terms of spectral curves.
4.6.5.
Let be a rational curve over with a unique singularity which is unibranched (i.e., the preimage of in the normalization is a single point). Let be the moduli space of fractional ideals for the completed local ring . Show that there is a canonical homeomorphism
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