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

A geometric realization of the asymptotic affine Hecke algebra

Roman Bezrukavnikov Department of Mathematics Massachusetts Institute of Technology
77 Massachusetts Avenue, Cambridge, MA 02139, USA
[email protected]
Ivan Karpov Department of Mathematics Massachusetts Institute of Technology
77 Massachusetts Avenue, Cambridge, MA 02139, USA
[email protected]
 and  Vasily Krylov Department of Mathematics Harvard University and CMSA
1 Oxford Street, Cambridge, MA 02138, USA
[email protected], [email protected]
Abstract.

A key tool for the study of an affine Hecke algebra \mathcal{H} is provided by Springer theory of the Langlands dual group via the realization of \mathcal{H} as equivariant KK-theory of the Steinberg variety. We prove a similar geometric description for Lusztig’s asymptotic affine Hecke algebra JJ identifying it with the sum of equivariant KK-groups of the squares of {\mathbb{C}}^{*}-fixed points in the Springer fibers, as conjectured by Qiu and Xi (the same result was also obtained by Oron Popp using different methods). As an application, we give a new geometric proof of Lusztig’s parametrization of irreducible representations of JJ. We also reprove Braverman-Kazhdan’s spectral description of JJ. As another application, we prove a description of the cocenters of \mathcal{H} and JJ conjectured by the first author with Braverman, Kazhdan and Varshavsky. The proof is based on a new algebraic description of JJ, which may be of independent interest.

1. Introduction

1.1. Affine Hecke algebra and asymptotic affine Hecke algebra

Let GG be a reductive algebraic group over \mathbb{C}. Let WfW_{f} be the Weyl group of GG and let Λ\Lambda be the coweight lattice of GG. The (extended) affine Weyl group WW of GG is defined as W:=WfΛW:=W_{f}\ltimes\Lambda. Let =G=(W)\mathcal{H}=\mathcal{H}_{G}=\mathcal{H}(W) be the affine Hecke algebra of GG, recall that \mathcal{H} is a deformation of the group algebra of WW over [𝐯,𝐯1]\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}], where 𝐯=𝐪1/2{\bf{v}}={\bf{q}}^{1/2} and 𝐯,𝐪{\bf{v}},{\bf{q}} are formal variables. Let {\mathbfcal{H}} be the complexification of \mathcal{H}. For qq\in{\mathbb{C}}^{*} we will denote by q:=[𝐯,𝐯1]q{\mathbfcal{H}}_{q}:={\mathbfcal{H}}\otimes_{{\mathbb{C}}[{\bf{v}},{\bf{v}}^{-1}]}{\mathbb{C}}_{q} the corresponding specialization of {\mathbfcal{H}}.

Kazhdan-Lusztig and Ginzburg showed that \mathcal{H} has a geometric realization: it can be identified with the convolution algebra KG×(𝒩~×𝒩𝒩~)K_{G^{\vee}\times\mathbb{C}^{*}}(\widetilde{\mathcal{N}}\times_{\mathcal{N}}\widetilde{\mathcal{N}}), where 𝒩~𝒩\widetilde{\mathcal{N}}\rightarrow\mathcal{N} is the Springer resolution for the Langlands dual group GG^{\vee}. Kazhdan and Lusztig used the relation between {\mathbfcal{H}} and geometry of the Steinberg variety 𝒩~×𝒩𝒩~\widetilde{\mathcal{N}}\times_{\mathcal{N}}\widetilde{\mathcal{N}} to parametrize simple q{\mathbfcal{H}}_{q}-modules for qq not a root of unity. It turns out that this parametrization extends to any qq\in\mathbb{C}^{*} that is not a root of the Poincaré polynomial PWP_{W} of WW (see [61]), moreover, the parametrization does not depend on qq.

Lusztig introduced the (based) algebra JJ called the asymptotic Hecke algebra, its basis is parametrized by WW, and there is an injective homomorphism ϕ:J[𝐯,𝐯𝟏]\phi\colon{\mathcal{H}}\hookrightarrow J\otimes_{\mathbb{Z}}\mathbb{Z}[\bf{v},{\bf{v}}^{-1}]. We set 𝐉:=J{\bf{J}}:=J\otimes_{\mathbb{Z}}\mathbb{C}.

Informally, one can think about the algebra JJ as the ‘‘limit’’ of \mathcal{H} as qq goes to 0. As discovered by Lusztig, representations of 𝐉{\bf{J}} are closely related to those of q{\mathbfcal{H}}_{q}, more precisely pulling back an irreducible 𝐉{\bf{J}}-module EE under the homomorphism ϕq:q𝐉\phi_{q}\colon{\mathbfcal{H}}_{q}\hookrightarrow{\bf{J}}, qq\in{\mathbb{C}}^{*} one gets a so-called standard module KqK_{q} over q{\mathbfcal{H}}_{q}111see [39], [40] for the case of generic qq; more generally, from [61, Theorem 3.2] and [19, Corollary 2.6] it follows that this holds for qq such that PW(q)0P_{W}(q)\neq 0. We reprove and strengthen this fact below. This connection explains the observation above that the parametrization of modules over the Hecke algebra q{\mathbfcal{H}}_{q} remains independent of the parameter qq, as long as qq is not a root of PWP_{W}.

The goal of this paper, achieved in Theorem A, is to prove a similar geometric description of the algebra JJ conjectured in [51] and derive some applications (Theorems B, C etc.). We note that Theorem A was also proved by Oron Popp in his PhD thesis [50] by a different method.

The coherent realization of \mathcal{H} is a manifestation of local Langlands duality used in the proof of a special case of local Langlands conjectures [36]. It would be very interesting to find an interpretation and a generalization for Theorem A in that context. In particular, according to a recent insight of Braverman and Kazhdan [19], 𝐉{\bf{J}} can be viewed as a ring of Iwahori bi-invariant distributions on the pp-adic group intermediate between the algebra of compactly supported distributions (well-known to be isomorphic to a specialization of {\mathbfcal{H}}) and Harish-Chandra Schwartz space of tempered distributions. In [19] one also finds a generalization of this definition to not necessarily Iwahori bi-invariant distribution. We hope that Theorem A admits an (at least conjectural) extension to that generality.

1.2. Filtrations on \mathcal{H} and two-sided cells

Let us now return to the identification KG×(𝒩~×𝒩𝒩~)K_{G^{\vee}\times\mathbb{C}^{*}}(\widetilde{{\mathcal{N}}}\times_{{\mathcal{N}}}\widetilde{{\mathcal{N}}})\simeq\mathcal{H}. The group GG^{\vee} acts on 𝒩{\mathcal{N}} with a finite number of orbits ordered by the closure order. For every orbit 𝕆e𝒩\mathbb{O}_{e}\subset\mathcal{N} consider its closure 𝕆¯e\overline{\mathbb{O}}_{e} and take its preimage in 𝒩~×𝒩𝒩~\widetilde{{\mathcal{N}}}\times_{{\mathcal{N}}}\widetilde{{\mathcal{N}}}. The G×G^{\vee}\times\mathbb{C}^{*}-equivariant KK-theory of this preimage is a term in the filtration of \mathcal{H} by the two-sided ideals e\mathcal{H}_{\leqslant e} indexed by nilpotent orbits 𝕆e{\mathbb{O}}_{e}. Each subquotient e:=e/<e\mathcal{H}_{e}:=\mathcal{H}_{\leqslant e}/\mathcal{H}_{<e} is a bimodule over \mathcal{H} that is clearly isomorphic to KZe×(e×e)K_{Z_{e}\times{\mathbb{C}}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e}), where e\mathcal{B}_{e} is the fiber of the Springer resolution over ee called the Springer fiber, ZeZ_{e} is the reductive part of the centralizer of ee in GG^{\vee} and the {\mathbb{C}}^{*}-action on e\mathcal{B}_{e} is defined using the Jacobson-Morozov Theorem. Set Re:=KZe×(pt)R_{e}:=K_{Z_{e}\times{\mathbb{C}}^{*}}(\operatorname{pt}), it acts naturally on e=KZe×(e×e)\mathcal{H}_{e}=K_{Z_{e}\times{\mathbb{C}}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e}).

Remark 1.2.1.

Note that e\mathcal{H}_{e} is also a ring (without a unit element). To see the ring structure geometrically, we need to identify e\mathcal{H}_{e} with KZe×(Λe×Λe;e×e)K_{Z_{e}\times{\mathbb{C}}^{*}}(\Lambda_{e}\times\Lambda_{e};\mathcal{B}_{e}\times\mathcal{B}_{e}), where Λe\Lambda_{e} is the Slodowy variety corresponding to ee and KZe×(Λe×Λe;e×e)K_{Z_{e}\times{\mathbb{C}}^{*}}(\Lambda_{e}\times\Lambda_{e};\mathcal{B}_{e}\times\mathcal{B}_{e}) is the KK-group of the category of Ze×Z_{e}\times{\mathbb{C}}^{*}-equivariant coherent sheaves on Λe×Λe\Lambda_{e}\times\Lambda_{e} supported on e×eΛe×Λe\mathcal{B}_{e}\times\mathcal{B}_{e}\subset\Lambda_{e}\times\Lambda_{e}, see [56, Theorem B.2].

There is a way to describe the filtration e\mathcal{H}_{\leqslant e}\subset\mathcal{H} above algebraically. Let {Cw|wW}\{C_{w}\,|\,w\in W\} be the canonical basis of \mathcal{H} introduced in [35] and generalized by Lusztig to the case of extended Weyl groups in [39]. Let w\mathcal{H}_{\leqslant w} be the minimal based (i.e., spanned over \mathbb{Z} by a subset of the canonical basis) two-sided ideal of \mathcal{H} that contains CwC_{w}. In [37] Lusztig introduced a notion of the two-sided cell in WW that can be characterized as follows: two elements w,wWw,w^{\prime}\in W lie in the same two-sided cell iff w=w\mathcal{H}_{\leqslant w}=\mathcal{H}_{\leqslant w^{\prime}}. We obtain a partition of WW into two-sided cells. To a cell 𝐜W{\bf{c}}\subset W there thus corresponds the two-sided ideal 𝐜\mathcal{H}_{\leqslant{\bf{c}}}. This is the so-called cell filtration. Let us denote by 𝐜\mathcal{H}_{\bf c} the corresponding subquotient.

We obtain a partial order on the set of two-sided cells defined as follows: 𝐜𝐜{\bf{c}}^{\prime}\leqslant{\bf{c}} iff 𝐜𝐜\mathcal{H}_{\leqslant{\bf{c}}^{\prime}}{{\subset}}\mathcal{H}_{\leqslant{\bf{c}}}. By a result of Lusztig (see [40]), the set of GG^{\vee}-orbits in 𝒩{\mathcal{N}} is in bijection with the set of two-sided cells in WW. Let us denote this bijection by LL. It then follows from results of Xi [62] and of the first author [10, Theorem 55, §11.3] that if 𝐜=L(e){\bf{c}}=L(e) then 𝐜=e\mathcal{H}_{\leqslant{\bf{c}}}=\mathcal{H}_{\leqslant e} and the LL bijection above is order-preserving. Thus the cell filtration will also be referred to as the geometric filtration.

1.2.1. The direct sum decomposition for the ring JJ

The ring JJ can be decomposed as the direct sum J=𝐜J𝐜J=\bigoplus_{{\bf{c}}}J_{\bf{c}}, where 𝐜W{\bf{c}}\subset W runs over the set of two-sided cells in WW.

For 𝐜=L(e){\bf{c}}=L(e) we have a natural homomorphism of algebras KZe×(Λe×Λe;e×e)=𝐜ϕ𝐜J𝐜[𝐯,𝐯1]K_{Z_{e}\times{\mathbb{C}}^{*}}(\Lambda_{e}\times\Lambda_{e};\mathcal{B}_{e}\times\mathcal{B}_{e})=\mathcal{H}_{{\bf{c}}}\xrightarrow{\phi_{{\bf{c}}}}J_{{\bf{c}}}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}]. This homomorphism becomes an isomorphism after tensoring by (𝐯){\mathbb{C}}({\bf{v}}). Let ϕ𝐜:J𝐜[𝐯,𝐯1]\phi^{\bf{c}}\colon\mathcal{H}\rightarrow J_{\bf{c}}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}] be the composition of ϕ:J[𝐯,𝐯1]\phi\colon\mathcal{H}\rightarrow J\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}] and the projection onto J𝐜[𝐯,𝐯1]J_{\bf{c}}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}].

1.3. Main results

1.3.1. Geometric description of JeJ_{e}

One of the main results of this paper is a geometric description of JeJ_{e} conjectured by Qiu and Xi in [51], see also Propp’s thesis [50, Theorem 1.5.2] for another proof.

Theorem A.

There exists an isomorphism of rings Je[𝐯,𝐯1]KZe×(e×e)J_{e}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}]\simeq K_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}^{\mathbb{C}^{*}}\times\mathcal{B}_{e}^{\mathbb{C}^{*}}).

Remark 1.3.1.

Note that JJ does not depend on qq, this corresponds to the fact that {\mathbb{C}}^{*} acts trivially on e×e\mathcal{B}_{e}^{{\mathbb{C}}^{*}}\times\mathcal{B}_{e}^{{\mathbb{C}}^{*}}. Specializing at 𝐯=1{\bf{v}}=1 we obtain an identification JeKZe(e×e)J_{e}\simeq K_{Z_{e}}(\mathcal{B}_{e}^{\mathbb{C}^{*}}\times\mathcal{B}_{e}^{\mathbb{C}^{*}}); however, specializing at 𝐯=1{\bf{v}}=-1 yields another isomorphism.

Remark 1.3.2.

In view of Theorem A, the category CohZG(e)(e×e)\operatorname{Coh}_{Z_{G^{\vee}}(e)}(\mathcal{B}_{e}^{\mathbb{C}^{*}}\times\mathcal{B}_{e}^{\mathbb{C}^{*}}) (or CohZe(e×e)\operatorname{Coh}_{Z_{e}}(\mathcal{B}_{e}^{\mathbb{C}^{*}}\times\mathcal{B}_{e}^{\mathbb{C}^{*}})) can be viewed as a categorification of the ring JeJ_{e}. Other categorifications of JeJ_{e} were proposed in [42], [15], [13] (these three are equivalent), and partially in [22]. We do not know how these different categorifications are related.

Let us briefly describe the idea of the proof of Theorem A. Recall the (injective) homomorphisms ϕ:J[𝐯,𝐯1]\phi\colon\mathcal{H}\hookrightarrow J\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}], ϕ𝐜:𝐜J𝐜[𝐯,𝐯1]\phi_{\bf{c}}\colon\mathcal{H}_{\bf{c}}\hookrightarrow J_{\bf{c}}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}]. They induce J𝐜[𝐯,𝐯1]\mathcal{H}-J_{{\bf{c}}}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}]-bimodule and 𝐜J𝐜\mathcal{H}_{\bf{c}}-J_{\bf{c}}-bimodule structures on J𝐜[𝐯,𝐯1]J_{{\bf{c}}}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}]. Moreover, considered as a left \mathcal{H} (resp. 𝐜\mathcal{H}_{\bf{c}})-module, it is isomorphic to 𝐜\mathcal{H}_{{\bf{c}}} (see Corollary 2.3.2). We then prove the following theorem.

Theorem B.

The right action of J𝐜[𝐯,𝐯1]J_{\bf{c}}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}] induces isomorphisms of algebras

J𝐜[𝐯,𝐯1]End(J𝐜[𝐯,𝐯1])oppEnd(𝐜)oppEnd𝐜Re(𝐜)opp,J_{{\bf{c}}}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}]\,\vphantom{j^{X^{2}}}\smash{\overset{\sim}{\vphantom{\rule{0.0pt}{1.99997pt}}\smash{\longrightarrow}}}\,\operatorname{End}_{\mathcal{H}}(J_{\mathbf{c}}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}])^{\mathrm{opp}}\simeq\operatorname{End}_{\mathcal{H}}(\mathcal{H}_{\bf{c}})^{\mathrm{opp}}\simeq\operatorname{End}^{R_{e}}_{\mathcal{H}_{\bf{c}}}(\mathcal{H}_{\bf{c}})^{\mathrm{opp}},

where End𝐜Re(𝐜)\operatorname{End}^{R_{e}}_{\mathcal{H}_{\bf{c}}}(\mathcal{H}_{\bf{c}}) stands for the set of endomorphisms, which are both ReR_{e}-and 𝐜\mathcal{H}_{{\bf{c}}}-linear.

Warning 1.3.3.

Note that End𝐜(𝐜)opp\operatorname{End}_{\mathcal{H}_{\bf{c}}}(\mathcal{H}_{\bf{c}})^{\mathrm{opp}} is not isomorphic to 𝐜\mathcal{H}_{\bf{c}} because 𝐜\mathcal{H}_{\bf{c}} is a ring without a unit. We only have the natural embedding 𝐜End𝐜(𝐜)opp\mathcal{H}_{\bf{c}}\hookrightarrow\operatorname{End}_{\mathcal{H}_{\bf{c}}}(\mathcal{H}_{\bf{c}})^{\mathrm{opp}} induced by the right multiplication.

Thus, in order to establish the identification between Ke:=KZe×(e×e)K_{e}:=K_{Z_{e}\times{\mathbb{C}}^{*}}(\mathcal{B}_{e}^{{\mathbb{C}}^{*}}\times\mathcal{B}_{e}^{{\mathbb{C}}^{*}}) and Je[𝐯,𝐯1]J_{e}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}] we need to construct a homomorphism KeEnd(𝐜)oppK_{e}\rightarrow\operatorname{End}_{\mathcal{H}}(\mathcal{H}_{\bf{c}})^{\mathrm{opp}} and prove that it is an isomorphism. To construct the aforementioned homomorphism, we need to find a Ke\mathcal{H}-K_{e}-bimodule that is isomorphic to e=KZe×(e×e)\mathcal{H}_{e}=K_{Z_{e}\times{\mathbb{C}}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e}) as a \mathcal{H}-module. The natural candidate is e:=KZe×(e×e)\mathcal{F}_{e}:=K_{Z_{e}\times{\mathbb{C}}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e}^{{\mathbb{C}}^{*}}). We prove that e\mathcal{F}_{e} is indeed isomorphic to e\mathcal{H}_{e}. To see that, we use the Bialynicki-Birula type decomposition of e\mathcal{B}_{e} by the attractors via the {\mathbb{C}}^{*}-action studied by De Concini, Lusztig and Procesi in [26]. In more detail, recall that if XX is a smooth projective variety with a {\mathbb{C}}^{*}-action then by the Bialynicki-Birula theorem attractors to the connected components of XX^{{\mathbb{C}}^{*}} are affine fibrations. It follows that the decomposition of XX by the attractors induces a filtration on the (complexified) KK-theory of XX with associated graded being isomorphic to the KK-theory of XX^{{\mathbb{C}}^{*}} if H(X,)H_{*}(X^{{\mathbb{C}}^{*}},\mathbb{C}) is generated by algebraic cycles. Variety e\mathcal{B}_{e} is projective but not smooth, so the Bialynicki-Birula theorem can not be applied to it directly. On the other hand by the results of [26], e\mathcal{B}_{e}^{{\mathbb{C}}^{*}} is smooth and attractors in e\mathcal{B}_{e} of the components of e\mathcal{B}_{e}^{{\mathbb{C}}^{*}} are indeed affine fibrations. So, we obtain a filtration on KZe×(e×e)K_{Z_{e}\times{\mathbb{C}}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e}^{{\mathbb{C}}^{*}}) with associated graded being isomorphic to KZe×(e×e)K_{Z_{e}\times{\mathbb{C}}^{*}}(\mathcal{B}_{e}^{{\mathbb{C}}^{*}}\times\mathcal{B}_{e}^{{\mathbb{C}}^{*}}). We then produce a splitting of this filtration. To this end we consider closures of the subvarieties AttrF:={(x,y)e×F|limt0tx=y}\operatorname{Attr}_{F}:=\{(x,y)\in\mathcal{B}_{e}\times F\,|\,\underset{t\rightarrow 0}{\operatorname{lim}}\,t\cdot x=y\}, where FeF\subset\mathcal{B}_{e}^{{\mathbb{C}}^{*}} are connected components of e\mathcal{B}_{e}^{{\mathbb{C}}^{*}}. We then consider natural morphisms π¯F:Attr¯FF\overline{\pi}_{F}\colon\overline{\operatorname{Attr}}_{F}\rightarrow F, ι¯F:Attr¯Fe\overline{\iota}_{F}\colon\overline{\operatorname{Attr}}_{F}\rightarrow\mathcal{B}_{e} and the splitting is given by F(Ide×ι¯F)(Ide×π¯F)\bigoplus_{F}(\operatorname{Id}_{\mathcal{B}_{e}}\times\overline{\iota}_{F})_{*}(\operatorname{Id}_{\mathcal{B}_{e}}\times\overline{\pi}_{F})^{*}. In other words, the splitting is given by the natural correspondences e×eIde×π¯Fe×Attr¯FIde×ι¯Fe×e\mathcal{B}_{e}\times\mathcal{B}_{e}^{{\mathbb{C}}^{*}}\xleftarrow{\operatorname{Id}_{\mathcal{B}_{e}}\times\overline{\pi}_{F}}\mathcal{B}_{e}\times\overline{\operatorname{Attr}}_{F}\xrightarrow{\operatorname{Id}_{\mathcal{B}_{e}}\times\overline{\iota}_{F}}\mathcal{B}_{e}\times\mathcal{B}_{e}.

So, we obtain a homomorphism θ:KeEnd(e)opp=Je[𝐯,𝐯1]\theta\colon K_{e}\rightarrow\operatorname{End}_{\mathcal{H}}(\mathcal{F}_{e})^{\mathrm{opp}}=J_{e}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}], and it remains to check that it is an isomorphism. Injectivity is easy since we have an identification F:KeeF\colon K_{e}\,\vphantom{j^{X^{2}}}\smash{\overset{\sim}{\vphantom{\rule{0.0pt}{1.99997pt}}\smash{\longrightarrow}}}\,\mathcal{F}_{e} (of right KeK_{e}-modules), and our homomorphism is then given by the action of KeK_{e} on itself via right multiplication. To check surjectivity, it is enough to show that any element of End(e)\operatorname{End}_{\mathcal{H}}(\mathcal{F}_{e}) is uniquely determined by its value on F(1)F(1). To see that we use the identifications

KeKeeeeeJe[𝐯,𝐯1]K_{e}\simeq_{K_{e}}\mathcal{F}_{e}\simeq_{\mathcal{H}_{e}}\mathcal{H}_{e}\simeq_{\mathcal{H}_{e}}J_{e}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}]

and the only thing to check is that the image of 1Ke1\in K_{e} in Je[𝐯,𝐯1]J_{e}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}] after these identifications is a (left) invertible element (we use Theorem B). It is clearly right invertible, but 𝐉e{\bf{J}}_{e} is left-Noetherian, so right invertible elements are left invertible.

1.3.2. Representation theory of 𝐉{\bf{J}} from the geometric perspective

Realization of JeJ_{e} as the convolution algebra KZe(e×e)K_{Z_{e}}(\mathcal{B}_{e}^{{\mathbb{C}}^{*}}\times\mathcal{B}_{e}^{{\mathbb{C}}^{*}}) allows one to apply geometric methods to representation theory of 𝐉{\bf{J}} (cf. [36] and [20, Sections 7, 8]). Our geometric realization has particularly favorable properties since e\mathcal{B}_{e}^{{\mathbb{C}}^{*}} is smooth and projective, and its homology is generated by algebraic cycles ([26]).

In [39] Lusztig classified irreducible modules over the algebra 𝐉{\bf{J}}. He proved that these modules are parametrized by the triples (s,e,ρ)(s,e,\rho) (up to conjugation), where e𝒩e\in\mathcal{N} is a nilpotent element, ss is a semisimple element of ZeZ_{e} and ρ\rho is an irreducible representation of Γes:=ZG(s,e)/ZG(s,e)0\Gamma^{s}_{e}:=Z_{G^{\vee}}(s,e)/Z_{G^{\vee}}(s,e)^{0}. Irreducible 𝐉{\bf{J}}-module corresponding to (s,e,ρ)(s,e,\rho) will be denoted by E(s,e,ρ)E(s,e,\rho). For a variety XX equipped with the action of an algebraic group HH we set 𝐊H(X):=KH(X){\bf{K}}_{H}(X):=K_{H}(X)\otimes_{\mathbb{Z}}\mathbb{C}. We reprove Lusztig’s description using our geometric approach by proving the following theorem.

Theorem C.

(a) Irreducible modules over 𝐉e=𝐊Ze(e×e){\bf{J}}_{e}={\bf{K}}_{Z_{e}}(\mathcal{B}_{e}^{{\mathbb{C}}^{*}}\times\mathcal{B}_{e}^{{\mathbb{C}}^{*}}) are all of the form 𝐊(e,s)ρ{\bf{K}}(\mathcal{B}_{e}^{{\mathbb{C}}^{*},s})_{\rho}.

(b) We have E(s,e,ρ)=𝐊(e,s)ρE(s,e,\rho)={\bf{K}}(\mathcal{B}_{e}^{{\mathbb{C}}^{*},s})_{\rho}.

Let us briefly outline the argument. Part (a)(a) of the Theorem follows from general considerations about modules over convolution algebras (see Section A.2 for the details). To prove part (b)(b), we recall that Lusztig’s parametrization works as follows (see [40]): E(s,e,ρ)E(s,e,\rho) is the unique irreducible module over 𝐉{\bf{J}} such that ϕqE(s,e,ρ)\phi_{q}^{*}E(s,e,\rho) is isomorphic to K(s,e,ρ,q):=𝐊(esq)K(s,e,\rho,q):={\bf{K}}(\mathcal{B}_{e}^{sq}) for a generic qq. So, in order to prove part (b)(b) we need to check that ϕq(𝐊(e,s)ρ)𝐊(eqs)ρ\phi^{*}_{q}({\bf{K}}(\mathcal{B}_{e}^{{\mathbb{C}}^{*},s})_{\rho})\simeq{\bf{K}}(\mathcal{B}_{e}^{qs})_{\rho} for a generic qq. We prove that this holds for any qq. This implies part (b)(b) of the Theorem and also shows that ϕqE(s,e,ρ)=K(s,e,ρ,q)\phi_{q}^{*}E(s,e,\rho)=K(s,e,\rho,q) for every qq\in{\mathbb{C}}^{*} (this result is new for qq being a root of PWP_{W}).

1.3.3. Proof of Braverman-Kazhdan’s theorem

In [19] the authors described the algebra 𝐉e{\bf{J}}_{e} in spectral terms by formulating a version of the matrix Paley-Wiener theorem for 𝐉e{\bf{J}}_{e} (see [19, Theorem 1.8 (3)]). We reprove their theorem using our geometric approach.222We were not able to follow some steps in the argument of [19] (see footnotes 22, 33 below), we fill in the details not found in loc. cit using our present geometric methods. Stefan Dawydiak [24] developed an alternative algebraic approach to completing the proof of that theorem. Let us first recall the content of [19, Theorem 1.8 (3)].

Pick a semisimple element sZes\in Z_{e}. Let MGM^{\vee}\subset G^{\vee} be a Levi containing ss and such that eLieMe\in\operatorname{Lie}M^{\vee}. In [19, Section 1.2] the authors consider a certain family of q{\mathbfcal{H}}_{q}-modules over ZM:=Z(M)0Z_{M}:=Z(M^{\vee})^{0} that we will denote by 𝒱(ZM,s,ρ,q)\mathcal{V}(Z_{M},s,\rho,q). Its fiber over 1ZM1\in Z_{M} is IndMK(s,e,ρ,q)\operatorname{Ind}_{{\mathbfcal{H}}_{M}}^{{\mathbfcal{H}}}K(s,e,\rho,q).

Theorem 1.3.4 (Braverman-Kazhdan).

Let 𝒮e\mathcal{S}_{e} be a subalgebra of M,s,ρEnd𝒪(ZM)rat.𝒱(ZM,s,ρ,q)\prod_{M,s,\rho}\operatorname{End}^{\mathrm{rat}.}_{\mathcal{O}(Z_{M})}\mathcal{V}(Z_{M},s,\rho,q) (where the product is taken over all compact ss) given by the following conditions:
a) any φ𝒮e\varphi\in\mathcal{S}_{e} does not have poles at the points of families 𝒱(ZM,s,ρ,q)\mathcal{V}(Z_{M},s,\rho,q) which correspond to (cf. loc. cit.) non-strictly positive characters of Levi subgroups;
b) the endomorphisms φ\varphi are compatible in some precise sense (see Section 6.2 for details).

Then 𝒮eJe\mathcal{S}_{e}\simeq J_{e}.

Before sketching our proof of this theorem, let us discuss certain families of modules over {\mathbfcal{H}}, e{\mathbfcal{H}}_{e}, 𝐉e[𝐯,𝐯1]{\bf{J}}_{e}\otimes{\mathbb{C}}[{\bf{v}},{\bf{v}}^{-1}] that should be considered as geometric counterparts of the Braverman-Kazhdan’s families 𝒱(ZM,s,ρ,q)\mathcal{V}(Z_{M},s,\rho,q).

Here by a ‘‘geometric’’ family we mean the result of the following construction. Let CZG(s,e)C\subset Z_{G^{\vee}}(s,e) be a torus and let M:=ZG(C)0M^{\vee}:=Z_{G^{\vee}}(C)^{0} be the corresponding Levi. Let ρ\rho be an irreducible representation of ΓM:=ZM(s,e)/ZM(s,e)0\Gamma_{M}:=Z_{M^{\vee}}(s,e)/Z_{M^{\vee}}(s,e)^{0}. Set H:=C,sH:=\langle C,s\rangle (the smallest closed diagonalizable subgroup of GG^{\vee} containing ss and CC). We have two families over C×C\times{\mathbb{C}}^{*}:

𝒦(C,s):=𝐊H×(e)|Cs×,(C,s):=𝐊H×(e)|Cs×.\mathcal{K}(C,s):={\bf{K}}_{H\times{\mathbb{C}}^{*}}(\mathcal{B}_{e})|_{Cs\times{\mathbb{C}}^{*}},\,\mathcal{L}(C,s):={\bf{K}}_{H\times{\mathbb{C}}^{*}}(\mathcal{B}_{e}^{{\mathbb{C}}^{*}})|_{Cs\times{\mathbb{C}}^{*}}.

The algebras {\mathbfcal{H}}, e{\mathbfcal{H}}_{e} act naturally on 𝒦(C,s)\mathcal{K}(C,s) and the algebra 𝐉e=𝐊Ze×(e×e){\bf{J}}_{e}={\bf{K}}_{Z_{e}\times{\mathbb{C}}^{*}}(\mathcal{B}_{e}^{{\mathbb{C}}^{*}}\times\mathcal{B}_{e}^{{\mathbb{C}}^{*}}) acts naturally on (C,s)\mathcal{L}(C,s). We also have the natural action of ΓM\Gamma_{M} on the families 𝒦(C,s)\mathcal{K}(C,s), 𝒦(L,s)\mathcal{K}(L,s). Taking ρ\rho-multiplicity spaces we obtain the families 𝒦(C,s,ρ):=𝒦(C,s)ρ\mathcal{K}(C,s,\rho):=\mathcal{K}(C,s)_{\rho}, (C,s,ρ):=(C,s)ρ\mathcal{L}(C,s,\rho):=\mathcal{L}(C,s)_{\rho}. One can show that for (χ,q)C×(\chi,q)\in C\times{\mathbb{C}}^{*} we have ΓM\Gamma_{M}-equivariant identifications:

𝒦(C,s)|(χ,q)=K(sχ,e,q),(C,s)|{χ}×=E(sχ,e)[𝐯,𝐯1],\mathcal{K}(C,s)|_{(\chi,q)}=K(s\chi,e,q),\,\mathcal{L}(C,s)|_{\{\chi\}\times{\mathbb{C}}^{*}}=E(s\chi,e)\otimes{\mathbb{C}}[{\bf{v}},{\bf{v}}^{-1}],

so these families are nothing but the families of (directs sums of) standard or irreducible modules over our algebras. We prove the following theorem that should be considered as a version ‘‘in families’’ of the identification ϕqE(s,e,ρ)K(s,e,ρ,q)\phi_{q}^{*}E(s,e,\rho)\simeq K(s,e,\rho,q) above333Existence of an action of 𝐉{\bf{J}} on 𝒦(C,s)\mathcal{K}(C,s) that is algebraic in ss is claimed and used in [19], see the proof of Theorem 2.4 (3)(3) in loc. cit., but not checked there in detail. Our Theorem D fills in the details.

Theorem D.

We have a natural ΓM\Gamma_{M}-equivariant isomorphism of the families of {\mathbfcal{H}} and 𝐜{\mathbfcal{H}}_{\bf{c}}-modules respectively:

ϕ𝐜(C,s)𝒦(C,s),ϕ𝐜(C,s)𝐜𝒦(C,s).\phi^{{\bf{c}}*}\mathcal{L}(C,s)\simeq_{{\mathbfcal{H}}}\mathcal{K}(C,s),\,\phi^{*}_{{\bf{c}}}\mathcal{L}(C,s)\simeq_{{\mathbfcal{H}}_{\bf{c}}}\mathcal{K}(C,s).

Let us now return to the proof of Theorem 1.3.4. We show that the families 𝒦(ZM,s,ρ,q)\mathcal{K}(Z_{M},s,\rho,q), 𝒱(ZM,s,ρ,q)\mathcal{V}(Z_{M},s,\rho,q) can be identified after restricting to some open dense subset 𝒰ZM\mathcal{U}\subset Z_{M} containing all non-strictly positive characters. Using Theorem D, it then follows that we are reduced to proving the following proposition.

Proposition 1.3.5.

Let \mathcal{E} be the subalgebra of M,sEnd𝒪(ZM)(ZM,s,q)=:E~\prod_{M,s}\operatorname{End}_{{\mathcal{O}}(Z_{M})}\mathcal{L}(Z_{M},s,q)=:\tilde{E} consisting of elements ϕ=(ϕ(M,s))M,s\phi=(\phi(M,s))_{M,s} satisfying the following property:

for any conjugate pair sχtχs\chi\sim t\chi^{\prime} (here χZL\chi^{\prime}\in Z_{L} for a Levi subgroup LL^{\vee}, so that tLt\in L^{\vee}, and eLieLe\in\operatorname{Lie}L^{\vee}), the following equality holds:

ϕ(M,s)χ=ϕ(L,t)χ.\phi(M,s)_{\chi}=\phi(L,t)_{\chi^{\prime}}.

Then Je\mathcal{E}\simeq J_{e} via the action map α:Je\alpha\colon J_{e}\to\mathcal{E}.

To prove this proposition, we show that the homomorphism α:𝐉e\alpha\colon{\bf{J}}_{e}\rightarrow\mathcal{E} becomes an isomorphism after completion at every point of Spec𝐊Ze(pt)\operatorname{Spec}{\bf{K}}_{Z_{e}}(\operatorname{pt}) by describing completions of 𝐉e{\bf{J}}_{e}, \mathcal{E} as explicit subalgebras of the completion of E~\tilde{E}.444Our argument is largely parallel to the original proof of [19], the difference is that we consider completion at the points of Spec𝐊Ze(pt)\operatorname{Spec}{\bf{K}}_{Z_{e}}(\operatorname{pt}) which is an exact functor, so, in particular, the natural embedding E~\mathcal{E}\subset\tilde{E} remains embedding after completions. In general it does not induce an embedding of fibers (since the fibers of 𝐉e{\bf{J}}_{e} are not semisimple in general, see Section 4.4.2), so the last paragraph of the proof of [19, Theorem 1.8] does not hold as stated. We use our geometric description of JeJ_{e} and localization theorem in KK-theory to analyze the completion of 𝐉e{\bf{J}}_{e}.

1.3.4. Description of the cocenter of {\mathbfcal{H}}

In the last part of the text we use our geometric approach to JeJ_{e} to obtain a description of the cocenter of 𝐉e{\bf{J}}_{e} conjectured in [12, Section 6.2]. Let us formulate the theorem.

Let CommZe\operatorname{Comm}_{Z_{e}} be the commuting variety for ZeZ_{e} (with the reduced scheme structure). Set 𝒞Ze:=CommZe//Ze\mathcal{C}_{Z_{e}}:=\operatorname{Comm}_{Z_{e}}/\!/Z_{e}.

Theorem E.

Let 𝒪a(e)\mathcal{O}^{a}(e) be the space consisting of regular functions ff on 𝒞Ze\mathcal{C}_{Z_{e}}, subject to the following properties:

a) for any semisimple sZes\in Z_{e}, f|{s}×ZZe(s)f|_{\{s\}\times Z_{Z_{e}}(s)} is locally constant (and, hence, gives a well-defined function fsf_{s} on the component group Γes\Gamma_{e}^{s} of ZZe(s)Z_{Z_{e}}(s);

b) fsf_{s} is a sum of characters of the group Γes\Gamma_{e}^{s} arising in 𝐊(es){\bf{K}}(\mathcal{B}_{e}^{s}).

Then 𝒪a(e)𝐉e/[𝐉e,𝐉e]\mathcal{O}^{a}(e)\simeq{\bf{J}}_{e}/[{\bf{J}}_{e},{\bf{J}}_{e}].

Remark 1.3.6.

Recall that by [12, Theorem 1] the homomorphism ϕq:q𝐉=e𝐉e\phi_{q}\colon{\mathbfcal{H}}_{q}\rightarrow{\bf{J}}=\bigoplus_{e}{\bf{J}}_{e} induces an isomorphism on the level of cocenters for qq not a root of unity. This means that Theorem E gives a description of the cocenter q/[q,q]{\mathbfcal{H}}_{q}/[{\mathbfcal{H}}_{q},{\mathbfcal{H}}_{q}] for qq not a root of unity.

1.4. Structure of the paper

The paper is organized as follows. In Section 2 we recall definitions and known properties of the algebras \mathcal{H}, JJ. Section 3 is devoted to the proof of Theorem B. In Section 4 we prove Theorem A. In Section 5 we study representations of 𝐉{\bf{J}}, construct families of (irreducible) modules over it and prove Theorems C, D. In Section 6, we prove [19, Theorem 1.8 (3)] by our methods. In Section 7 we prove Theorem E i.e. describe the cocenter of 𝐉e{\bf{J}}_{e}. Appendix A contains proofs of various general facts about equivariant KK-theory that we need for our arguments.

1.5. Acknowledgements

We gratefully acknowledge helpful input from Dan Ciubotaru, Stefan Dawydiak, Pavel Etingof, Michael Finkelberg, Victor Ginzburg, Mikhail Goltsblat, William Graham, Do Kien Hoang, David Kazhdan, Ivan Losev, Jakub Löwit, Victor Ostrik, Oron Propp, Vadim Vologodsky and Zhiwei Yun. We are grateful to George Lusztig for useful comments on the first version of the text.

I. K. especially thanks Dmitrii Zakharov for explaining the material of Proposition 6.6.1 and Michael Finkelberg for his -nical generosity.

R.B. was partly supported by the NSF grant DMS-2101507.

2. Generalities on \mathcal{H} and JJ

2.1. The affine Hecke algebra

Let WW be as above and let SWS\subset W be the set of simple reflections. Let :W0\ell\colon W\rightarrow\mathbb{Z}_{\geqslant 0} be the length function on WW. The algebra =G\mathcal{H}=\mathcal{H}_{G} is an algebra over [𝐯,𝐯1]\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}] with basis {Tw}wW\{T_{w}\}_{w\in W}. Multiplication in this algebra is determined by the relations TxTy=TxyT_{x}T_{y}=T_{xy} when (xy)=(x)+(y)\ell(xy)=\ell(x)+\ell(y) and (Tsv2)(Ts+1)=0(T_{s}-v^{2})(T_{s}+1)=0 for sSs\in S. Recall that :={\mathbfcal{H}}:=\mathcal{H}\otimes_{\mathbb{Z}}{\mathbb{C}} is the complexification of \mathcal{H}.

2.1.1. Lusztig’s cells

There is a partition of WW into the union of the so-called cells. We start with a brief recollection of it.

Let \leqslant be the strong Bruhat order. First of all, two elements xx and yy of WW are said to be connected, if either x<yx<y or y<xy<x; and degPx,y=|(x)(y)|12\operatorname{deg}P_{x,y}=\frac{|\ell(x)-\ell(y)|-1}{2}: in particular, (x)(y)\ell(x)-\ell(y) is odd. (Here, we denote by Px,yP_{x,y} the corresponding Kazhdan-Lusztig polynomial, see [35] and [39, Section 1.2].)

To each wWw\in W corresponds the so-called left descending set Dl(w)={sS|sww}D^{l}(w)=\{s\in S\ |\ sw\leqslant w\}.

Now, for x,yWx,y\in W we say that xLyx\leqslant_{L}y, if there is a chain of elements (x1=x,,xl,,xk=y)(x_{1}=x,\ldots,x_{l},\ldots,x_{k}=y), s.t. all neighbours in it are connected, and DMissing Operator(xi)Dl(xi+1)D^{l}(x_{i})\setminus D^{l}(x_{i+1})\neq\emptyset.

It is well-known that L\leqslant_{L} is a partial preorder, and the corresponding equivalence classes are called left cells: this is the work by Lusztig ([37]).

Right cells are subsets of the form Ψ1\Psi^{-1} for Ψ\Psi a left cell. Their definition can also be given in a similar fashion to the one above. We should only replace Dl(w)D^{l}(w) by Dr(w):={sS|wsw}D^{r}(w):=\{s\in S\ |\ ws\leqslant w\}.

Finally, there are also two-sided cells (we will denote the set of them by 𝒞\mathcal{C}). They are defined in almost the same fashion as right or left ones; now the appropriate condition for neighbors in (x1,,xl,)(x_{1},\ldots,x_{l},\ldots) is that either Dl(xi)Dl(xi+1)D^{l}(x_{i})\setminus D^{l}(x_{i+1})\neq\emptyset or Dr(xi)Dr(xi+1)D^{r}(x_{i})\setminus D^{r}(x_{i+1})\neq\emptyset for each ii.

2.1.2. Two-sided cells and nilpotent conjugacy classes

The famous theorem by Lusztig (cf. [40]) says that two-sided cells in WW are in bijection with the conjugacy classes of nilpotent elements in the Lie algebra 𝔤\mathfrak{g}^{\vee} of the Langlands dual group GG^{\vee}. In particular, this gives an order on cells (via the closure order on the nilpotent orbits). This order coincides with the natural one (this was conjectured by Lusztig and proved by the first author in [9, Theorem 4 (b)]).

The so-called aa-function sends every two-sided cell 𝐜\bf c to the dimension of the corresponding Springer fiber e\mathcal{B}_{e}.

It has a combinatorial meaning as well. Namely, recall that \mathcal{H} is the affine Hecke algebra of GG and let us denote by Cx,xW,C_{x},\ x\in W, the Kazhdan-Lusztig basis elements in \mathcal{H}. Then, if hx,y,z[𝐯,𝐯1]h_{x,y,z}\in\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}] are structure constants of \mathcal{H} with respect to {Cw}\{C_{w}\}, a(𝐜)-a(\bf c) is the lowest possible degree of non-zero term in hx,y,zh_{x,y,z}, x,y𝐜x,y\in\bf c, zWz\in W.

2.2. Asymptotic affine Hecke algebra

Now we recall Lusztig’s definition of the asymptotic affine Hecke algebra JJ (cf. loc. cit.). It has a basis {tw|wW}\{t_{w}\,|\,w\in W\}. In this basis the structure constant γx,y,z\gamma_{x,y,z} is the constant term of the polynomial 𝐯a(z)hx,y,z1{\bf{v}}^{a(z)}h_{x,y,z^{-1}}: txty=zγx,y,ztz1t_{x}t_{y}=\sum_{z}\gamma_{x,y,z}t_{z^{-1}}.

It is shown in loc. cit. that JJ is actually an associative algebra. Moreover, denoting the span over [𝐯,𝐯1]{\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}]} of {tx|x𝐜}\{t_{x}\ |\ x\in\bf c\} by J𝐜J_{\bf c}, one gets a decomposition of JJ as a direct sum (product) of rings:

J=𝐜𝒞J𝐜.J=\bigoplus_{\bf c\in\mathcal{C}}J_{\bf c}.

We will sometimes denote J𝐜J_{\bf c} by JeJ_{e} for 𝐜\bf c corresponding to the nilpotent orbit 𝕆e=Ge{\mathbb{O}}_{e}=G^{\vee}e (cf. Section 2.1.2). The same notation will be used for subscripts and superscripts of various morphisms in Section 2.3.

2.2.1. Distinguished involutions and the unit element in J𝐜J_{\bf{c}}

The set of distinguished involutions: 𝒟W\mathcal{D}\subset W consists of all elements dWd\in W, such that a(w)(w)=2degP1,w.a(w)-\ell(w)=2\operatorname{deg}P_{1,w}. It is known that its elements are actually involutions, and each one-sided cell contains exactly one of them.

The element 𝟏=d𝒟td{\bf 1}=\sum_{d\in\mathcal{D}}t_{d} is the unit element in the algebra JJ, while for 𝐜𝒞{\bf{c}}\in\mathcal{C}, the element 𝟏𝐜:=d𝐜𝒟td{\bf 1}_{\bf{c}}:=\sum_{d\in{\bf{c}}\cap\mathcal{D}}{{t_{d}}} is the unit element in J𝐜J_{\bf{c}}.

2.3. Relation between \mathcal{H} and JJ

Recall Lusztig’s homomorphism (see [39]) ϕ:J[𝐯,𝐯1]\phi\colon\mathcal{H}\to J\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}],

Cwd𝒟,zW,a(z)=a(d)hw,d,ztzC_{w}\mapsto\sum\limits_{d\in\mathcal{D},z\in W,\ a(z)=a(d)}h_{w,d,z}t_{z}

and the additive isomorphism ψ:J[𝐯,𝐯1],Cwtw\psi\colon\mathcal{H}\,\vphantom{j^{X^{2}}}\smash{\overset{\sim}{\vphantom{\rule{0.0pt}{1.99997pt}}\smash{\longrightarrow}}}\,J\otimes\mathbb{\mathbb{Z}}[{\bf{v}},{\bf{v}}^{-1}],\ C_{w}\mapsto t_{w}.

For a two-sided cell 𝐜𝒞\bf c\in\mathcal{C} we set

J𝐜:=𝐜LR𝐜J𝐜.J_{\leqslant\bf c}:=\bigoplus\limits_{{\bf c^{\prime}}\leqslant_{LR}\bf c}J_{\bf c^{\prime}}.

Obviously ψ(J𝐜[𝐯,𝐯1])=𝐜\psi(J_{\leqslant\bf c}[{\bf{v}},{\bf{v}}^{-1}])=\mathcal{H}_{\leqslant\bf c} is the corresponding term of the cell filtration.

The coincidence of the geometric and the cell filtration implies that 𝐜\mathcal{H}_{\leqslant\bf c} can be identified with the G×G^{\vee}\times{\mathbb{C}}^{*}-equivariant KK-theory of the preimage of 𝕆¯e\overline{\mathbb{O}}_{e} in 𝒩~×𝒩𝒩~\widetilde{{\mathcal{N}}}\times_{{\mathcal{N}}}\widetilde{{\mathcal{N}}} and 𝐜=KZe×(e×e)\mathcal{H}_{\bf c}=K_{Z_{e}\times{\mathbb{C}}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e}), see [56], [62] for details.

Finally, note that homomorphism ϕ\phi induces homomorphisms

ϕ𝐜:𝐜J𝐜[𝐯,𝐯1],ϕ𝐜:J𝐜[𝐯,𝐯1]\phi_{\bf c}\colon\mathcal{H}_{\bf c}\to J_{\bf c}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}],\,\phi^{\bf c}\colon\mathcal{H}\to J_{\bf c}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}]

(ϕ𝐜\phi^{\bf c} is the composition of ϕ\phi and the projection to the direct summand).

The isomorphism ψ\psi restricts to the isomorphisms ψ𝐜:𝐜J𝐜[𝐯,𝐯1]\psi_{\bf c}\colon\mathcal{H}_{\bf c}\,\vphantom{j^{X^{2}}}\smash{\overset{\sim}{\vphantom{\rule{0.0pt}{1.99997pt}}\smash{\longrightarrow}}}\,J_{\bf c}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}].

2.3.1. Bimodule structure on J[𝐯,𝐯1]J\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}]

A homomorphism of rings RSR\to S equips SS with the structure of an RSR-S bimodule. In particular, the homomorphism ϕ𝐜\phi^{\bf{c}} defines a J𝐜[𝐯,𝐯1]\mathcal{H}-J_{\bf{c}}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}] - bimodule structure on J𝐜[𝐯,𝐯1]J_{\bf{c}}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}]. Similarly, ϕ𝐜\phi_{\bf{c}} defines a 𝐜J𝐜[𝐯,𝐯1]\mathcal{H}_{\bf{c}}-J_{\bf{c}}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}] - bimodule structure on J𝐜[𝐯,𝐯1]J_{\bf{c}}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}].

In this section, we will give an alternative description of the (left) action of \mathcal{H} on J𝐜[𝐯,𝐯1]J_{\bf{c}}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}] as above. We will then conclude that this \mathcal{H}-module is isomorphic to 𝐜\mathcal{H}_{\bf{c}}. We will obtain similar statements for the 𝐜\mathcal{H}_{\bf{c}}-module J𝐜[𝐯,𝐯1]J_{\bf{c}}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}].

Let ht𝐜\operatorname{ht}_{\bf c} denote the projection 𝐜𝐜{\mathcal{H}}_{\leqslant\bf c}\twoheadrightarrow{\mathcal{H}}_{\bf c}. Following [41, Section 3] let us define the action of \mathcal{H} on 𝐜𝐜J𝐜[𝐯,𝐯1]\bigoplus_{{\bf{c}}^{\prime}\leqslant{\bf{c}}}J_{{\bf{c}}}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}] by the formula:

(2.3.1) Cxty=ψ(CxCy).C_{x}\cdot t_{y}=\psi(C_{x}C_{y}).

Note that directly from the definitions ψ:𝐜𝐜𝐜J𝐜[𝐯,𝐯1]\psi\colon\mathcal{H}_{\leqslant{\bf{c}}}\,\vphantom{j^{X^{2}}}\smash{\overset{\sim}{\vphantom{\rule{0.0pt}{1.99997pt}}\smash{\longrightarrow}}}\,\bigoplus_{{\bf{c}}^{\prime}\leqslant{\bf{c}}}J_{{\bf{c}}}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}] is the isomorphism of \mathcal{H}-modules (in particular, it follows that the formula (2.3.1) indeed defines the action of \mathcal{H}).

Note now that ψ(<𝐜)=𝐜<𝐜J𝐜[𝐯,𝐯1]\psi(\mathcal{H}_{<{\bf{c}}})=\bigoplus_{{\bf{c}}^{\prime}<{\bf{c}}}J_{{\bf{c}}}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}] is an \mathcal{H}-submodule of 𝐜𝐜J𝐜[𝐯,𝐯1]\bigoplus_{{\bf{c}}^{\prime}\leqslant{\bf{c}}}J_{{\bf{c}}}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}]. Modding out by this submodule, we obtain the action of \mathcal{H} on J𝐜[𝐯,𝐯1]J_{{\bf{c}}}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}] such that for y𝐜y\in\bf c we have:

(2.3.2) Cxty=ψ𝐜(ht𝐜(CxCy)).C_{x}\cdot t_{y}=\psi_{{\bf{c}}}(\operatorname{ht_{\bf c}}(C_{x}C_{y})).

It induces the action 𝐜J𝐜[𝐯,𝐯1]\mathcal{H}_{\bf{c}}\curvearrowright J_{\bf{c}}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}].

Lemma 2.3.1.

(a) Two \mathcal{H}-actions on J𝐜[𝐯,𝐯1]J_{\bf{c}}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}] defined above (the one coming from ϕ𝐜\phi^{\bf{c}} and the one given by the formula (2.3.2)) coincide.

(b) Two 𝐜\mathcal{H}_{\bf{c}}-actions on J𝐜[𝐯,𝐯1]J_{\bf{c}}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}] defined above (the one coming from ϕ𝐜\phi_{\bf{c}} and the one given by the formula (2.3.2)) coincide.

Proof.

It is enough to prove part (a), part (b) will follow.

Let us prove part (a). We need to check that

Cxty=ϕ𝐜(Cx)tyC_{x}\cdot t_{y}=\phi^{{\bf{c}}}(C_{x})t_{y}

for xWx\in W and y𝐜y\in{\bf{c}}. So, our goal is to check that

(2.3.3) ψ𝐜(ht𝐜(CxCy))=ϕ𝐜(Cx)ty.\psi_{\bf{c}}(\operatorname{ht}_{\bf{c}}(C_{x}C_{y}))=\phi^{\bf{c}}(C_{x})t_{y}.

Note that directly from the definitions we have:

(2.3.4) ϕ𝐜(Cw)=d𝒟𝐜ψ𝐜(ht𝐜(CwCd)).\phi^{\bf{c}}(C_{w})=\sum_{d\in\mathcal{D}\cap{\bf{c}}}\psi_{\bf{c}}(\operatorname{ht}_{\bf{c}}(C_{w}C_{d})).

The following formula follows from [38, 2.4 (d)] (see also [12, Proof of Proposition 2]):

(2.3.5) ψ𝐜(ht𝐜(Cx1Cx2))tx3=ψ𝐜(Cx1ψ𝐜1(tx2tx3)).\psi_{\bf c}(\operatorname{ht}_{\bf{c}}(C_{x_{1}}C_{x_{2}}))t_{x_{3}}=\psi_{\bf c}(C_{x_{1}}\psi_{\bf c}^{-1}(t_{x_{2}}t_{x_{3}})).

Now, setting x1=xx_{1}=x, x3=yx_{3}=y, summing (2.3.5) over x2𝒟𝐜x_{2}\in\mathcal{D}\cap\bf c and using (2.3.4) together with the fact that 𝟏𝐜=x2𝒟𝐜tx2{\bf{1}}_{{\bf{c}}}=\sum_{x_{2}\in\mathcal{D}\cap{\bf{c}}}t_{x_{2}} is the identity element of J𝐜J_{\bf{c}} we obtain the desired equation (2.3.3). ∎

Corollary 2.3.2.

(a) The \mathcal{H}-module J𝐜[𝐯,𝐯1]J_{\bf{c}}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}] defined via ϕ𝐜\phi^{\bf{c}} is isomorphic to 𝐜\mathcal{H}_{\bf{c}}.

(b) The 𝐜\mathcal{H}_{\bf{c}}-module J𝐜[𝐯,𝐯1]J_{\bf{c}}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}] defined via ϕ𝐜\phi_{\bf{c}} is isomorphic to 𝐜\mathcal{H}_{\bf{c}}.

Proof.

It follows from Lemma 2.3.1 that ψ\psi is the desired isomorphism of \mathcal{H} (resp. 𝐜\mathcal{H}_{\bf{c}}) - modules. ∎

2.4. Description of JeJ_{e} via finite centrally extended sets

A key ingredient in our proof of the Theorem B presented in the next section is the finite-set realization of J𝐜J_{\bf{c}}, which we now recall.

The following was conjectured (in a slightly different form) by Lusztig [40]. It was proven in [60] in type A and in [15] in general.

Definition 2.4.1.

A centrally extended finite HH-set (for a reductive group HH) is a finite HH-set YY equipped with some central extension 1𝔾mStabH(y)~StabH(y)11\to\mathbb{G}_{m}\to\widetilde{\operatorname{Stab}_{H}(y)}\to\operatorname{Stab}_{H}(y)\to 1 for every yYy\in Y such that for every gGg\in G the conjugation isomorphism Cg:StabH(y)StabH(gy)C_{g}\colon\operatorname{Stab}_{H}(y)\to\operatorname{Stab}_{H}(gy) extends to an isomorphism between these central extensions satisfying natural compatibilities.

For a centrally extended finite HH-set YY, one defines an equivariant sheaf on YY as a usual HH-equivariant sheaf \mathcal{F} with an additional data of StabH~(y)\widetilde{\operatorname{Stab}_{H}}(y)-action on y\mathcal{F}_{y}, which induces an action of 𝔾m\mathbb{G}_{m}-factor by identity character. This allows us to consider KH(pt)K_{H}(\operatorname{pt})-algebra KH(Y)K_{H}(Y).

Theorem 2.4.2 ([15]).

Let 𝐜W{\bf{c}}\subset W be a two-sided cell in WW and let e𝒩e\in\mathcal{N} denote the corresponding nilpotent element. Then there exists a centrally extended finite ZeZ_{e}-set YY such that J𝐜KZe(Y×Y)J_{\bf c}\simeq K_{Z_{e}}(Y\times Y).

In particular, one has a central embedding KZe(pt)J𝐜K_{Z_{e}}(\operatorname{pt})\to J_{\bf c}, which will be of crucial importance for us.

Remark 2.4.3.

For G=PGLnG=PGL_{n} the situation simplifies. First, the centralizers in question are connected so the ZeZ_{e}-action is trivial.

Also, it turns out that no non-trivial central extensions appear. Thus, we arrive at the following statement proved in [60]: J𝐜J_{\bf c} can be realized as a matrix algebra over the ring KZe(pt)=K0(Zemod)K_{Z_{e}}(\operatorname{pt})=K_{0}(Z_{e}-\operatorname{mod}). Note that JJ corresponding to PGLn\operatorname{PGL}_{n} in our notations is JJ corresponding to the extended affine Weyl group associated with SLn\operatorname{SL}_{n} in the notations of [60].

2.5. Description of e\mathcal{H}_{e} and JeJ_{e} via bimodules over non-commutative Springer

We now recall, following [13], the relation between the above algebras and the non-commutative Springer resolution AA introduced in [7], [14].

Recall that AA is an algebra equipped with a derived equivalence

Db(AmodG×)Db(CohG×(𝒩~)),D^{b}(A-\operatorname{mod}_{G^{\vee}\times\mathbb{C}^{*}})\simeq D^{b}(\operatorname{Coh}_{G\times\mathbb{C}^{*}}(\widetilde{\mathcal{N}})),

where the subscript denotes the equivariant structure. It has a canonical filtration indexed by the poset of nilpotent conjugacy classes in 𝒩\mathcal{N} equipped with the closure order.

We will denote the associated graded piece Ae/A<eA_{\leqslant e}/A_{<e} by AeA_{e}. It is known (cf. [14]) that K0(AebimodZe×)K_{0}(A_{e}-\operatorname{bimod}_{Z_{e}\times\mathbb{C}^{*}}) is isomorphic to e\mathcal{H}_{e}. Now, it follows from [13, Section 8] that there exists an isomorphism of KZe×(pt)=ReK_{Z_{e}\times\mathbb{C}^{*}}(\operatorname{pt})=R_{e}-algebras:

(2.5.1) Je[𝐯,𝐯1]K0(AebimodssZe×),J_{e}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}]\simeq K_{0}(A_{e}-\operatorname{bimod}^{\operatorname{ss}}_{Z_{e}\times\mathbb{C}^{*}}),

where in the right hand side we take the KK-group of the category of semisimple bimodules. Here the morphism ψ\psi can be identified with the tautological isomorphism between the KK-theories of the categories of semisimple and all finite length modules. The ReR_{e}-action on Je[𝐯,𝐯1]J_{e}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}] comes from the identification (2.5.1).

For future use we record the following.

Corollary 2.5.1.

The morphisms ϕe\phi_{e} and ψe\psi_{e} are ReR_{e}-linear.

Proof.

For ψe\psi_{e} this is immediate from the above. To deduce the claim for ϕe\phi_{e}, recall that for any xWx\in W, ϕ(Cx)=d𝐜𝒟ψ(CxCd)\phi(C_{x})=\sum_{d\in{\bf{c}}\cap\mathcal{D}}\psi(C_{x}C_{d}), the sum running over distinguished involutions inside the cell associated to ee. Now the claim follows, since the ReR_{e}-action commutes with both ψ\psi and the multiplication in e\mathcal{H}_{e}. ∎

Remark 2.5.2.

See [39, Proposition 6] for a related statement.

Remark 2.5.3.

Note that e\mathcal{H}_{e} is not a unital algebra, in particular, the ReR_{e}-action on e\mathcal{H}_{e} does not come from a homomorphism ReeR_{e}\to\mathcal{H}_{e}.

3. Proof of Theorem B

3.1.

Now we are ready to prove Theorem B. Recall that it claims that we have isomorphisms of algebras:

J𝐜[𝐯,𝐯1]End(J𝐜[𝐯,𝐯1])oppEnd(𝐜)oppEndRe𝐜(𝐜)opp.J_{{\bf{c}}}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}]\,\vphantom{j^{X^{2}}}\smash{\overset{\sim}{\vphantom{\rule{0.0pt}{1.99997pt}}\smash{\longrightarrow}}}\,\operatorname{End}_{\mathcal{H}}(J_{\mathbf{c}}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}])^{\mathrm{opp}}\simeq\operatorname{End}_{\mathcal{H}}(\mathcal{H}_{\bf{c}})^{\mathrm{opp}}\simeq\operatorname{End}^{R_{e}}_{\mathcal{H}_{\bf{c}}}(\mathcal{H}_{\bf{c}})^{\mathrm{opp}}.

We have already observed in Corollary 2.3.2 that J𝐜[𝐯,𝐯1]𝐜J_{{\bf{c}}}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}]\simeq\mathcal{H}_{\bf{c}} as left \mathcal{H} (and 𝐜\mathcal{H}_{\bf{c}})-modules so it remains to check that the natural homomorphisms

(3.1.1) J𝐜[𝐯,𝐯1]End(J𝐜[𝐯,𝐯1])opp,J𝐜[𝐯,𝐯1]End𝐜Re(J𝐜[𝐯,𝐯1])oppJ_{{\bf{c}}}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}]\rightarrow\operatorname{End}_{\mathcal{H}}(J_{\mathbf{c}}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}])^{\mathrm{opp}},\,J_{\bf{c}}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}]\rightarrow\operatorname{End}_{\mathcal{H}_{\bf{c}}}^{R_{e}}(J_{\bf{c}}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}])^{\mathrm{opp}}

are actually isomorphisms. Note that both of them are injective since J𝐜J_{\bf{c}} is unital. So, it remains to check that the homomorpisms in (3.1.1) are surjective. We start from the first one, the proof for the second one is completely analogous.

First of all, we claim that it is enough to check that the homomorphism

(3.1.2) 𝐉𝐜[𝐯,𝐯1]End(𝐉𝐜[𝐯,𝐯1])opp{\bf{J}}_{{\bf{c}}}\otimes\mathbb{C}[{\bf{v}},{\bf{v}}^{-1}]\rightarrow\operatorname{End}_{{\mathbfcal{H}}}({\bf{J}}_{\mathbf{c}}\otimes\mathbb{C}[{\bf{v}},{\bf{v}}^{-1}])^{\mathrm{opp}}

is surjective. Indeed, since J𝐜J_{\bf{c}} is a free \mathbb{Z}-module, we have embeddings

J𝐜𝐉𝐜,End(J𝐜[𝐯,𝐯1])oppEnd(𝐉𝐜[𝐯,𝐯1])opp.J_{\bf{c}}\hookrightarrow{\bf{J}}_{\bf{c}},~{}\operatorname{End}_{\mathcal{H}}(J_{\bf{c}}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}])^{\mathrm{opp}}\hookrightarrow\operatorname{End}_{{\mathbfcal{H}}}({\bf{J}}_{\bf{c}}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}])^{\mathrm{opp}}.

Now the homomorphism (3.1.2) is given by 𝐉𝐜[𝐯,𝐯1]aa{\bf{J}}_{\bf{c}}\otimes{\mathbb{C}}[{\bf{v}},{\bf{v}}^{-1}]\ni a\mapsto-\cdot a, so if the operator a-\cdot a lies in End(J𝐜[𝐯,𝐯1])oppEnd(𝐉𝐜[𝐯,𝐯1])opp\operatorname{End}_{\mathcal{H}}(J_{\bf{c}}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}])^{\mathrm{opp}}\hookrightarrow\operatorname{End}_{{\mathbfcal{H}}}({\bf{J}}_{\bf{c}}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}])^{\mathrm{opp}} then a=1aJ𝐜[𝐯,𝐯1]a=1\cdot a\in J_{\bf{c}}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}].

So, we need to prove the surjectivity of (3.1.2). The proof consists of several steps.

A) Recall that 𝐉𝐜{\bf{J}}_{\bf{c}} is an algebra with a unit element (see Section 2.2.1). To prove the proposition it is sufficient to show vanishing of an {\mathbfcal{H}}-endomorphism of 𝐉𝐜[𝐯,𝐯1]{\bf{J}}_{\bf{c}}\otimes\mathbb{C}[{\bf{v}},{\bf{v}}^{-1}] which sends 𝟏𝐜{\bf 1}_{\bf{c}} to 0. This, in turn, can be reformulated as follows:

(3.1.3) V𝐜:=Hom(𝐉𝐜[𝐯,𝐯1]/Im(ϕ𝐜),𝐉𝐜[𝐯,𝐯1])=0,V_{\bf c}:=\operatorname{Hom}_{{\mathbfcal{H}}}({\bf{J}}_{\bf c}\otimes\mathbb{C}[{\bf v},{\bf v}^{-1}]/\operatorname{Im}(\phi_{\bf c}),{\bf{J}}_{\bf c}\otimes\mathbb{C}[{\bf v},{\bf v}^{-1}])=0,

where abusing notations we denote the complexification of ϕ𝐜\phi_{\bf{c}} by the same letter.

B) To fix ideas, we first prove the equality (3.1.3) in the special case when G=GLnG=\operatorname{GL}_{n}. In this case centralizers of all nilpotent elements are connected, so 𝐑e=𝐊Ze×(pt){\bf{R}}_{e}={\bf{K}}_{Z_{e}\times{\mathbb{C}}^{*}}(\operatorname{pt}) has no zero-divisors. Let us denote by 𝐑e{\bf{R}}_{e}^{{}^{\prime}} the image of the restriction homomorphism 𝐑:=𝐊G×(pt)𝐊Ze×(pt)=𝐑e{\bf{R}}:={\bf{K}}_{G^{\vee}\times\mathbb{C}^{*}}(\operatorname{pt})\to{\bf{K}}_{Z_{e}\times{\mathbb{C}}^{*}}(\operatorname{pt})={\bf{R}}_{e}.

Now, since ϕ𝐜\phi_{\bf c} is injective and 𝐑e{\bf{R}}_{e}-linear (cf. [12], and Corollary 2.5.1), ϕ𝐜(𝐜)\phi_{\bf c}({\mathbfcal{H}}_{\bf c}) has the same 𝐑e{\bf{R}}_{e}- (and, hence, 𝐑e{\bf{R}}_{e}^{{}^{\prime}}-) rank as 𝐜{\mathbfcal{H}}_{\bf{c}}, thus it is an 𝐑e{\bf{R}}_{e}^{{}^{\prime}}-submodule of full rank in 𝐉𝐜{\bf{J}}_{\bf{c}}. (These ranks are finite by [39, Proposition 6].)

Moreover, the bijective map ψe\psi_{e} is 𝐑e{\bf{R}}_{e}-linear as well (see Corollary 2.5.1), so rk𝐑eIm(ϕ𝐜)=rk𝐑e𝐉𝐜[𝐯,𝐯1]\operatorname{rk}_{{\bf{R}}_{e}^{{}^{\prime}}}\operatorname{Im}(\phi_{\bf c})=\operatorname{rk}_{{\bf{R}}_{e}^{{}^{\prime}}}{\bf{J}}_{\bf c}\otimes\mathbb{C}[{\bf{v}},{\bf{v}}^{-1}].

It means that 𝐉𝐜[𝐯,𝐯1]/Im(ϕ𝐜){\bf{J}}_{\bf c}\otimes\mathbb{C}[{\bf{v}},{\bf{v}}^{-1}]/\operatorname{Im}(\phi_{\bf c}) is 𝐑e{\bf{R}}_{e}^{{}^{\prime}}-torsion.

Since 𝐑e{\bf{R}}_{e}^{{}^{\prime}} has no zero divisors (centralizers being connected) and 𝐉𝐜[𝐯,𝐯1]{\bf{J}}_{\bf c}\otimes\mathbb{C}[{\bf{v}},{\bf{v}}^{-1}] is a torsion-free module over 𝐑e{\bf{R}}_{e} (cf. Lemma 3.1.1 below) the statement follows once we check that an {\mathbfcal{H}}-linear morphism is 𝐑e{\bf{R}}_{e}^{{}^{\prime}}-linear. The latter point follows from the well-known isomorphism between 𝐑{\bf{R}} and the center of {\mathbfcal{H}} [5] and compatibility between the 𝐑{\bf{R}}- and 𝐑e{\bf{R}}_{e}-actions on e{\mathbfcal{H}}_{e} (cf. [60]).

C) We now consider the general case, then Ze=:ZZ_{e}=:Z may have several connected components which we denote by ZjZ_{j}, jΓ:=π0(Z)j\in\Gamma:=\pi_{0}(Z). Each conjugacy class Ωi\Omega_{i} in Γ\Gamma gives rise to an idempotent aia_{i} in 𝐑e{\bf{R}}_{e}, so that 𝐑e=iai𝐑e{\bf{R}}_{e}=\bigoplus_{i}a_{i}{\bf{R}}_{e}, and ai𝐑ea_{i}{\bf{R}}_{e} (and, hence, ai𝐑ea_{i}{\bf{R}}_{e}^{{}^{\prime}}) has no zero-divisors. The desired statement follows as above from the following two lemmas.

Lemma 3.1.1.

ai𝐉𝐜[𝐯,𝐯1]a_{i}{\bf{J}}_{\bf c}\otimes\mathbb{C}[{\bf{v}},{\bf{v}}^{-1}] is a torsion-free module over ai𝐑ea_{i}{\bf{R}}_{e}.

Lemma 3.1.2.

Any 𝐜{\mathbfcal{H}}_{\bf{c}}-linear (in particular, any {\mathbfcal{H}}-linear) endomorphism of 𝐉𝐜[𝐯,𝐯1]{\bf{J}}_{\bf c}\otimes\mathbb{C}[{\bf{v}},{\bf{v}}^{-1}] preserves the decomposition 𝐉𝐜[𝐯,𝐯1]=iai𝐉𝐜[𝐯,𝐯1].{\bf{J}}_{\bf c}\otimes\mathbb{C}[{\bf{v}},{\bf{v}}^{-1}]=\bigoplus_{i}a_{i}{\bf{J}}_{\bf c}\otimes\mathbb{C}[{\bf{v}},{\bf{v}}^{-1}].

D) Proof of Lemma 3.1.1. In Section 2.4 we recalled the realization of J𝐜J_{\bf c} as ZeZ_{e}-equivariant KK-theory of a centrally extended finite set. We claim that the statement of Lemma 3.1.1 is true for any 𝐑e{\bf{R}}_{e}-module arising that way.

It suffices to consider the case when the action of ZeZ_{e} on the finite set is transitive. If the corresponding central extension is trivial then the complexified equivariant KK-theory is identified with the ring of conjugation invariant regular functions on a finite index subgroup HZe×H\subset Z_{e}\times\mathbb{C}^{*}; in general it is a subspace in the ring of invariant functions on a finite covering H~\tilde{H} of such a subgroup HH. The statement is clear since a component in H~\tilde{H} maps surjectively to a component in Ze×Z_{e}\times\mathbb{C}^{*}.

E) Proof of Lemma 3.1.2. As follows from Corollary 2.5.1, we have decomposition

𝐜=iai𝐜.{\mathbfcal{H}}_{\bf c}=\bigoplus_{i}a_{i}{\mathbfcal{H}}_{\bf c}.

Now consider the 𝐜{\mathbfcal{H}}_{\bf{c}}-action on 𝐉𝐜[𝐯,𝐯1]{\bf{J}}_{\bf c}\otimes\mathbb{C}[{\bf{v}},{\bf{v}}^{-1}] via ϕ𝐜\phi_{\bf{c}} (cf. Section 2.3.1), since aia_{i}’s are orthogonal we get:

ai𝐜(aj𝐉𝐜[𝐯,𝐯1])=0a_{i}\mathcal{{\mathbfcal{H}}}_{\bf c}\cdot(a_{j}{\bf{J}}_{\bf c}\otimes\mathbb{C}[{\bf{v}},{\bf{v}}^{-1}])=0

for iji\neq j. Thus ai𝐉𝐜a_{i}{\bf{J}}_{\bf c} consists of all elements in 𝐉𝐜{\bf{J}}_{\bf c} annihilated by aj𝐜a_{j}{\mathbfcal{H}}_{\bf c} for all jij\neq i, so the decomposition of 𝐉𝐜{\bf{J}}_{\bf c} from Lemma 3.1.2 is stable under 𝐜{\mathbfcal{H}}_{\bf c}-linear endomorphisms. ∎

This finishes the proof of Theorem B.

4. Proof of Theorem A

4.1. A homomorphism θ:KeJe[𝐯,𝐯1]\theta\colon K_{e}\rightarrow J_{e}\otimes{\mathbb{C}}[{\bf{v}},{\bf{v}}^{-1}]

4.1.1.

Set

Ke:=KZe×(e×e),𝐊e:=𝐊Ze×(e×e).K_{e}:=K_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}^{\mathbb{C}^{*}}\times\mathcal{B}_{e}^{\mathbb{C}^{*}}),~{}{\bf{K}}_{e}:={\bf{K}}_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}^{\mathbb{C}^{*}}\times\mathcal{B}_{e}^{\mathbb{C}^{*}}).

Clearly, KeK_{e}, is a (unital) algebra w.r.t. the natural convolution product. Our goal in this section is to provide an injective homomorphism θ:KeJe[𝐯,𝐯1]\theta\colon K_{e}\to J_{e}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}]; it will be shown in Section 4.2 that θ\theta is an isomorphism.

We will construct θ\theta by introducing an Ke\mathcal{H}-K_{e}-bimodule e\mathcal{F}_{e}, checking that ee\mathcal{F}_{e}\simeq_{\mathcal{H}}\mathcal{H}_{e}, and invoking Theorem B.

Let us recall first (cf. [20], [36]) that \mathcal{H} can be identified with G×G^{\vee}\times\mathbb{C}^{*}-equivariant KK-theory KG×(St)K_{G^{\vee}\times\mathbb{C}^{*}}(\operatorname{St}) of the Steinberg variety St:=𝒩~×𝒩𝒩~\operatorname{St}:=\widetilde{{\mathcal{N}}}\times_{{\mathcal{N}}}\widetilde{{\mathcal{N}}}. We have the projection π:St𝒩\pi\colon\operatorname{St}\to\mathcal{N}; for a locally closed subvariety Z𝒩Z\subset\mathcal{N} we will denote π1(Z)\pi^{-1}(Z) by StZSt\operatorname{St}_{Z}\subset\operatorname{St}.

From [62] and [10, Theorem 55, §11.3] it follows that e\mathcal{H}_{e} is identified with KG×(St𝕆e)=KZe×(e×e)K_{G^{\vee}\times\mathbb{C}^{*}}(\operatorname{St}_{\mathbb{O}_{e}})=K_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e}); the isomorphism is compatible with the action of KZe×(pt)K_{Z_{e}\times\mathbb{C}^{*}}(\operatorname{pt}). Here, the \mathbb{C}^{*}-action comes from the element hh of an 𝔰𝔩2\mathfrak{sl}_{2} triple (e,h,f)(e,h,f); we fix such a triple. It is well-known that the centralizer ZG(e,h,f)Z_{G^{\vee}}(e,h,f) is a maximal reductive subgroup in ZG(e)Z_{G^{\vee}}(e), thus it is identified with ZeZ_{e}.

Let p:𝒩~𝒩p\colon\widetilde{\mathcal{N}}\to\mathcal{N} be the Springer resolution and Σe:=e+𝔷𝔤(f)𝔤\Sigma_{e}:=e+\mathfrak{z}_{\mathfrak{g}^{\vee}}(f)\subseteq\mathfrak{g}^{\vee} the Slodowy slice, we also let Λe𝒩~\Lambda_{e}\subset\widetilde{{\mathcal{N}}} denote p1(Σe𝒩)p^{-1}(\Sigma_{e}\cap\mathcal{N}). This is a smooth variety called Slodowy variety. We have eΛe\mathcal{B}_{e}\subseteq\Lambda_{e}; moreover, Λe=e\Lambda_{e}^{\mathbb{C}^{*}}=\mathcal{B}_{e}^{\mathbb{C}^{*}} (see e.g. [44, Section 1.8]).

Now we define

e:=KZe×(e×e)=KZe×(Λe×e;e×e).\mathcal{F}_{e}:=K_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e}^{\mathbb{C}^{*}})=K_{Z_{e}\times\mathbb{C}^{*}}(\Lambda_{e}\times\mathcal{B}_{e}^{\mathbb{C}^{*}};\mathcal{B}_{e}\times\mathcal{B}_{e}^{\mathbb{C}^{*}}).
Lemma 4.1.1.

Both KeK_{e} and e\mathcal{F}_{e} are torsion-free as \mathbb{Z}-modules.

Proof.

It follows from Proposition 4.1.5 below that there exists an identification ee\mathcal{F}_{e}\simeq\mathcal{H}_{e} of left \mathcal{H}-modules, so, in particular, they are isomorphic as \mathbb{Z}-modules. Let 𝐜W{\bf{c}}\subset W be the two-sided cell corresponding to ee. Recall that e=𝐜/<𝐜\mathcal{H}_{e}=\mathcal{H}_{\leqslant\bf{c}}/\mathcal{H}_{<\bf{c}} which is free over \mathbb{Z} with a basis consisting of {[Cw]𝐯k|w𝐜,k}\{[C_{w}]{\bf{v}}^{k}\,|\,w\in{\bf{c}},\,k\in\mathbb{Z}\}. We conclude that e\mathcal{F}_{e} is also free over \mathbb{Z}. Proposition 4.1.14 claims that there exists an isomorphism eKe\mathcal{F}_{e}\simeq K_{e}, implying that KeK_{e} is also free over \mathbb{Z}. ∎

4.1.2. The Ke\mathcal{H}-K_{e}-bimodule structure on e\mathcal{F}_{e}

Proposition 4.1.2.

There are natural commuting actions of \mathcal{H} and KeK_{e} on e\mathcal{F}_{e}.

Proof.

Proposition essentially follows from [43]. We sketch the proof for the reader’s convenience.

Let us define the action * of \mathcal{H} on e\mathcal{F}_{e}. Consider the diagram.

𝒩~×𝒩~×e=:X{{\widetilde{\mathcal{N}}\times\widetilde{\mathcal{N}}\times\mathcal{B}_{e}^{\mathbb{C}^{*}}=:X}}𝒩~×𝒩~{{\widetilde{\mathcal{N}}\times\widetilde{\mathcal{N}}}}𝒩~×e,{{\widetilde{\mathcal{N}}\times\mathcal{B}_{e}^{\mathbb{C}^{*}}},}p12\scriptstyle{p_{12}}p23\scriptstyle{p_{23}}

where p12p_{12}, p23p_{23} are the projections onto the corresponding factors.

Pick hKG×(𝒩~×𝒩𝒩~)h\in\mathcal{H}\simeq K_{G^{\vee}\times\mathbb{C}^{*}}(\widetilde{{\mathcal{N}}}\times_{{\mathcal{N}}}\widetilde{{\mathcal{N}}}) and aKZe×(e×e)a\in K_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e}^{\mathbb{C}^{*}}). Now
1) aa can be viewed as a class of an equivariant complex 𝒜\mathcal{A} on 𝒩~×e\widetilde{\mathcal{N}}\times\mathcal{B}_{e}^{\mathbb{C}^{*}} with support on e×e\mathcal{B}_{e}\times\mathcal{B}_{e}^{\mathbb{C}^{*}};
2) hh can be viewed as a class of an equivariant complex 𝒢\mathcal{G} on 𝒩~×𝒩~\widetilde{\mathcal{N}}\times\widetilde{\mathcal{N}} with support on 𝒩~×𝒩𝒩~\widetilde{\mathcal{N}}\times_{\mathcal{N}}\widetilde{\mathcal{N}}.

We define 𝒢𝒜=p13(p12(𝒢)𝕃Xp23(𝒜))\mathcal{G}*\mathcal{A}=p_{13*}(p_{12}^{*}(\mathcal{G})\otimes^{\mathbb{L}}_{X}p_{23}^{*}(\mathcal{A})), where p13:X𝒩~×ep_{13}\colon X\rightarrow\widetilde{{\mathcal{N}}}\times\mathcal{B}_{e}^{{\mathbb{C}}^{*}} is the projection onto the first and the third factors. Then we set ha:=[𝒢𝒜]h*a:=[\mathcal{G}*\mathcal{A}] (cf. ).

The fact that this defines an action of \mathcal{H} commuting with the action of KeK_{e} (defined below) follows by diagram chase.

Let us define the action of Ke=KZe×(e×e)K_{e}=K_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}^{\mathbb{C}^{*}}\times\mathcal{B}_{e}^{\mathbb{C}^{*}}) on

e=KZe×(Λe×e;e×e)=KZe×(𝒩~×e;e×e).\mathcal{F}_{e}=K_{Z_{e}\times\mathbb{C}^{*}}(\Lambda_{e}\times\mathcal{B}_{e}^{\mathbb{C}^{*}};\mathcal{B}_{e}\times\mathcal{B}_{e}^{\mathbb{C}^{*}})=K_{Z_{e}\times\mathbb{C}^{*}}(\widetilde{\mathcal{N}}\times\mathcal{B}_{e}^{\mathbb{C}^{*}};\mathcal{B}_{e}\times\mathcal{B}_{e}^{\mathbb{C}^{*}}).

Consider the following diagram:

Λe×e×e{{\Lambda_{e}\times\mathcal{B}_{e}^{\mathbb{C}^{*}}\times\mathcal{B}_{e}^{\mathbb{C}^{*}}}}Λe×e{{\Lambda_{e}\times\mathcal{B}_{e}^{\mathbb{C}^{*}}}}Λe×e{{\Lambda_{e}\times\mathcal{B}_{e}^{\mathbb{C}^{*}}}}e×e{{\mathcal{B}_{e}^{\mathbb{C}^{*}}\times\mathcal{B}_{e}^{\mathbb{C}^{*}}}}q13\scriptstyle{q_{13}}q12\scriptstyle{q_{12}}q23\scriptstyle{q_{23}}

it is clear that for QeQ\in\mathcal{F}_{e} and RKeR\in K_{e} the formula:

QR=q13(q12Qq23R)Q*R=q_{13*}(q_{12}^{*}Q\otimes q_{23}^{*}R)

gives a well-defined right KeK_{e}-action on e\mathcal{F}_{e}.

Lemma 4.1.3.

The induced action of e\mathcal{H}_{\leqslant e}\subset\mathcal{H} on e\mathcal{F}_{e} factors through the action of e\mathcal{H}_{e}.

Proof.

Let us first of all recall that we have the identification eKZe×(e×e)\mathcal{H}_{e}\simeq K_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e}). After this identification, the realization of e\mathcal{H}_{e} as a subquotient e/<e\mathcal{H}_{\leqslant e}/\mathcal{H}_{<e} of KG×(St)\mathcal{H}\simeq K_{G^{\vee}\times\mathbb{C}^{*}}(\operatorname{St}) can be described as follows. We have e=KG×(St𝕆¯e)\mathcal{H}_{\leqslant e}=K_{G^{\vee}\times\mathbb{C}^{*}}(\operatorname{St}_{\overline{\mathbb{O}}_{e}}), then the embedding KG×(St𝕆¯e)KG×(St)K_{G^{\vee}\times\mathbb{C}^{*}}(\operatorname{St}_{\overline{\mathbb{O}}_{e}})\hookrightarrow K_{G^{\vee}\times\mathbb{C}^{*}}(\operatorname{St}) is given by the pushforward for the closed embedding St𝕆¯eSt\operatorname{St}_{\overline{\mathbb{O}}_{e}}\hookrightarrow\operatorname{St} and the quotient KG×(St𝕆¯e)KZe×(e×e)=KG×(St𝕆e)K_{G^{\vee}\times\mathbb{C}^{*}}(\operatorname{St}_{\overline{\mathbb{O}}_{e}})\twoheadrightarrow K_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e})=K_{G^{\vee}\times\mathbb{C}^{*}}(\operatorname{St}_{\mathbb{O}_{e}}) is given by the restriction to the open subset St𝕆eSt𝕆¯e\operatorname{St}_{\mathbb{O}_{e}}\hookrightarrow\operatorname{St}_{\overline{\mathbb{O}}_{e}}.

So, our goal is to check that <e=KG×(St𝕆¯e𝕆e)\mathcal{H}_{<e}=K_{G^{\vee}\times\mathbb{C}^{*}}(\operatorname{St}_{\overline{\mathbb{O}}_{e}\setminus\mathbb{O}_{e}}) acts trivially on e=KZe×(e×e)\mathcal{F}_{e}=K_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e}^{\mathbb{C}^{*}}). Indeed, let 𝒢\mathcal{G} be an equivariant complex representing some class in KG×(St𝕆¯e𝕆e)K_{G^{\vee}\times\mathbb{C}^{*}}(\operatorname{St}_{\overline{\mathbb{O}}_{e}\setminus\mathbb{O}_{e}}), then the support of 𝒢\mathcal{G} is contained in St𝕆¯e𝕆e\operatorname{St}_{\overline{\mathbb{O}}_{e}\setminus\mathbb{O}_{e}}. Let 𝒜\mathcal{A} be an equivariant complex on 𝒩~×e\widetilde{\mathcal{N}}\times\mathcal{B}_{e}^{\mathbb{C}^{*}} representing a class in KZe×(e×e)K_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e}^{\mathbb{C}^{*}}), then the support of 𝒜\mathcal{A} is contained in e×e\mathcal{B}_{e}\times\mathcal{B}_{e}^{\mathbb{C}^{*}}. We conclude that the support of p12(𝒢)𝕃Xp23(𝒜)p_{12}^{*}(\mathcal{G})\otimes^{\mathbb{L}}_{X}p_{23}^{*}(\mathcal{A}) (see the notations from the proof of Proposition 4.1.2) is contained in p231(e×e)p121(St𝕆¯e𝕆e)=p_{23}^{-1}(\mathcal{B}_{e}\times\mathcal{B}_{e}^{\mathbb{C}^{*}})\cap p_{12}^{-1}(\operatorname{St}_{\overline{\mathbb{O}}_{e}\setminus\mathbb{O}_{e}})=\varnothing, hence the (derived) tensor product above is equal to zero. ∎

4.1.3. Alternative description of the action of e\mathcal{H}_{e} on e\mathcal{F}_{e}

We can define the convolution algebra structure on KZe×(e×e)K_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e}) by identifying it with KZe×(Λe×Λe;e×e)K_{Z_{e}\times\mathbb{C}^{*}}(\Lambda_{e}\times\Lambda_{e};\mathcal{B}_{e}\times\mathcal{B}_{e}), the KK-group of Ze×Z_{e}\times\mathbb{C}^{*}-equivariant coherent sheaves on Λe×Λe\Lambda_{e}\times\Lambda_{e} with set-theoretic support on e×e\mathcal{B}_{e}\times\mathcal{B}_{e} (cf. [56]). This algebra is isomorphic to e\mathcal{H}_{e} (see [56, Theorem B.2]). Let us describe the action KZe×(e×e)e=KZe×(e×e)K_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e})\curvearrowright\mathcal{F}_{e}=K_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e}^{\mathbb{C}^{*}}) geometrically. Consider the following diagram

Λe×Λe×e{{\Lambda_{e}\times\Lambda_{e}\times\mathcal{B}_{e}^{\mathbb{C}^{*}}}}Λe×Λe{{\Lambda_{e}\times\Lambda_{e}}}Λe×e{{\Lambda_{e}\times\mathcal{B}_{e}^{\mathbb{C}^{*}}}}Λe×e{{\Lambda_{e}\times\mathcal{B}_{e}^{\mathbb{C}^{*}}}}π12\scriptstyle{\pi_{12}}π23\scriptstyle{\pi_{23}}π13\scriptstyle{\pi_{13}}

Since e\mathcal{B}_{e} is proper, Λe\Lambda_{e} is smooth, and e\mathcal{B}_{e}^{\mathbb{C}^{*}} is smooth and proper, it is clear that for

[𝒫]KZe×(e×e),[𝒬]e=KZe×(e×e)[\mathcal{P}]\in K_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e}),~{}[\mathcal{Q}]\in\mathcal{F}_{e}=K_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e}^{\mathbb{C}^{*}})

the formula:

(4.1.1) [𝒫][𝒬]=[π13(π12𝒫π23𝒬)][\mathcal{P}]*[\mathcal{Q}]=[\pi_{13*}(\pi_{12}^{*}\mathcal{P}\otimes\pi_{23}^{*}\mathcal{Q})]

give a well-defined left KZe×(e×e)K_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e})-action on e\mathcal{F}_{e}.

Proposition 4.1.4.

After the identification eKZe×(e×e)\mathcal{H}_{e}\simeq K_{Z_{e}\times\mathbb{C^{*}}}(\mathcal{B}_{e}\times\mathcal{B}_{e}), the action given by (4.1.1) coincides with the action of e\mathcal{H}_{e} induced by the action of \mathcal{H} on e\mathcal{F}_{e} described in Proposition 4.1.2.

Proof.

Same argument as in the proof of [56, Theorem B.2] works. ∎

4.1.4.

In view of Theorem B, the next key proposition yields the desired homomorphism

(4.1.2) θ:KeEnd(e)opp=End(e)opp=Je[𝐯,𝐯1].\theta\colon K_{e}\to\operatorname{End}_{\mathcal{H}}(\mathcal{F}_{e})^{\mathrm{opp}}=\operatorname{End}_{\mathcal{H}}(\mathcal{H}_{e})^{\mathrm{opp}}=J_{e}\otimes\mathbb{C}[{\bf{v}},{\bf{v}}^{-1}].
Proposition 4.1.5.

We have a canonical isomorphism of left \mathcal{H}-modules: ee\mathcal{F}_{e}\simeq_{\mathcal{H}}\mathcal{H}_{e}.

Remark 4.1.6.

Notice that e\mathcal{F}_{e} splits as a direct sum indexed by components of e\mathcal{B}_{e}^{\mathbb{C}^{*}}, thus Proposition 4.1.5 implies such a decomposition for eKZe×(e×e)\mathcal{H}_{e}\simeq K^{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e}).

Similar direct sum decompositions are found in the literature. In particular, [30, Theorem 2.10] provides a semi-orthogonal decomposition for the derived category of equivariant coherent sheaves in a rather general situation, which implies a direct sum decomposition for the Grothendieck group. Existence of such a decomposition compatible with the convolution action would imply Proposition 4.1.5.555The similarity between Theorem 4.1.5 and results of [30] was pointed out to us by Do Kien Hoang.

On the other hand, [44, Lemma 14.9] essentially proves that for G=SLnG^{\vee}=SL_{n} the Bialynicky-Birula filtration on KT×(e)K_{T\times\mathbb{C}^{*}}(\mathcal{B}_{e}) can be split by means of the action of standard generators of Hecke algebra on KT×(e,1)K_{T\times\mathbb{C}^{*}}(\mathcal{B}_{e,1}).

We do not know if either of the two approaches yields a proof of Proposition 4.1.5.

Remark 4.1.7.

The proof of Proposition 4.1.5 below can be generalized to yield a similar direct sum decomposition for an arbitrary projective variety XX with a {\mathbb{C}}^{*}-action such that for every connected component FXF\subset X^{{\mathbb{C}}^{*}} the attractor of this component is smooth (in particular, the component itself is smooth) and the analog of Lemma 4.1.13 below holds.

Before proceeding to the proof of the Proposition we recall some geometric properties of the variety e\mathcal{B}_{e} found in [26]. Recall that we fix an 𝔰𝔩2\mathfrak{sl}_{2}-triple e,h,f𝔤e,h,f\in\mathfrak{g}^{\vee}. The adjoint action of hh on 𝔤\mathfrak{g}^{\vee} induces the decomposition 𝔤=i𝔤i\mathfrak{g}^{\vee}=\bigoplus_{i\in\mathbb{Z}}\mathfrak{g}^{\vee}_{i}. Let LL^{\vee} (resp. PP) be the connected algebraic subgroup of GG^{\vee} whose Lie algebra is 𝔤0\mathfrak{g}^{\vee}_{0} (resp. i0𝔤i\bigoplus_{i\geqslant 0}\mathfrak{g}^{\vee}_{i}). Recall that WfW_{f} is the Weyl group of GG^{\vee}, let WLWfW_{L}\subset W_{f} be the Weyl group of LL^{\vee}. Recall that =w¯WL\WfLwB/B\mathcal{B}^{\mathbb{C}^{*}}=\bigsqcup_{\bar{w}\in W_{L}\backslash W_{f}}L^{\vee}wB/B. For w¯WL\Wf\bar{w}\in W_{L}\backslash W_{f} we set w¯:=PwB/B\mathcal{B}_{\bar{w}}:=PwB/B\subset\mathcal{B}. We also set e,w¯:=w¯e\mathcal{B}_{e,\bar{w}}:=\mathcal{B}_{\bar{w}}\cap\mathcal{B}_{e}. Note that e,w¯\mathcal{B}_{e,\bar{w}} consists of points yey\in\mathcal{B}_{e} such that limt0tyLw¯B/B\underset{t\rightarrow 0}{\operatorname{lim}}\,t\cdot y\in L^{\vee}\bar{w}B/B, where the action of \mathbb{C}^{*} comes from the cocharacter of the center of LL^{\vee} that integrates hh.

Lemma 4.1.8.

Variety e,w¯\mathcal{B}_{e,\bar{w}} is smooth.

Proof.

[26, Proposition 3.2]. ∎

Warning 4.1.9.

Variety e,w¯\mathcal{B}_{e,\bar{w}} may be disconnected. It may also be empty.

Lemma 4.1.10.

The map xlimt0txx\mapsto\underset{t\rightarrow 0}{\operatorname{lim}}\,t\cdot x is an affine fibration e,w¯e,w¯\mathcal{B}_{e,\bar{w}}\twoheadrightarrow\mathcal{B}_{e,\bar{w}}^{{\mathbb{C}}^{*}}.

Proof.

Follows from [18, Theorem 4.1] (the assumptions of this theorem are satisfied for X=e,w¯X=\mathcal{B}_{e,\bar{w}} by Lemma 4.1.8 together with [54, Corollary 2 in Section 3]).

Another argument (that works only over fields of characteristic 0) can be found in [26, Section 3.4], where the authors use the result of Bass-Haboush (see [26, Section 1.5]) to obtain the statement. Finally, another argument (that works over fields of arbitrary characteristic) is given in [45, Section 5]. We are grateful to George Lusztig for pointing out this reference to us. ∎

Definition 4.1.11.

A partition of a variety XX as a finite union of locally closed subvarieties XiX_{i} is said to be an α\alpha-partition if the subvarieties in the partition can be indexed X1,,XnX_{1},\ldots,X_{n} in such a way that X1XkX_{1}\cup\ldots\cup X_{k} is closed in XX for k=1,,nk=1,\ldots,n.

Lemma 4.1.12.

Varieties e,w¯\mathcal{B}_{e,\bar{w}} form an α\alpha-partition of e\mathcal{B}_{e}.

Proof.

See [26, Section 3.4]. ∎

Proof.

(of Proposition 4.1.5) Step 1. By Lemma 4.1.12 varieties e,w¯\mathcal{B}_{e,\bar{w}} form an α\alpha-partition of e\mathcal{B}_{e}, i.e. there exists a labeling e,w¯i\mathcal{B}_{e,\bar{w}_{i}}, i=1,,ni=1,\ldots,n of these varieties such that i=1ke,w¯i\bigcup_{i=1}^{k}\mathcal{B}_{e,\bar{w}_{i}} is closed for k=1,,nk=1,\ldots,n.

This yields a filtration (that we denote by Φ\Phi) on KZe×(e×e)K_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e}).

Moreover, the locally closed strata e,w¯\mathcal{B}_{e,\bar{w}} are the attracting varieties for e,w¯\mathcal{B}_{e,\bar{w}}^{\mathbb{C}^{*}}, which is a union of components of e\mathcal{B}_{e}^{\mathbb{C}^{*}}. By Lemma 4.1.10 the map πw¯:e,w¯e,w¯\pi_{\bar{w}}\colon\mathcal{B}_{e,\bar{w}}\to\mathcal{B}_{e,\bar{w}}^{\mathbb{C}^{*}} is a vector bundle, hence it induces (cf. loc. cit.) an isomorphism πw¯:KZe×(e,w¯)KZe×(e,w¯)\pi_{\bar{w}}^{*}\colon K_{Z_{e}\times{\mathbb{C}}^{*}}(\mathcal{B}_{e,\bar{w}}^{{\mathbb{C}}^{*}})\,\vphantom{j^{X^{2}}}\smash{\overset{\sim}{\vphantom{\rule{0.0pt}{1.99997pt}}\smash{\longrightarrow}}}\,K_{Z_{e}\times{\mathbb{C}}^{*}}(\mathcal{B}_{e,\bar{w}}).

We claim that one gets a canonical isomorphism of KZe×(pt)K_{Z_{e}\times\mathbb{C}^{*}}(\operatorname{pt})-modules: grΦKZe×(e×e)=e\operatorname{gr_{\Phi}}K_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e})=\mathcal{F}_{e}.

To check this, one has to show that all of the maps Φk:=KZe×(e×(i=1ke,w¯i))KZe×(e×e)\Phi_{k}:=K_{Z_{e}\times{\mathbb{C}}^{*}}(\mathcal{B}_{e}\times(\bigcup_{i=1}^{k}\mathcal{B}_{e,\bar{w}_{i}}))\to K_{Z_{e}\times{\mathbb{C}}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e}) are injective.

For this, see the Lemma 4.1.13 below.

In order to finish the proof of the Proposition it now suffices to provide an e\mathcal{H}_{e}- and ReR_{e}-equivariant splitting of Φ\Phi.

Step 2. For every w¯\bar{w}, πw¯\pi_{\bar{w}} together with the locally closed embedding ιw¯:e,w¯e\iota_{\bar{w}}\colon\mathcal{B}_{e,\bar{w}}\hookrightarrow\mathcal{B}_{e} fits into the diagram:

e,w¯{{\mathcal{B}_{e,\bar{w}}}}e{{\mathcal{B}_{e}}}e,w¯.{{\mathcal{B}_{e,\bar{w}}^{\mathbb{C}^{*}}.}}πw¯\scriptstyle{\pi_{\bar{w}}}ιw¯\scriptstyle{\iota_{\bar{w}}}

We claim that Im(ιw¯×πw¯)\operatorname{Im}(\iota_{\bar{w}}\times\pi_{\bar{w}}) is locally closed inside e×e,w\mathcal{B}_{e}\times\mathcal{B}_{e,w}^{\mathbb{C}^{*}}, or, equivalently, inside ¯e,w¯×e,w¯\overline{\mathcal{B}}_{e,\bar{w}}\times\mathcal{B}_{e,\bar{w}}^{\mathbb{C}^{*}}. To see this, consider the diagram:

e,w¯{{\mathcal{B}_{e,\bar{w}}}}e,w¯×e,w¯{{\mathcal{B}_{e,\bar{w}}\times\mathcal{B}_{e,\bar{w}}^{\mathbb{C}^{*}}}}¯e,w¯×e,w¯,{{\overline{\mathcal{B}}_{e,\bar{w}}\times\mathcal{B}_{e,\bar{w}}^{\mathbb{C}^{*}}},}

in which the first map is the graph map for πw¯\pi_{\bar{w}} which is clearly closed, and the second one is the open embedding. The claim follows.

Now let Γw¯\Gamma_{\bar{w}} be the closure of the image of ιw¯×πw¯\iota_{\bar{w}}\times\pi_{\bar{w}} inside e×e,w¯\mathcal{B}_{e}\times\mathcal{B}_{e,\bar{w}}^{\mathbb{C}^{*}}. It is a proper Ze×Z_{e}\times\mathbb{C}^{*}-subvariety with an open subset isomorphic to e,w¯\mathcal{B}_{e,\bar{w}}. Moreover, there are canonical Ze×Z_{e}\times\mathbb{C}^{*}-equivariant projections ι~w¯\widetilde{\iota}_{\bar{w}} and π~w¯\widetilde{\pi}_{\bar{w}} from Γw¯\Gamma_{\bar{w}} to e\mathcal{B}_{e} and e,w¯\mathcal{B}_{e,\bar{w}}^{\mathbb{C}^{*}} respectively.

Step 3. Now induction by k=1,,nk=1,\ldots,n shows that the map of the form

i=1n(Ide×ι~w¯i)(Ide×π~w¯i)\bigoplus_{i=1}^{n}(\operatorname{Id}_{\mathcal{B}_{e}}\times\widetilde{\iota}_{\bar{w}_{i}})_{*}(\operatorname{Id}_{\mathcal{B}_{e}}\times\widetilde{\pi}_{\bar{w}_{i}})^{*}

splits the filtration. (We recall that Ide×πw¯\operatorname{Id}_{\mathcal{B}_{e}}\times\pi_{\bar{w}} is a vector bundle, hence it induces an isomorphism in equivariant KK-theory.)

Note that (Ide×π~w¯)(\operatorname{Id}_{\mathcal{B}_{e}}\times\widetilde{\pi}_{\bar{w}})^{*} is well-defined: pullback in KK-theory is well-defined for morphisms with smooth target or a base change of such morphisms, while Ide×π~w¯\operatorname{Id}_{\mathcal{B}_{e}}\times\widetilde{\pi}_{\bar{w}} is a base change of the morphism π~w¯\widetilde{\pi}_{\bar{w}} with smooth target e,w¯\mathcal{B}_{e,\bar{w}}^{\mathbb{C}^{*}}.

Step 4. The argument similar to the one in the last paragraph shows that \mathcal{H} has a canonical left convolution action on KZe×(e×Γw¯)K_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}\times\Gamma_{\bar{w}}). Moreover, (Ide×ι~w¯)(\operatorname{Id}_{\mathcal{B}_{e}}\times\widetilde{\iota}_{\bar{w}})_{*} and (Ide×π~w¯)(\operatorname{Id}_{\mathcal{B}_{e}}\times\widetilde{\pi}_{\bar{w}})^{*} are \mathcal{H}-equivariant.

Let us, for example, prove this for (Ide×π~w¯)(\operatorname{Id}_{\mathcal{B}_{e}}\times\widetilde{\pi}_{\bar{w}})^{*}.

We have:

(4.1.3) 𝒩~×𝒩~×Γw{{\widetilde{\mathcal{N}}\times\widetilde{\mathcal{N}}\times\Gamma_{w}}}𝒩~×𝒩~×e{{\widetilde{\mathcal{N}}\times\widetilde{\mathcal{N}}\times\mathcal{B}_{e}^{\mathbb{C}^{*}}}}𝒩~×Γw¯{{\widetilde{\mathcal{N}}\times\Gamma_{\bar{w}}}}𝒩~×e{{\widetilde{\mathcal{N}}\times\mathcal{B}_{e}^{\mathbb{C}^{*}}}}Id×Id×π~w¯\scriptstyle{{\operatorname{Id}\times\operatorname{Id}\times\widetilde{\pi}_{\bar{w}}}}π13\scriptstyle{{\pi_{13}}}Id×π~w¯\scriptstyle{{\operatorname{Id}\times\widetilde{\pi}_{\bar{w}}}}q13\scriptstyle{q_{13}}

We need to prove that for [𝒜]KG×(𝒩~×𝒩𝒩~)[\mathcal{A}]\in K_{G^{\vee}\times\mathbb{C}^{*}}(\widetilde{\mathcal{N}}\times_{\mathcal{N}}\widetilde{\mathcal{N}}) and [𝒢]KZe×(𝒩~×e;e×e)[\mathcal{G}]\in K_{Z_{e}\times\mathbb{C}^{*}}(\widetilde{\mathcal{N}}\times\mathcal{B}_{e}^{\mathbb{C}^{*}};\mathcal{B}_{e}\times\mathcal{B}_{e}^{\mathbb{C}^{*}}) the following equality holds:

(4.1.4) [(Id×π~w¯)π13(π12𝒜𝕃π23𝒢)]=[q13(q12𝒜𝕃q23(Id×π~w¯)𝒢)],\Big{[}(\operatorname{Id}\times\widetilde{\pi}_{\bar{w}})^{*}\pi_{13*}(\pi_{12}^{*}\mathcal{A}\otimes^{\mathbb{L}}\pi_{23}^{*}\mathcal{G})\Big{]}=\Big{[}q_{13*}(q_{12}^{*}\mathcal{A}\otimes^{\mathbb{L}}q_{23}^{*}(\operatorname{Id}\times\widetilde{\pi}_{\bar{w}})^{*}\mathcal{G})\Big{]},

where πij\pi_{ij}, (resp. qijq_{ij}), {i,j}{1,2,3}\{i,j\}\subset\{1,2,3\} are projections of 𝒩~×𝒩~×Γw\widetilde{\mathcal{N}}\times\widetilde{\mathcal{N}}\times\Gamma_{w} (resp. 𝒩~×𝒩~×e\widetilde{\mathcal{N}}\times\widetilde{\mathcal{N}}\times\mathcal{B}_{e}^{\mathbb{C}^{*}}) to the corresponding factors.

The equality (4.1.4) is clear since the diagram (4.1.3) is Cartesian, the map π13\pi_{13} is flat, and the result of [58, Proposition 1.4] holds.

This finishes the proof. ∎

Lemma 4.1.13.

The natural maps KZe×(e×(i=1ke,w¯i))KZe×(e×e)K_{Z_{e}\times{\mathbb{C}}^{*}}(\mathcal{B}_{e}\times(\bigcup_{i=1}^{k}\mathcal{B}_{e,\bar{w}_{i}}))\to K_{Z_{e}\times{\mathbb{C}}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e}) are injective for all kk.

Proof.

Step 1. To prove the sought-for injectivity, it suffices to construct the map

κ:KZe×(e×e)KZe×(e×e)\kappa\colon K_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e})\to K_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e}^{\mathbb{C}^{*}})

so that the maps κ(Ide×ι~w¯i)(Ide×π~w¯i)\kappa\circ(\operatorname{Id}_{\mathcal{B}_{e}}\times\widetilde{\iota}_{\bar{w}_{i}})_{*}(\operatorname{Id}_{\mathcal{B}_{e}}\times\widetilde{\pi}_{\bar{w}_{i}})^{*} composed with the projections KZe×(e×e)KZe×(e×e,w¯i)K_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e}^{\mathbb{C}^{*}})\rightarrow K_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e,\bar{w}_{i}}^{\mathbb{C}^{*}}) are injective for all ii.

We will make use of the Drinfeld-Gaitsgory degeneration (cf. [25, Section 2]), and the specialization in the equivariant KK-theory (cf. [20, Section 5.3]).

First of all, we briefly recall both constructions.

Step 2. First, for an algebraic variety MM equipped with a \mathbb{C}^{*}-action, the DG-degeneration is the family M~\widetilde{M}\to\mathbb{C} so that M~t\widetilde{M}_{t} can be canonically identified with MM, and M~0w¯iMw¯i+×Mw¯iMw¯i\widetilde{M}_{0}\simeq\bigsqcup_{\bar{w}_{i}}M_{\bar{w}_{i}}^{+}\times_{M_{\bar{w}_{i}}^{\mathbb{C}^{*}}}M_{\bar{w}_{i}}^{-}. Here, for the given Mw¯iM_{\bar{w}_{i}}^{\mathbb{C}^{*}}, Mw¯i+M_{\bar{w}_{i}}^{+} stands for the corresponding attracting set, and Mw¯iM_{\bar{w}_{i}}^{-} stands for the repelling one. Moreover, there exists a canonical global trivialization of M~\widetilde{M} over 𝔾m\mathbb{G}_{m}: πM:M~|𝔾mM×𝔾m\pi_{M}\colon\widetilde{M}|_{\mathbb{G}_{m}}\simeq M\times\mathbb{G}_{m}.

By inspection of the constructions from loc. cit., one sees that for ZZ equipped with the action of an algebraic group HH (so that the HH-action commutes with \mathbb{C}^{*}), the family also carries an action of HH compatible with the trivial action on the base.

Second, if XX\to\mathbb{C} is any HH-equivariant algebraic family (for example, the one above), the map lim0:KH(XX0)KH(X0)\operatorname{lim}_{0}\colon K_{H}(X\setminus X_{0})\to K_{H}(X_{0}) can be defined.

Step 3. Now, let us consider e×e\mathcal{B}_{e}\times\mathcal{B}_{e} with the \mathbb{C}^{*}-action on the second factor as MM.

The desired map κ\kappa is as follows.

For any class []KZe×(M)[\mathcal{F}]\in K_{Z_{e}\times\mathbb{C}^{*}}(M), one may consider the class

π+lim0πM([𝒪𝔾m])KZe×(M+)KZe×(M)\pi_{+*}\operatorname{lim}_{0}\pi_{M}^{*}([\mathcal{F}\boxtimes\mathcal{O}_{\mathbb{G}_{m}}])\in K_{Z_{e}\times\mathbb{C}^{*}}(M^{+})\simeq K_{Z_{e}\times\mathbb{C}^{*}}(M^{\mathbb{C}^{*}})

for π+\pi_{+} being the projection M+×MMM+M^{+}\times_{M^{\mathbb{C}^{*}}}M^{-}\to M^{+}.

Let us now prove the fact that the composition κ(Ide×ι~w¯i)(Ide×π~w¯i)[𝒢]\kappa\circ(\operatorname{Id}_{\mathcal{B}_{e}}\times\widetilde{\iota}_{\bar{w}_{i}})_{*}(\operatorname{Id}_{\mathcal{B}_{e}}\times\widetilde{\pi}_{\bar{w}_{i}})^{*}[\mathcal{G}] restricted to M+w¯iM^{+}_{\bar{w}_{i}} is non-zero for a non-zero [𝒢][\mathcal{G}].

Let SMS\subset M be the support of (Ide×ι~w¯i)(Ide×π~w¯i)[𝒢](\operatorname{Id}_{\mathcal{B}_{e}}\times\widetilde{\iota}_{\bar{w}_{i}})_{*}(\operatorname{Id}_{\mathcal{B}_{e}}\times\widetilde{\pi}_{\bar{w}_{i}})^{*}[\mathcal{G}]. Clearly, SS is contained in Im(e×ι~w¯i)\operatorname{Im}(\mathcal{B}_{e}\times\widetilde{\iota}_{\bar{w}_{i}}) so it lies in the closure M+w¯i¯\overline{M^{+}_{\bar{w}_{i}}} to the component of MM^{\mathbb{C}^{*}} corresponding to w¯i\bar{w}_{i}. Set Z:=M+w¯i¯Z:=\overline{M^{+}_{\bar{w}_{i}}} and let Z~\widetilde{Z} be the Drinfeld-Gaitsgory interpolation of ZZ. Note that Z~\widetilde{Z} is closed in M~\widetilde{M} by [25, Proposition 2.3.2 (i)]. Note also that M+w¯iZ+×ZZ=Z0M^{+}_{\bar{w}_{i}}\subset Z^{+}\times_{Z^{\mathbb{C}^{*}}}Z^{-}=Z_{0} is a union of connected components of Z0Z_{0}.

Let j0j_{0} be an evident open (and closed) embedding M+w¯iZ+×ZZ=Z0M^{+}_{\bar{w}_{i}}\hookrightarrow Z^{+}\times_{Z^{\mathbb{C}^{*}}}Z^{-}=Z_{0}, and let jtj_{t} be an evident open embedding M+w¯iZtM^{+}_{\bar{w}_{i}}\to Z_{t}. They glue to an open embedding j:M+w¯i×Z~j\colon M^{+}_{\bar{w}_{i}}\times\mathbb{C}\rightarrow\widetilde{Z}. Let j0j_{\neq 0} be the restriction of jj to M+w¯i×𝔾mM^{+}_{\bar{w}_{i}}\times\mathbb{G}_{m}.

Then,

(4.1.5) j0lim0πM((Ide×ι~w¯i)(Ide×π~w¯i)[𝒢]𝒪𝔾m)=\displaystyle j_{0}^{*}\operatorname{lim}_{0}\pi_{M}^{*}((\operatorname{Id}_{\mathcal{B}_{e}}\times\widetilde{\iota}_{\bar{w}_{i}})_{*}(\operatorname{Id}_{\mathcal{B}_{e}}\times\widetilde{\pi}_{\bar{w}_{i}})^{*}[\mathcal{G}]\boxtimes\mathcal{O}_{\mathbb{G}_{m}})=
(4.1.6) lim0j0π((Ide×ι~w¯i)(Ide×π~w¯i)[𝒢]𝒪𝔾m)=\displaystyle\operatorname{lim}_{0}j_{\neq 0}^{*}\pi^{*}((\operatorname{Id}_{\mathcal{B}_{e}}\times\widetilde{\iota}_{\bar{w}_{i}})_{*}(\operatorname{Id}_{\mathcal{B}_{e}}\times\widetilde{\pi}_{\bar{w}_{i}})^{*}[\mathcal{G}]\boxtimes\mathcal{O}_{\mathbb{G}_{m}})=
(4.1.7) [𝒢]0\displaystyle[\mathcal{G}]\neq 0

Here in the second step we use that the embedding j0j_{\neq 0} degenerates, via the Drinfeld-Gaitsgory family, to the embedding j0j_{0} and take the limit for the corresponding trivial family M+×M^{+}\times\mathbb{C}, – and [20, Lemma 5.3.6] (cf. also [20, Theorem 5.3.9]). We have also omitted the Thom isomorphism πw¯i\pi_{\bar{w}_{i}}^{*} from the notation.

The claim follows.

The proof of the following fact is similar to the above discussion.

Proposition 4.1.14.

As a right KeK_{e}-module, e\mathcal{F}_{e} is isomorphic to KeK_{e}.

Since e\mathcal{B}_{e}^{\mathbb{C}^{*}} is smooth, KeK_{e} is a unital algebra, where the unit element is the class [𝒪Δe][\mathcal{O}_{\Delta_{\mathcal{B}_{e}^{{\mathbb{C}}^{*}}}}] of the structure sheaf of the diagonal Δee×e\Delta_{\mathcal{B}_{e}^{{\mathbb{C}}^{*}}}\subset\mathcal{B}_{e}^{{\mathbb{C}}^{*}}\times\mathcal{B}_{e}^{{\mathbb{C}}^{*}}.

Corollary 4.1.15.

The map θ\theta is injective.

Proof.

Let Ξ:Kee\Xi\colon K_{e}\,\vphantom{j^{X^{2}}}\smash{\overset{\sim}{\vphantom{\rule{0.0pt}{1.99997pt}}\smash{\longrightarrow}}}\,\mathcal{F}_{e} be the identification of right KeK_{e}-modules (see Proposition 4.1.14 above). It follows from the definitions that for xKex\in K_{e}, we have x=Ξ1(θ(x)(Ξ(1)))x=\Xi^{-1}(\theta(x)(\Xi(1))), i.e., xx is uniquely determined by θ(x)\theta(x). The claim follows. ∎

4.2. Surjectivity of θ\theta

To finish the proof of Theorem A it remains to show that θ\theta is surjective. Recall that Ξ:Kee\Xi\colon K_{e}\,\vphantom{j^{X^{2}}}\smash{\overset{\sim}{\vphantom{\rule{0.0pt}{1.99997pt}}\smash{\longrightarrow}}}\,\mathcal{F}_{e} is the isomorphism of Proposition 4.1.14, it is compatible with the ReR_{e}-action and the right KeK_{e}-action.

Consider the identifications KeeeJe[𝐯,𝐯1]K_{e}\,\vphantom{j^{X^{2}}}\smash{\overset{\sim}{\vphantom{\rule{0.0pt}{1.99997pt}}\smash{\longrightarrow}}}\,\mathcal{F}_{e}\simeq\mathcal{H}_{e}\simeq J_{e}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}] given, respectively, by Ξ\Xi, by Proposition 4.1.5 and by Corollary 2.3.2. Let aJe[𝐯,𝐯1]a\in J_{e}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}] be the image of 1Ke1\in K_{e}.

Lemma 4.2.1.

The element aJe[𝐯,𝐯1]a\in J_{e}\otimes{\mathbb{C}}[{\bf{v}},{\bf{v}}^{-1}] is left invertible.

Proof.

It follows from Proposition 4.1.14 that the element Ξ(1)\Xi(1) generates e\mathcal{F}_{e} under the right action of KeK_{e}, hence, also under the right action of Je[𝐯,𝐯1]=End(e)oppJ_{e}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}]=\operatorname{End}_{\mathcal{H}}(\mathcal{F}_{e})^{\mathrm{opp}}; here we use that the KeK_{e}-action on ee\mathcal{F}_{e}\simeq_{\mathcal{H}}\mathcal{H}_{e} comes from the homomorphism KeEnd(e)oppK_{e}\rightarrow\operatorname{End}_{\mathcal{H}}(\mathcal{F}_{e})^{\mathrm{opp}}. Generators of a free rank one right module are exactly right invertible elements. We conclude that there exists bJe[𝐯,𝐯1]b\in J_{e}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}] such that ab=1ab=1. Recall that 𝐉e{\bf{J}}_{e} is the complexification of JeJ_{e}. Since 𝐉e{\bf{J}}_{e} is left-Noetherian (being a finite module over a Noetherian central subalgebra, see e.g. [39, Proposition 1.6]), it is a Dedekind-finite ring, so the right invertible element aa is also left invertible. In particular, it follows that the element bJe[𝐯,𝐯1]b\in J_{e}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}] is the left inverse to aa. ∎

Corollary 4.2.2.

An element φEnd(e)\varphi\in\operatorname{End}_{\mathcal{H}}(\mathcal{F}_{e}) is uniquely determined by its value on Ξ(1)\Xi(1).

Proof.

By Theorem B such an endomorphism φ\varphi is given by right multiplication by an element of Je[𝐯,𝐯1]J_{e}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}], since aa is left invertible, the claim follows. ∎

Since Ξ(1)\Xi(1) is a free generator of e\mathcal{F}_{e} as a right KeK_{e}-module, Corollary 4.2.2 implies surjectivity of θ\theta. This completes the proof of Theorem A, establishing the isomorphism

(4.2.1) KeJe[𝐯,𝐯1].K_{e}\simeq J_{e}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}].
Remark 4.2.3.

As pointed out above, by Theorem A the monoidal category Db(CohZe(e×e))D^{b}(\operatorname{Coh}_{Z_{e}}(\mathcal{B}_{e}\times\mathcal{B}_{e})) can be viewed as a categorification of the algebra JeJ_{e}. It would be interesting to find a compatible categorification of the homomorphism ϕ𝐜:𝐜J𝐜[𝐯,𝐯1]\phi_{{\bf{c}}}\colon\mathcal{H}_{\bf{c}}\rightarrow J_{\bf{c}}\otimes{\mathbb{C}}[{\bf{v}},{\bf{v}}^{-1}]. We do not have a proposal for a monoidal functor categorifying ϕ𝐜\phi_{\bf c}; however, our proof provides a categorification for the bimodule structure on the target ring arising from that homomorphism: the above argument shows this bimodule is identified with e\mathcal{F}_{e}, the bimodule category Db(CohZe×(e×e))D^{b}(\operatorname{Coh}_{Z_{e}\times{\mathbb{C}}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e}^{{\mathbb{C}}^{*}})) is its categorification.

Corollary 4.2.4.

We have EndeRe(e)opp=Je[𝐯,𝐯1].\operatorname{End}_{\mathcal{H}_{e}}^{R_{e}}(\mathcal{F}_{e})^{\mathrm{opp}}=J_{e}\otimes\mathbb{C}[{\bf{v}},{\bf{v}}^{-1}].

Proof.

Follows from Theorem B together with Proposition 4.1.5. ∎

4.3. Towards the geometric description of ϕ𝐜\phi^{\bf{c}} and ϕ𝐜\phi_{\bf{c}}

4.3.1.

It follows from the definitions that the homomorphisms

ϕ𝐜:KG×(𝒩~×𝒩𝒩~)KZe×(e×e),ϕ𝐜:KZe×(e×e)KZe×(e×e)\phi^{\bf{c}}\colon K_{G^{\vee}\times\mathbb{C}^{*}}(\widetilde{\mathcal{N}}\times_{\mathcal{N}}\widetilde{\mathcal{N}})\rightarrow K_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}^{\mathbb{C}^{*}}\times\mathcal{B}_{e}^{\mathbb{C}^{*}}),~{}\phi_{\bf{c}}\colon K_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e})\rightarrow K_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}^{\mathbb{C}^{*}}\times\mathcal{B}_{e}^{\mathbb{C}^{*}})

can be described as follows. Let

Ξ:Ke=KZe×(e×e)KZe×(e×e)=e\Xi\colon K_{e}=K_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}^{\mathbb{C}^{*}}\times\mathcal{B}_{e}^{\mathbb{C}^{*}})\,\vphantom{j^{X^{2}}}\smash{\overset{\sim}{\vphantom{\rule{0.0pt}{1.99997pt}}\smash{\longrightarrow}}}\,K_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e}^{\mathbb{C}^{*}})=\mathcal{F}_{e}

be the identification given by i=1n(ι~w¯i×Ide)(π~w¯i×Ide)\bigoplus_{i=1}^{n}(\widetilde{\iota}_{\bar{w}_{i}}\times\operatorname{Id}_{\mathcal{B}_{e}^{\mathbb{C}^{*}}})_{*}(\widetilde{\pi}_{\bar{w}_{i}}\times\operatorname{Id}_{\mathcal{B}_{e}^{\mathbb{C}^{*}}})^{*} (see the notation in the proof of Step 33 of Proposition 4.1.5).

Proposition 4.3.1.

The homomorphism ϕ𝐜\phi^{\bf{c}} is given by

(4.3.1) KG×(𝒩~×𝒩𝒩~)xΞ1(xΞ([Δe]))KZe×(e×e),K_{G^{\vee}\times\mathbb{C}^{*}}(\widetilde{\mathcal{N}}\times_{\mathcal{N}}\widetilde{\mathcal{N}})\ni x\mapsto\Xi^{-1}(x*\Xi([\Delta_{\mathcal{B}_{e}^{\mathbb{C}^{*}}}]))\in K_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}^{\mathbb{C}^{*}}\times\mathcal{B}_{e}^{\mathbb{C}^{*}}),

where * is the convolution action of KG×(𝒩~×𝒩𝒩~)K_{G^{\vee}\times\mathbb{C}^{*}}(\widetilde{\mathcal{N}}\times_{\mathcal{N}}\widetilde{\mathcal{N}}) on KZe×(e×e)K_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e}^{\mathbb{C}^{*}}).

The homomorphism ϕ𝐜\phi_{\bf{c}} has the same description with * being replaced by the convolution action of KZe×(e×e)K_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e}) on KZe×(e×e)K_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e}^{\mathbb{C}^{*}}).

Proof.

The homomorphism ϕ𝐜\phi^{\mathbf{c}} is given by the bimodule e=KZe×(e×e)\mathcal{F}_{e}=K_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e}^{\mathbb{C}^{*}}). Namely, for xKG×(𝒩~×𝒩𝒩~)x\in K_{G^{\vee}\times\mathbb{C}^{*}}(\widetilde{\mathcal{N}}\times_{\mathcal{N}}\widetilde{\mathcal{N}}), its image in Ke=KZe×(e×e)K_{e}=K_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}^{\mathbb{C}^{*}}\times\mathcal{B}_{e}^{\mathbb{C}^{*}}) under ϕ𝐜\phi^{\bf{c}} is obtained as follows: we consider the operator xEndKe(e)oppx*-\in\operatorname{End}_{K_{e}}(\mathcal{F}_{e})^{\mathrm{opp}} and use the isomorphism Ξ:Kee\Xi\colon K_{e}\,\vphantom{j^{X^{2}}}\smash{\overset{\sim}{\vphantom{\rule{0.0pt}{1.99997pt}}\smash{\longrightarrow}}}\,\mathcal{F}_{e} of right KeK_{e}-modules to identify EndKe(e)opp=EndKe(Ke)opp=Ke\operatorname{End}_{K_{e}}(\mathcal{F}_{e})^{\mathrm{opp}}=\operatorname{End}_{K_{e}}(K_{e})^{\mathrm{opp}}=K_{e}. Then, ϕ𝐜(x)\phi^{\bf{c}}(x) is the element of KeK_{e} corresponding to xx*-. In other words, ϕ𝐜(x)\phi^{\bf{c}}(x) is the element of KeK_{e} such that

xΞ(y)=Ξ(yϕ𝐜(x))x*\Xi(y)=\Xi(y*\phi^{\bf{c}}(x))

for every yKey\in K_{e}. Substituting y=[Δe]y=[\Delta_{\mathcal{B}_{e}^{\mathbb{C}^{*}}}], we conclude that ϕ𝐜(x)=Ξ1(xΞ([Δe]))\phi^{\bf{c}}(x)=\Xi^{-1}(x*\Xi([\Delta_{\mathcal{B}_{e}^{\mathbb{C}^{*}}}])). ∎

Remark 4.3.2.

Note that Ξ([Δe])\Xi([\Delta_{\mathcal{B}_{e}^{\mathbb{C}^{*}}}]) can be explicitly described, it is equal to the structure sheaf of the disjoint union iΓw¯i\bigsqcup_{i}\Gamma_{\bar{w}_{i}}.

4.3.2.

To every character λ\lambda of the maximal torus of GG^{\vee} we can associate the corresponding induced line bundle 𝒪(λ)\mathcal{O}_{\mathcal{B}}(\lambda). Let 𝒪𝒩~(λ)\mathcal{O}_{\widetilde{\mathcal{N}}}(\lambda) be the pull back of 𝒪(λ)\mathcal{O}_{\mathcal{B}}(\lambda) to 𝒩~=T\widetilde{\mathcal{N}}=T^{*}\mathcal{B}. Let Δ:𝒩~𝒩~×𝒩𝒩~\Delta\colon\widetilde{\mathcal{N}}\hookrightarrow\widetilde{\mathcal{N}}\times_{\mathcal{N}}\widetilde{\mathcal{N}} be the diagonal embedding. The elements [Δ𝒪(λ)]KG×(𝒩~×𝒩𝒩~)[\Delta_{*}\mathcal{O}_{\mathcal{B}}(\lambda)]\in K_{G^{\vee}\times\mathbb{C}^{*}}(\widetilde{\mathcal{N}}\times_{\mathcal{N}}\widetilde{\mathcal{N}})\simeq\mathcal{H} form the so-called ‘‘lattice part’’ of the affine Hecke algebra \mathcal{H} (c.f. [20, Section 7]).

Remark 4.3.3.

Note that the action of [Δ𝒪(λ)][\Delta_{*}\mathcal{O}_{\mathcal{B}}(\lambda)] on KZe×(e×e)K_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e}^{\mathbb{C}^{*}}) is given by (𝒪e(λ)𝒪e)(\mathcal{O}_{\mathcal{B}_{e}}(\lambda)\boxtimes\mathcal{O}_{\mathcal{B}_{e}^{\mathbb{C}^{*}}})\otimes-.

We will denote by 𝒪e(λ)\mathcal{O}_{\mathcal{B}_{e}^{\mathbb{C}^{*}}}(\lambda) the restriction of 𝒪(λ)\mathcal{O}_{\mathcal{B}}(\lambda) to e\mathcal{B}_{e}^{\mathbb{C}^{*}}. Abusing the notation, we will denote by Δ𝒪e(λ)\Delta_{*}\mathcal{O}_{\mathcal{B}_{e}^{\mathbb{C}^{*}}}(\lambda) the push forward of 𝒪e(λ)\mathcal{O}_{\mathcal{B}_{e}^{\mathbb{C}^{*}}}(\lambda) under the diagonal embedding Δ:ee×e\Delta\colon\mathcal{B}_{e}^{\mathbb{C}^{*}}\hookrightarrow\mathcal{B}_{e}^{\mathbb{C}^{*}}\times\mathcal{B}_{e}^{\mathbb{C}^{*}}.

Warning 4.3.4.

It is not true in general that for yKG×(𝒩~×𝒩𝒩~)y\in K_{G^{\vee}\times\mathbb{C}^{*}}(\widetilde{{\mathcal{N}}}\times_{{\mathcal{N}}}\widetilde{{\mathcal{N}}}), the element ϕ𝐜([Δ𝒪𝒩~(λ)]y)\phi^{\bf{c}}([\Delta_{*}\mathcal{O}_{\widetilde{\mathcal{N}}}(\lambda)]*y) coincides with ϕ𝐜(y)Δ𝒪e(λ)\phi^{\bf{c}}(y)*\Delta_{*}\mathcal{O}_{\mathcal{B}_{e}^{\mathbb{C}^{*}}}(\lambda).

Informally, the reason that leads to the Warning can be illustrated by the following example. Consider 1\mathbb{P}^{1} with the action of \mathbb{C}^{*} on it given by [a:b][ta:b][a:b]\mapsto[ta:b]. Then, the set (1)(\mathbb{P}^{1})^{\mathbb{C}^{*}} consits of two points 0=[0:1]0=[0:1], =[1:0]\infty=[1:0]. The identification Ξ:K({0,})K(1)\Xi\colon K(\{0,\infty\})\,\vphantom{j^{X^{2}}}\smash{\overset{\sim}{\vphantom{\rule{0.0pt}{1.99997pt}}\smash{\longrightarrow}}}\,K(\mathbb{P}^{1}) is induced by the correspondence Γ:={0}×1×\Gamma:=\{0\}\times\mathbb{P}^{1}\sqcup\infty\times\infty. It follows from the definition that this identification is given by

[0][𝒪1],[][],[\mathbb{C}_{0}]\mapsto[\mathcal{O}_{\mathbb{P}^{1}}],~{}[\mathbb{C}_{\infty}]\mapsto[\mathbb{C}_{\infty}],

where by p\mathbb{C}_{p} we denote the skyscraper sheaf at the point point pp.

We see that

[𝒪1]=Ξ([0])=Ξ([𝒪1(1)|{0}])[𝒪1(1)]Ξ([0])=[𝒪1(1)],[\mathcal{O}_{\mathbb{P}^{1}}]=\Xi([\mathbb{C}_{0}])=\Xi([\mathcal{O}_{\mathbb{P}^{1}}(1)|_{\{0\}}])\neq[\mathcal{O}_{\mathbb{P}^{1}}(1)]\otimes\Xi([\mathbb{C}_{0}])=[\mathcal{O}_{\mathbb{P}^{1}}(1)],

i.e., Ξ\Xi does not commute with tensoring by the line bundle 𝒪1(1)\mathcal{O}_{\mathbb{P}^{1}}(1).

let us now work out the details of the similar phenomenon in the real situtation, i.e., for 𝔤=𝔰𝔩3\mathfrak{g}=\mathfrak{sl}_{3}. Let ee be the subregular nilpotent. Then, the Springer fiber e\mathcal{B}_{e} has two irreducible components labeled by simple roots α1\alpha_{1}, α2\alpha_{2}. Each of these components is isomorphic to 1\mathbb{P}^{1} so we will denote the ii’th component by 1i\mathbb{P}^{1}_{i}. Components 11\mathbb{P}^{1}_{1}, 12\mathbb{P}^{1}_{2} intersect transversally at one point to be denoted pp. The set e\mathcal{B}_{e}^{\mathbb{C}^{*}} consists of three points, one of them is pp and two other q1,q2q_{1},q_{2} are such that qi1iq_{i}\in\mathbb{P}^{1}_{i}.

It is easy to see that the correspondence Γ:=iΓw¯i\Gamma:=\bigsqcup_{i}\Gamma_{\bar{w}_{i}} is equal to

(4.3.2) Γ=11×{q1}12×{q2}{p}×{p}(1112)×{p,q1,q2}\Gamma={\mathbb{P}}^{1}_{1}\times\{q_{1}\}\sqcup{\mathbb{P}}^{1}_{2}\times\{q_{2}\}\sqcup\{p\}\times\{p\}\subset(\mathbb{P}^{1}_{1}\cup\mathbb{P}^{1}_{2})\times\{p,q_{1},q_{2}\}

Let π~\widetilde{\pi} be the projection of the correspondence (4.3.2) onto e={p,q1.q2}\mathcal{B}_{e}^{\mathbb{C}^{*}}=\{p,q_{1}.q_{2}\} and let ι~\widetilde{\iota} be the projection of (4.3.2) onto e=1112\mathcal{B}_{e}=\mathbb{P}^{1}_{1}\cup\mathbb{P}^{1}_{2}.

Let λ\lambda be any dominant weight, let k1,k2k_{1},k_{2}\in\mathbb{Z} be such that 𝒪(λ)|1i=𝒪1i(ki)\mathcal{O}_{\mathcal{B}}(\lambda)|_{\mathbb{P}^{1}_{i}}=\mathcal{O}_{\mathbb{P}^{1}_{i}}(k_{i}). Take y=1y=1 (where yy is as in Warning 4.3.4.)

Hence, it follows from (4.3.1) that the homomorphism ϕ𝐜\phi^{\bf{c}} sends [Δ𝒪𝒩~(λ)][\Delta_{*}\mathcal{O}_{\widetilde{\mathcal{N}}(\lambda)}] to (for the first equality see Remark 4.3.3 above)

Ξ1([Δ𝒪𝒩~(λ)][𝒪iΓw¯i])=Ξ1([(𝒪e(λ)𝒪e)𝒪iΓw¯i])==Ξ1([𝒪11×{q1}(k1)𝒪12×{q2}(k2)𝒪{p}×{p}]).\Xi^{-1}([\Delta_{*}\mathcal{O}_{\widetilde{\mathcal{N}}(\lambda)}]*[\mathcal{O}_{\bigsqcup_{i}\Gamma_{\bar{w}_{i}}}])=\Xi^{-1}([(\mathcal{O}_{\mathcal{B}_{e}}(\lambda)\boxtimes\mathcal{O}_{\mathcal{B}_{e}^{\mathbb{C}^{*}}})\otimes\mathcal{O}_{\bigsqcup_{i}\Gamma_{\bar{w}_{i}}}])=\\ =\Xi^{-1}([\mathcal{O}_{\mathbb{P}^{1}_{1}\times\{q_{1}\}}(k_{1})\sqcup\mathcal{O}_{\mathbb{P}^{1}_{2}\times\{q_{2}\}}(k_{2})\sqcup\mathcal{O}_{\{p\}\times\{p\}}]).

So, our goal is to compare

Ξ(Δ[𝒪e(λ)])=Ξ(Δ[𝒪{q1,q2,p}])with[𝒪11×{q1}(k1)𝒪12×{q2}(k2)𝒪{p}×{p}].\Xi(\Delta_{*}[\mathcal{O}_{\mathcal{B}_{e}^{\mathbb{C}^{*}}}(\lambda)])=\Xi(\Delta_{*}[\mathcal{O}_{\{q_{1},q_{2},p\}}])~{}\text{with}~{}[\mathcal{O}_{\mathbb{P}^{1}_{1}\times\{q_{1}\}}(k_{1})\oplus\mathcal{O}_{\mathbb{P}^{1}_{2}\times\{q_{2}\}}(k_{2})\oplus\mathcal{O}_{\{p\}\times\{p\}}].

Recall that Ξ=(ι~×Ide)(π~×Ide)\Xi=(\widetilde{\iota}\times\operatorname{Id}_{\mathcal{B}_{e}^{\mathbb{C}^{*}}})_{*}(\widetilde{\pi}\times\operatorname{Id}_{\mathcal{B}_{e}^{\mathbb{C}^{*}}})^{*}, it follows that Ξ(Δ[𝒪{q1,q2,p}])\Xi(\Delta_{*}[\mathcal{O}_{\{q_{1},q_{2},p\}}]) is equal to

[𝒪Γ]=[𝒪11×{q1}12×{q2}{p×p}].[\mathcal{O}_{\Gamma}]=[\mathcal{O}_{\mathbb{P}^{1}_{1}\times\{q_{1}\}\sqcup\mathbb{P}^{1}_{2}\times\{q_{2}\}\sqcup\{p\times p\}}].

Clearly,

[𝒪11×{q1}12×{q2}{p×p}][𝒪11×{q1}(k1)𝒪12×{q2}(k2)𝒪{p}×{p}][\mathcal{O}_{\mathbb{P}^{1}_{1}\times\{q_{1}\}\sqcup\mathbb{P}^{1}_{2}\times\{q_{2}\}\sqcup\{p\times p\}}]\neq[\mathcal{O}_{\mathbb{P}^{1}_{1}\times\{q_{1}\}}(k_{1})\oplus\mathcal{O}_{\mathbb{P}^{1}_{2}\times\{q_{2}\}}(k_{2})\oplus\mathcal{O}_{\{p\}\times\{p\}}]

when λ0\lambda\neq 0 (i.e., k10k_{1}\neq 0 or k20k_{2}\neq 0.)

4.4. Case e=0e=0

4.4.1.

Let us consider the case e=0e=0 in more detail. We then have e=\mathcal{B}_{e}=\mathcal{B}, the action of \mathbb{C}^{*} on \mathcal{B} is trivial. It follows that the splitting Ξ\Xi is also trivial and the correspondence iΓw¯i\bigsqcup_{i}\Gamma_{\bar{w}_{i}} is equal to Δ\Delta_{\mathcal{B}} (the diagonal Δ×\Delta_{\mathcal{B}}\hookrightarrow\mathcal{B}\times\mathcal{B}). Let 𝐜0{\bf{c}}_{0} be the two-sided cell corresponding to e=0e=0. Theorem A in this case establishes the isomorphism:

KG×(×)J𝐜0[𝐯,𝐯1]K_{G^{\vee}\times\mathbb{C}^{*}}(\mathcal{B}\times\mathcal{B})\simeq J_{{\bf{c}}_{0}}\otimes\mathbb{Z}[{\bf{v}},{\bf{v}}^{-1}]

such that the homomorphism ϕ𝐜0:KG×(𝒩~×𝒩𝒩~)KG×(×)\phi^{{\bf{c}}_{0}}\colon K_{G^{\vee}\times\mathbb{C}^{*}}(\widetilde{\mathcal{N}}\times_{\mathcal{N}}\widetilde{\mathcal{N}})\rightarrow K_{G^{\vee}\times\mathbb{C}^{*}}(\mathcal{B}\times\mathcal{B}) is given by

xx[Δ],x\mapsto x*[\Delta_{\mathcal{B}}],

where * corresponds to the standard action of KG×(𝒩~×𝒩𝒩~)K_{G^{\vee}\times\mathbb{C}^{*}}(\widetilde{\mathcal{N}}\times_{\mathcal{N}}\widetilde{\mathcal{N}}) on KG×(×)K_{G^{\vee}\times\mathbb{C}^{*}}(\mathcal{B}\times\mathcal{B}) via convolution.

Remark 4.4.1.

The map xx[Δ]x\mapsto x*[\Delta_{\mathcal{B}}] can be explicitly described as follows. Let us consider 𝔑:=(𝒩~×𝒩𝒩~)×𝒩\mathfrak{N}:=\left(\widetilde{\mathcal{N}}\times_{\mathcal{N}}\widetilde{\mathcal{N}}\right)\times\mathcal{N} together with its three projections pip_{i} onto each of the factors. Let also π\pi be the natural map 𝔑𝒩~×𝒩𝒩~\mathfrak{N}\to\widetilde{\mathcal{N}}\times_{\mathcal{N}}\widetilde{\mathcal{N}}

Unwrapping the definitions, one gets x[Δ]=(p1×p3)(π(x)𝕃(π2×π3)𝒪Δ)x*[\Delta_{\mathcal{B}}]=(p_{1*}\times p_{3*})(\pi^{*}(x)\otimes^{\mathbb{L}}(\pi_{2}\times\pi_{3})^{*}\mathcal{O}_{\Delta}): here we have identified KG×(T×T)K_{G^{\vee}\times\mathbb{C}^{*}}(T^{*}\mathcal{B}\times T^{*}\mathcal{B}) and KG×(×)K_{G^{\vee}\times\mathbb{C}^{*}}(\mathcal{B}\times\mathcal{B}) by means of the Thom isomorphism τ\tau.

Let us also consider the natural map δ:𝒩~×𝒩𝒩~𝔑,(a,b)(a,b,b).\delta:\widetilde{\mathcal{N}}\times_{\mathcal{N}}\widetilde{\mathcal{N}}\to\mathfrak{N},\ (a,b)\mapsto(a,b,b). By the projection formula, x[Δ]=(p1×p3)δ(x)=j(x)x*[\Delta_{\mathcal{B}}]=(p_{1*}\times p_{3*})\delta_{*}(x)=j_{*}(x) for the natural map j:𝒩~×𝒩𝒩~𝒩~×𝒩~j:\widetilde{\mathcal{N}}\times_{\mathcal{N}}\widetilde{\mathcal{N}}\to\widetilde{\mathcal{N}}\times\widetilde{\mathcal{N}}.

The identification KG(×)J𝐜0K_{G^{\vee}}(\mathcal{B}\times\mathcal{B})\simeq J_{{\bf{c}}_{0}} was already obtained in [59] (when the derived subgroup of GG^{\vee} is simply-connected) and in [49] (in general). Moreover, in [63] the homomorphism ϕ𝐜0:KG(×)KG(×)J𝐜0\phi^{{\bf{c}}_{0}}\colon K_{G^{\vee}}(\mathcal{B}\times\mathcal{B})\rightarrow K_{G^{\vee}}(\mathcal{B}\times\mathcal{B})\simeq J_{{\bf{c}}_{0}} was described for GG such that the derived subgroup of GG^{\vee} is simply-connected. For such GG^{\vee}, KG(×)K_{G^{\vee}}(\mathcal{B}\times\mathcal{B}) is isomorphic to the matrix algebra KG()KG(pt)KG()K_{G^{\vee}}(\mathcal{B})\otimes_{K_{G^{\vee}}(\operatorname{pt})}K_{G^{\vee}}(\mathcal{B}) (see [36, Proposition 1.6]), this identification is used in construction of ϕ𝐜0\phi^{{\bf{c}}_{0}} given in [63]).

The geometric construction in loc. cit. is equivalent to the one from the Remark above (cf. Theorem 3.5 in loc. cit.) up to the conjugation by some explicit matrix.

4.4.2. Example when a specialization of 𝐉𝐜0{\bf{J}}_{{\bf{c}}_{0}} at sSpec𝐊Ze(pt)s\in\operatorname{Spec}{\bf{K}}_{Z_{e}}(\operatorname{pt}) is not semisimple

It is known that 𝐉𝐜0{\bf{J}}_{{\bf{c}}_{0}} is not isomorphic to a matrix algebra over 𝐊Ze(pt){\bf{K}}_{Z_{e}}(\operatorname{pt}) in general (see [60, Section 8.3]). Actually, it is even not true that the fiber of 𝐉𝐜0{\bf{J}}_{{\bf{c}}_{0}} at every point of Spec𝐊Ze(pt)\operatorname{Spec}{\bf{K}}_{Z_{e}}(\operatorname{pt}) is semisimple (c.f. [12, Corollary 1] and [55, Example 4.4]).

For example, for G=SL2G=\operatorname{SL}_{2}, and s=[diag(1,1)]Spec𝐊PGL2(pt)s=[\operatorname{diag}(1,-1)]\in\operatorname{Spec}{\bf{K}}_{\operatorname{PGL}_{2}}(\operatorname{pt}), the fiber of 𝐉𝐜0𝐊PGL2(1×1){\bf{J}}_{{\bf{c}}_{0}}\simeq{\bf{K}}_{\operatorname{PGL}_{2}}(\mathbb{P}^{1}\times\mathbb{P}^{1}) at ss is four-dimensional but the are only two irreducible representations of 𝐊PGL2(1×1)s{\bf{K}}_{\operatorname{PGL}_{2}}(\mathbb{P}^{1}\times\mathbb{P}^{1})_{s} and both of them are one-dimensional. To see this, note that by Theorem C (see Section 5.1 and Proposition A.2.1 below), irreducible modules over 𝐊PGL2(1×1)s{\bf{K}}_{\operatorname{PGL}_{2}}(\mathbb{P}^{1}\times\mathbb{P}^{1})_{s} are all of the form K((1)s)ρK((\mathbb{P}^{1})^{s})_{\rho}, where ρ\rho is an irreducible representation of ZPGL2(s)/ZPGL2(s)0=/2Z_{\operatorname{PGL}_{2}}(s)/Z_{\operatorname{PGL}_{2}}(s)^{0}=\mathbb{Z}/2\mathbb{Z} acting simply transitively on (1)s={0,}(\mathbb{P}^{1})^{s}=\{0,\infty\}. So, K((1)s)K((\mathbb{P}^{1})^{s}) is two dimensional and is the direct sum of two irreducible one-dimensional 𝐊PGL2(1×1)s{\bf{K}}_{\operatorname{PGL}_{2}}(\mathbb{P}^{1}\times\mathbb{P}^{1})_{s}-modules.

5. Representation theory of JeJ_{e} from the geometric perspective.

5.1. Simple modules over 𝐉e{\bf{J}}_{e}

The goal of this Section is classification of simple 𝐉e{\bf{J}}_{e}-modules resulting in the proof of Theorem C.

We first recall Lusztig’s classification [40, Theorem 4.2]. It is shown in loc. cit. that for every pair (s,ρ)(s,\rho) of a semisimple element sZes\in Z_{e} and ρIrrep(ZZe(s)/ZZe(s)0)\rho\in\operatorname{Irrep}(Z_{Z_{e}}(s)/{Z_{Z_{e}}(s)^{0}}) there exists unique irreducible 𝐉e{\bf{J}}_{e}-module E(s,e,ρ)E(s,e,\rho) characterized by the following property: ϕeqE(s,e,ρ)\phi^{e*}_{q}E(s,e,\rho) is isomorphic to K(e,s,ρ,q):=𝐊(sqe)ρK(e,s,\rho,q):={\bf{K}}(\mathcal{B}^{sq}_{e})_{\rho} for generic qq. Moreover, Lusztig proved that every irreducible 𝐉e{\bf{J}}_{e}-module is of the form E(s,e,ρ)E(s,e,\rho) for some (s,ρ)(s,\rho) as above.

Our goal is to prove that:

(a) irreducible modules over 𝐉e=𝐊Ze(e×e){\bf{J}}_{e}={\bf{K}}_{Z_{e}}(\mathcal{B}_{e}^{{\mathbb{C}}^{*}}\times\mathcal{B}_{e}^{{\mathbb{C}}^{*}}) are of the form 𝐊(e,s)ρ{\bf{K}}(\mathcal{B}_{e}^{{\mathbb{C}}^{*},s})_{\rho},

(b) we have: E(s,e,ρ)=𝐊(e,s)ρE(s,e,\rho)={\bf{K}}(\mathcal{B}_{e}^{{\mathbb{C}}^{*},s})_{\rho}.

Part (a)(a) of this theorem immediately follows from Proposition A.2.1. Part (b)(b) is a consequence of the next Lemma. Set Γ=Γse:=ZZe(s)/ZZe(s)0\Gamma=\Gamma^{s}_{e}:=Z_{Z_{e}}(s)/{Z_{Z_{e}}(s)^{0}}.

Lemma 5.1.1.

There are canonical isomorphisms of eΓse\mathcal{H}_{e}-\Gamma^{s}_{e}- and Γse\mathcal{H}-\Gamma^{s}_{e}-modules respectively:

ϕe,q𝐊(e,s)𝐊(esq),ϕqe𝐊(e,s)𝐊(esq).\phi_{e,q}^{*}{\bf{K}}(\mathcal{B}_{e}^{\mathbb{C}^{*},s})\simeq{\bf{K}}(\mathcal{B}_{e}^{sq}),\,\phi_{q}^{e*}{\bf{K}}(\mathcal{B}_{e}^{\mathbb{C}^{*},s})\simeq{\bf{K}}(\mathcal{B}_{e}^{sq}).
Proof.

We prove the claim for ϕeq\phi^{e*}_{q}, the argument for ϕe,q\phi_{e,q}^{*} follows.

We have checked (see Remark 4.2.3) that the Je[𝐯,𝐯1]\mathcal{H}-J_{e}\otimes{\mathbb{C}}[{\bf{v}},{\bf{v}}^{-1}]-bimodule corresponding to the homomorphism ϕe\phi^{e} is identified with e\mathcal{F}_{e}, thus we have a canonical isomorphism:

ϕeMeKeM\phi^{e*}M\simeq\mathcal{\mathcal{F}}_{e}\otimes_{K_{e}}M

holding for any Je[𝐯,𝐯1]J_{e}\otimes\mathbb{C}[{\bf{v}},{\bf{v}}^{-1}]-module MM.

Thus we are reduced to constructing an isomorphism:

(5.1.1) 𝐊Ze×(e×e)q𝐊e𝐊(e,s)𝐊(esq).{\bf{K}}_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e}^{\mathbb{C}^{*}})_{q}\otimes_{{\bf{K}}_{e}}{\bf{K}}(\mathcal{B}_{e}^{\mathbb{C}^{*},s})\simeq_{{\mathbfcal{H}}}{\bf{K}}(\mathcal{B}_{e}^{sq}).

Let us denote by CC the product of our \mathbb{C}^{*} and s\langle s\rangle (Zariski closure of the cyclic subgroup generated by ss). Setting 𝐊es:=𝐊C(e×e)Γ{\bf{K}}_{e}^{s}:={\bf{K}}_{C}(\mathcal{B}_{e}^{\mathbb{C}^{*}}\times\mathcal{B}_{e}^{\mathbb{C}^{*}})^{\Gamma}, we have an isomorphism of {\mathbfcal{H}}-modules:

(5.1.2) 𝐊Ze×(e×e)𝐊e𝐊(e,s)q𝐊C(e×e)Γ𝐊es𝐊(e,s)q,{\bf{K}}_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e}^{\mathbb{C}^{*}})\otimes_{{\bf{K}}_{e}}{\bf{K}}(\mathcal{B}_{e}^{\mathbb{C}^{*},s})_{q}\simeq_{{\mathbfcal{H}}}{\bf{K}}_{C}(\mathcal{B}_{e}\times\mathcal{B}_{e}^{\mathbb{C}^{*}})^{\Gamma}\otimes_{{\bf{K}}_{e}^{s}}{\bf{K}}(\mathcal{B}_{e}^{\mathbb{C}^{*},s})_{q},

this follows from the fact that 𝐊Ze×(e×e){\bf{K}}_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e}^{\mathbb{C}^{*}}) (respectively, 𝐊C(e×e)Γ{\bf{K}}_{C}(\mathcal{B}_{e}\times\mathcal{B}_{e}^{\mathbb{C}^{*}})^{\Gamma}) is a free rank one module over 𝐊e{\bf{K}}_{e} (respectively, 𝐊es{\bf{K}}_{e}^{s}), both sides of (5.1.2) are isomorphic to 𝐊(e,s)q{\bf{K}}(\mathcal{B}_{e}^{\mathbb{C}^{*},s})_{q}.

Consider

Me,s,q:=𝐊C(e×e)Γq𝐊es𝐊(e,s).{M}_{e,s,q}:={\bf{K}}_{C}(\mathcal{B}_{e}\times\mathcal{B}_{e}^{\mathbb{C}^{*}})^{\Gamma}_{q}\otimes_{{\bf{K}}_{e}^{s}}{\bf{K}}(\mathcal{B}_{e}^{\mathbb{C}^{*},s}).

Here the central subalgebra 𝐊C(pt)𝐊es{\bf{K}}_{C}(\operatorname{pt})\subset{\bf{K}}_{e}^{s} acts on the right factor via the character corresponding to sqsq, we will denote reduction by that character by the subscript sq\bullet_{sq}. By Section A.1.5 we get (cf. also (5.2.2) below):

(5.1.3) 𝐊C(e×e)Γsq=𝐊(esq×e,s)Γ=ρIrrep(Γ)𝐊(esq)ρ𝐊(e,s)ρ,{\bf{K}}_{C}(\mathcal{B}_{e}\times\mathcal{B}_{e}^{\mathbb{C}^{*}})^{\Gamma}_{sq}={\bf{K}}(\mathcal{B}_{e}^{sq}\times\mathcal{B}_{e}^{\mathbb{C}^{*},s})^{\Gamma}=\bigoplus_{\rho\in\operatorname{Irrep}(\Gamma)}{\bf{K}}(\mathcal{B}_{e}^{sq})_{\rho}\otimes{\bf{K}}(\mathcal{B}_{e}^{\mathbb{C}^{*},s})_{\rho^{*}},

where in the last equality we use [20, Theorem 5.6.1] (the implication (b)(a)(b)\Rightarrow(a), for the equivariance under the trivial group).

Strictly speaking, to use it, we should establish the Kunneth formula for the smooth variety e,s\mathcal{B}_{e}^{\mathbb{C}^{*},s}. However, it immediately follows from the second of the following two identifications:

(5.1.4) K(esq×esq)H(esq×esq),K(e,s×e,s)H(e,s×e,s).K(\mathcal{B}_{e}^{sq}\times\mathcal{B}_{e}^{sq})\simeq H_{*}(\mathcal{B}_{e}^{sq}\times\mathcal{B}_{e}^{sq}),\ K(\mathcal{B}_{e}^{{\mathbb{C}}^{*},s}\times\mathcal{B}_{e}^{{\mathbb{C}}^{*},s})\simeq H_{*}(\mathcal{B}_{e}^{{\mathbb{C}}^{*},s}\times\mathcal{B}_{e}^{{\mathbb{C}}^{*},s}).

For the first of these two equations, note that e×e\mathcal{B}_{e}\times\mathcal{B}_{e} is also a Springer fiber (for G×GG^{\vee}\times G^{\vee}). Thus, H(esq×esq)H_{*}(\mathcal{B}_{e}^{sq}\times\mathcal{B}_{e}^{sq}) is isomorphic to the Chow group A(esq×esq)A_{*}(\mathcal{B}_{e}^{sq}\times\mathcal{B}_{e}^{sq}), cf. [26, Theorem 3.9]. Moreover, the Chow group A(esq×esq)A_{*}(\mathcal{B}_{e}^{sq}\times\mathcal{B}_{e}^{sq}) is isomorphic to the KK-theory K(esq×esq)K(\mathcal{B}_{e}^{sq}\times\mathcal{B}_{e}^{sq}) by [16, Theorem III.1 (b)]).

The second identification also follows from [26] and [16] in a similar fashion.

Now, the results of section A.1.5 also imply that

(5.1.5) (𝐊es)sq=ρIrrep(Γ)End𝐊(es)ρ.({\bf{K}}_{e}^{s})_{sq}=\bigoplus_{\rho\in\operatorname{Irrep}(\Gamma)}\operatorname{End}{\bf{K}}(\mathcal{B}_{es}^{\mathbb{C}^{*}})_{\rho}.

Now (5.1.5) acts on (5.1.3) via the action on the second tensor factor.

Thus

Me,s,q=ρIrrep(Γ)𝐊(esq)ρ(𝐊(e,s)ρEnd𝐊(e,s)ρ𝐊(e,s))=M_{e,s,q}=\bigoplus_{\rho\in\operatorname{Irrep}(\Gamma)}{\bf{K}}(\mathcal{B}_{e}^{sq})_{\rho}\otimes({\bf{K}}(\mathcal{B}_{e}^{\mathbb{C}^{*},s})_{\rho^{*}}\otimes_{\operatorname{End}{\bf{K}}(\mathcal{B}_{e}^{\mathbb{C}^{*},s})_{\rho}}{\bf{K}}(\mathcal{B}_{e}^{\mathbb{C}^{*},s}))=
=ρIrrep(Γ)𝐊(esq)ρ(𝐊(e,s)ρEnd𝐊(e,s)ρ(𝐊(e,s)ρρ))=ρIrrep(Γ)𝐊(esq)ρρ.=\bigoplus_{\rho\in\operatorname{Irrep}(\Gamma)}{\bf{K}}(\mathcal{B}_{e}^{sq})_{\rho}\otimes({\bf{K}}(\mathcal{B}_{e}^{\mathbb{C}^{*},s})_{\rho^{*}}\otimes_{\operatorname{End}{\bf{K}}(\mathcal{B}_{e}^{\mathbb{C}^{*},s})_{\rho}}({\bf{K}}(\mathcal{B}_{e}^{\mathbb{C}^{*},s})_{\rho}\otimes\rho))=\bigoplus_{\rho\in\operatorname{Irrep}(\Gamma)}{\bf{K}}(\mathcal{B}_{e}^{sq})_{\rho}\otimes\rho.

5.2. Families of modules over JeJ_{e} and e\mathcal{H}_{e}

In the previous section we proved that for any qq\in{\mathbb{C}}^{*} there exists an identification ϕqeE(e,s,ρ)=K(e,s,ρ,q)\phi_{q}^{e*}E(e,s,\rho)=K(e,s,\rho,q).

(It should be noted that for qq being not a root of the Poincaré polynomial of WfW_{f}, this was already shown in [19] using the algebraic result from [61].)

Now, we will explain that this collection of isomorphisms fits into an algebraic family, using our geometric results on JJ.

This fact is stated in [19] and used in the proof of [19, Theorem 1.8 (3)], cf. footnote 3 and section 6 below.

5.2.1.

We start by introducing certain algebraic family whose fibers are of the form K(e,s,q):=ρIrrepΓseρK(e,s,ρ,q).K(e,s,q):=\bigoplus_{\rho\in\operatorname{Irrep}\Gamma^{s}_{e}}\rho\otimes K(e,s,\rho,q).

If ZeZ_{e} is connected modulo the center of GG^{\vee} and has simply connected derived subgroup, then the KK-theory 𝐊Ze×(e){\bf{K}}_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}) yields a family which has such fibers for all ss: the specialization at the maximal ideal of a semisimple conjugacy class γsq\gamma_{sq} of some sqsq, is

𝐊Ze×(e)γsq𝐊(esq)=K(e,s,q)=ρIrrepΓseρK(e,s,ρ,q),{\bf{K}}_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e})_{\gamma_{sq}}\simeq{\bf{K}}(\mathcal{B}_{e}^{sq})=K(e,s,q)=\bigoplus_{\rho\in\operatorname{Irrep}\Gamma^{s}_{e}}\rho\otimes K(e,s,\rho,q),

where the first isomorphism follows from localization theorem [28] (see Appendix A).

In general, the situation is more complicated.

Remark 5.2.1.

Let Γ\Gamma be a reductive group which is either disconnected or not simply connected. For a semisimple conjugacy class γg\gamma_{g} of some gΓg\in\Gamma the specialization 𝐊Γ(X)γg{\bf{K}}_{\Gamma}(X)_{\gamma_{g}} may be not isomorphic to 𝐊(Xg){\bf{K}}(X^{g}). In particular, the finite group ZΓ(g)/ZΓ(g)0Z_{\Gamma}(g)/Z_{\Gamma}(g)^{0} acts trivially on the former, while it may act non-trivially on the latter (cf. [12, 5.2]).

Nor do we have an isomorphism666However, we have (cf. Appendix A.1): (5.2.1) 𝐊Γ(X)γg𝐊ZΓ(g)(Xg)g.{\bf{K}}_{\Gamma}(X)_{\gamma_{g}}\simeq{\bf{K}}_{Z_{\Gamma}(g)}(X^{g})_{g}. between 𝐊Γ(X)γg{\bf{K}}_{\Gamma}(X)_{\gamma_{g}} and the invariants 𝐊(Xg)ZΓ(g)ZΓ(g)0{\bf{K}}(X^{g})^{\frac{Z_{\Gamma}(g)}{Z_{\Gamma}(g)^{0}}}.

A counterexample is provided by PGL2\operatorname{PGL}_{2} acting on X=1X=\mathbb{P}^{1}. Let ss be the element [diag(1,1)]PGL2[\operatorname{diag}(1,-1)]\in\operatorname{PGL}_{2}, let TT be the diagonal torus and Z=ZPGL2(s)Z=Z_{\operatorname{PGL}_{2}}(s). Then, Xs=Z/TX^{s}=Z/T is a set of two points. Thus, 𝐊Z(Xs)=𝐊T(pt)=Λ{\bf{K}}_{Z}(X^{s})={\bf{K}}_{T}(\operatorname{pt})=\mathbb{C}\Lambda, where Λ\Lambda is a character lattice. It is a module of rank 22 over 𝐊Z(pt){\bf{K}}_{Z}(\operatorname{pt}), so 𝐊Z(Xs)s{\bf{K}}_{Z}(X^{s})_{s} has dimension (at least) two. On the other hand, 𝐊(Xs){\bf{K}}(X^{s}) is the one-dimensional space of /2\mathbb{Z}/2\mathbb{Z}-invariants.

5.2.2.

Being unable to define a single family with required fibers, we instead consider (following [19]) a collection of families defined for every semisimple sZes\in Z_{e}.

We fix such ss, and a torus CZZe(s)C\subset Z_{Z_{e}}(s), and proceed to construct a family of modules over Spec(𝐊C×(pt))\operatorname{Spec}({\bf{K}}_{C\times\mathbb{C}^{*}}(\operatorname{pt})) whose specialization to a point (χ,q)C×(\chi,q)\in C\times\mathbb{C}^{*} is isomorphic to 𝐊(esχq)=H(esχq,){\bf{K}}(\mathcal{B}_{e}^{s\chi q})=H_{*}(\mathcal{B}_{e}^{s\chi q},\mathbb{C}). (The latter equality holds since the homology group H(esχq,)H_{*}(\mathcal{B}_{e}^{s\chi q},\mathbb{Z}) is isomorphic to the Chow group A(esχq)A_{*}(\mathcal{B}_{e}^{s\chi q}); cf. [26, Theorem 3.9] and the Chow group A(esχq)A_{*}(\mathcal{B}_{e}^{s\chi q}) is isomorphic to the KK-theory K(esχq)K(\mathcal{B}_{e}^{s\chi q}) by [16, Theorem III.1 (b)]).

Let us denote the diagonalizable group C,s\langle C,s\rangle by HH. We set 𝒦(C,s):=𝐊H×(e)|Cs×\mathcal{K}(C,s):={\bf{K}}_{H\times\mathbb{C}^{*}}(\mathcal{B}_{e})|_{Cs\times\mathbb{C}^{*}}, where the subscript refers to restriction to the closed subset of Spec(𝐊H×(pt))\operatorname{Spec}({\bf{K}}_{H\times\mathbb{C}^{*}}(\operatorname{pt})).

From Section A.1.5 it follows that for any χ\chi as above we have:

(5.2.2) 𝒦(C,s)^χq𝐊H×(e)^sχq𝐊H×(esχq)^1𝐊H0×(esχq)^1,\widehat{\mathcal{K}(C,s)}^{\chi q}\simeq\widehat{{\bf{K}}_{H\times\mathbb{C}^{*}}(\mathcal{B}_{e})}^{s\chi q}\simeq\widehat{{\bf{K}}_{H\times\mathbb{C}^{*}}(\mathcal{B}_{e}^{s\chi q})}^{1}\simeq\widehat{{\bf{K}}_{H^{0}\times\mathbb{C}^{*}}(\mathcal{B}_{e}^{s\chi q})}^{1},

where by ^t\widehat{\ }\ ^{t} we mean completion at the maximal ideal of an element tt and H0H^{0} denotes the identity component in HH.

Applying localization theorem to the torus H0×H^{0}\times{\mathbb{C}}^{*} we conclude that the specialization 𝒦(C,s)χq\mathcal{K}(C,s)_{\chi q} is identified with 𝐊(esχq){\bf{K}}(\mathcal{B}_{e}^{s\chi q}).

The ring =𝐊Ze×(e×e){\mathbfcal{H}}={\bf{K}}_{Z_{e}\times\mathbb{C}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e}) acts naturally on 𝒦(C,s)\mathcal{K}(C,s), it is easy to see that the action on the fiber coincides with one introduced in Proposition A.2.1 (cf. also [20, 5.11.7, 5.11.10]).

Remark 5.2.2.

The module 𝒦(C,s)\mathcal{K}(C,s) is similar to the semiperiodic module of [13]; it can be viewed as its generalization for sZG(T0)s\notin Z_{G}(T_{0}), in the notation of loc. cit..

In a similar manner, one can form a family (C,s):=𝐊H×(e)|Cs×\mathcal{L}(C,s):={\bf{K}}_{H\times{\mathbb{C}}^{*}}(\mathcal{B}_{e}^{\mathbb{C}^{*}})|_{Cs\times{\mathbb{C}}^{*}} over Spec(𝐊C×(pt))\operatorname{Spec}({\bf{K}}_{C\times\mathbb{C}^{*}}(\operatorname{pt})) with fibers 𝐊(e,sχ){\bf{K}}(\mathcal{B}_{e}^{\mathbb{C}^{*},s\chi}) and equip it with an action of Je[𝐯,𝐯1]=𝐊eJ_{e}\otimes{\mathbb{C}}[{\bf{v}},{\bf{v}}^{-1}]={\bf{K}}_{e}. We set (C,s,q):=(C,s)|Cs×{q}\mathcal{L}(C,s,q):=\mathcal{L}(C,s)|_{Cs\times\{q\}}.

5.3.

Now, the natural goal is to obtain a ‘‘version in families’’ of Lemma 5.1.1 above, i.e. to prove that the family 𝒦(C,s)\mathcal{K}(C,s) is isomorphic to the pullback of the family (C,s)\mathcal{L}(C,s) under the homomorphism ϕe\phi_{e}.

We need the following auxiliary result.

Let us consider a natural morphism,

κ:𝐊Cs×(e)𝒪(C×)𝐊Cs×(e)𝐊Cs×(e×e).\kappa\colon{\bf{K}}_{Cs\times\mathbb{C}^{*}}(\mathcal{B}_{e}^{\mathbb{C}^{*}})\otimes_{\mathcal{O}(C\times\mathbb{C}^{*})}{\bf{K}}_{Cs\times\mathbb{C}^{*}}(\mathcal{B}_{e}^{\mathbb{C}^{*}})\to{\bf{K}}_{Cs\times\mathbb{C}^{*}}(\mathcal{B}_{e}^{\mathbb{C}^{*}}\times\mathcal{B}_{e}^{\mathbb{C}^{*}}).
Lemma 5.3.1.

κ\kappa is an isomorphism.

Proof.

It follows from [44, Theorem 1.14] (see also [26]) that 𝐊Cs×(e){\bf{K}}_{Cs\times\mathbb{C}^{*}}(\mathcal{B}_{e}) is a free module over 𝐊Cs×(pt)=𝒪(C×){\bf{K}}_{Cs\times\mathbb{C}^{*}}(\operatorname{pt})=\mathcal{O}(C\times\mathbb{C}^{*}).

Thus, it suffices to check the claim fiberwise. Similarly to calculation (5.2.2) it follows from:

K(e,sχ×e,sχ)K(e,sχ)K(e,sχ).K(\mathcal{B}_{e}^{\mathbb{C}^{*},s\chi}\times\mathcal{B}_{e}^{\mathbb{C}^{*},s\chi})\simeq K(\mathcal{B}_{e}^{\mathbb{C}^{*},s\chi})\otimes K(\mathcal{B}_{e}^{\mathbb{C}^{*},s\chi}).

The last isomorphism is clear from the argument similar to the formulas (5.1.4). ∎

Set M:=ZG(C)0M^{\vee}:=Z_{G^{\vee}}(C)^{0}, ΓM:=π0(ZM(e,s))\Gamma_{M}:=\pi_{0}(Z_{M^{\vee}}(e,s)).

Let us denote the algebra 𝐊Cs×(e×e){\bf{K}}_{Cs\times\mathbb{C}^{*}}(\mathcal{B}_{e}^{\mathbb{C}^{*}}\times\mathcal{B}_{e}^{\mathbb{C}^{*}}) by 𝐊(M){\bf{K}}(M).

Similarly to the proof of Lemma 5.1.1, one obtains a morphism

e𝐊e𝐊Cs×(e)𝐊Cs×(e×e)ΓM𝐊(M)ΓM𝐊Cs×(e).{\mathbfcal{H}}_{e}\otimes_{{\bf{K}}_{e}}{\bf{K}}_{Cs\times\mathbb{C}^{*}}(\mathcal{B}_{e}^{\mathbb{C}^{*}})\to{\bf{K}}_{Cs\times\mathbb{C}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e}^{\mathbb{C}^{*}})^{\Gamma_{M}}\otimes_{{\bf{K}}(M)^{\Gamma_{M}}}{\bf{K}}_{Cs\times\mathbb{C}^{*}}(\mathcal{B}_{e}^{\mathbb{C}^{*}}).

It is an isomorphism since it is an isomorphism on fibers and all involved modules are flat.

Now we claim that, moreover, 𝐊Cs×(e×e)ΓM𝐊(M)ΓM𝐊Cs×(e)𝐊Cs×(e){\bf{K}}_{Cs\times\mathbb{C}^{*}}(\mathcal{B}_{e}\times\mathcal{B}_{e}^{\mathbb{C}^{*}})^{\Gamma_{M}}\otimes_{{\bf{K}}(M)^{\Gamma_{M}}}{\bf{K}}_{Cs\times\mathbb{C}^{*}}(\mathcal{B}_{e}^{\mathbb{C}^{*}})\simeq{\bf{K}}_{Cs\times\mathbb{C}^{*}}(\mathcal{B}_{e}).

Indeed, we observe that

ρIrrepΓM𝐊Cs×(e)ρ𝒪(Cs×)𝐊Cs×(e)ρEnd𝒪(Cs×)𝐊Cs×(e)ρ=:𝒜,\bigoplus_{\rho\in\operatorname{Irrep}\Gamma_{M}}{\bf{K}}_{Cs\times\mathbb{C}^{*}}(\mathcal{B}_{e}^{\mathbb{C}^{*}})_{\rho}\otimes_{\mathcal{O}(Cs\times\mathbb{C}^{*})}{\bf{K}}_{Cs\times\mathbb{C}^{*}}(\mathcal{B}_{e}^{\mathbb{C}^{*}})_{\rho^{*}}\simeq\operatorname{End}_{\mathcal{O}(Cs\times\mathbb{C}^{*})}{\bf{K}}_{Cs\times\mathbb{C}^{*}}(\mathcal{B}_{e}^{\mathbb{C}^{*}})_{\rho}=:\mathcal{A},

which follows fiberwise from the argument similar to the equations (5.1.4).

Moreover, since the vector bundle (over the corresponding torus) 𝐊Cs×(e){\bf{K}}_{Cs\times\mathbb{C}^{*}}(\mathcal{B}_{e}^{\mathbb{C}^{*}}) is self-dual in a canonical way (thanks to Poincaré pairing), there exists the canonical morphism

ρ𝒪(C)𝐊Cs×(e)ρ𝒜𝐊Cs×(e).\rho^{*}\otimes\mathcal{O}(C)\to{\bf{K}}_{Cs\times\mathbb{C}^{*}}(\mathcal{B}_{e}^{\mathbb{C}^{*}})_{\rho}\otimes_{\mathcal{A}}{\bf{K}}_{Cs\times\mathbb{C}^{*}}(\mathcal{B}_{e}^{\mathbb{C}^{*}}).

One easily sees that, in fact, it establishes the isomorphism. Indeed, it is enough to check this fiberwise, and we reduce the statement to an elementary linear algebra observation.

We conclude:

𝐊Cs×(e)ρ𝒜𝐊Cs×(e)ρ𝒪(C).{\bf{K}}_{Cs\times\mathbb{C}^{*}}(\mathcal{B}_{e}^{\mathbb{C}^{*}})_{\rho}\otimes_{\mathcal{A}}{\bf{K}}_{Cs\times\mathbb{C}^{*}}(\mathcal{B}_{e}^{\mathbb{C}^{*}})\simeq\rho^{*}\otimes\mathcal{O}(C).

Now, as in Lemma 5.1.1, we obtain:

Theorem 5.3.2.

As eΓM{\mathbfcal{H}}_{e}-\Gamma_{M}-modules, ϕe(C,s)=𝐊Cs×(e).\phi_{e}^{*}\mathcal{L}(C,s)={\bf{K}}_{Cs\times\mathbb{C}^{*}}(\mathcal{B}_{e}).

Now the arguments in the previous chapter generalize verbatim to prove the following ΓM\Gamma_{M}-equivariant isomorphism.

Theorem 5.3.3.
(5.3.1) ϕe(C,s)𝐊Cs×(e).\phi^{e*}\mathcal{L}(C,s)\simeq_{{\mathbfcal{H}}}{\bf{K}}_{Cs\times\mathbb{C}^{*}}(\mathcal{B}_{e}).
Remark 5.3.4.

This proves the version of Corollary 2.6 in [19] ‘‘in families’’.

So, Theorem D is proven.

6. Braverman-Kazhdan’s spectral description of JeJ_{e}

6.1.

We are now ready to reprove Theorem 1.8 (3) in [19] addressing the points in footnotes 3, 4 using our geometric approach to 𝐉e{\bf{J}}_{e}.

Let 𝐆\bf G be a version of GG over a local field; let qq be a characteristic of a residue field; we assume that qq is large enough. (Our convention for bold letters is slightly different from the one in [19]).

Let 𝐌\bf M be a Levi subgroup of 𝐆\bf G. Let σ\sigma be an irreducible tempered representation of 𝐌\bf M and let χ\chi be a character of 𝐌\bf M.

Then (cf. [4, p.78]) all representations of the form Ind𝐆𝐌(σχ)\operatorname{Ind}^{\bf G}_{\bf M}(\sigma\otimes\chi) can be realized on the same vector space that we denote VσV_{\sigma}.

Let ZMZ_{M} be the connected component of 1Z(M)1\in Z(M^{\vee}) (recall that Z(M)Z(M^{\vee}) is the center of MM^{\vee}). Then we obtain an action of 𝐆\bf G on the trivial vector bundle over a torus ZMZ_{M}.

By taking Iwahori-invariants, we obtain a family of {\mathbfcal{H}}-modules over ZMZ_{M} (cf. [19, 1.2]): let us denote it by 𝒱(ZM,σ).\mathcal{V}(Z_{M},\sigma). In particular, setting σ=K(s,e,ρ,q)\sigma=K(s,e,\rho,q) we get a family that we denote by 𝒱(ZM,s,ρ,q)\mathcal{V}(Z_{M},s,\rho,q) (since ee is fixed it does not enter the notation).

6.2.

It is known that:
1) for any LL^{\vee} containing MM^{\vee}, the family 𝒱(ZL,s,ρ,q)\mathcal{V}(Z_{L},s,\rho,q) can be naturaly realized as a subfamily 𝒱(ZM,s,ρM,q)\mathcal{V}(Z_{M},s,\rho_{M},q), where ρM\rho_{M} is defined as pull back of ρ\rho under π0(ZM(e,s))π0(ZL(e,s))\pi_{0}(Z_{M^{\vee}}(e,s))\to\pi_{0}(Z_{L^{\vee}}(e,s));
2) for any compact ss^{\prime} and a triple (L,t,θ)(L^{\vee},t,\theta) conjugate to (M,s,ρ)(M^{\vee},s,\rho), the family 𝒱(ZM,s,ρ,q)\mathcal{V}(Z_{M},s,\rho,q) can be rationally identified with 𝒱(ZL,t,θ,q)\mathcal{V}(Z_{L},t,\theta,q) via the intertwining operator.

The next statement is an equivalent form of Theorem 1.8 (3) in [19].

Theorem 6.2.1.

Let Π=M,s,ρEndrat.𝒪(ZM)𝒱(ZM,s,ρ,q)\Pi=\prod_{M,s,\rho}\operatorname{End}^{\mathrm{rat}.}_{\mathcal{O}(Z_{M})}\mathcal{V}(Z_{M},s,\rho,q), where the product is taken over all triples M,s,ρM,\,s,\rho as above with ss compact. Let 𝒮eΠ\mathcal{S}_{e}\subset\Pi be the subalgebra consisting of elements ϕ\phi which satisfy the following conditions:
a) φ\varphi does not have poles at points of families 𝒱(ZM,s,ρ,q)\mathcal{V}(Z_{M},s,\rho,q) corresponding to non-strictly positive characters of Levi subgroups;
b) φ\varphi is compatible with 1) and 2) above.

Then 𝒮e𝐉e\mathcal{S}_{e}\simeq{\bf{J}}_{e}.

The goal of this section is to summarize the proof as an application of our geometric description of 𝐉e{\bf{J}}_{e}.

6.3.

Let e,M\mathcal{B}_{e,M^{\vee}} be the Springer fiber in MM^{\vee}. Then there is an embedding of the lowest (in the natural ‘‘Bialynicki-Birula’’ order) component of ZMZ_{M}-fixed points: i:e,Mei\colon\mathcal{B}_{e,M^{\vee}}\to\mathcal{B}_{e}. We have a natural morphism

(6.3.1) i:𝐊ZM,s×(e,M)q|ZMs×𝐊ZM,s×(e)q|ZMs×.i_{*}\colon{\bf{K}}_{\langle Z_{M},s\rangle\times\mathbb{C}^{*}}(\mathcal{B}_{e,M})_{q}|_{Z_{M}s\times{\mathbb{C}}^{*}}\to{\bf{K}}_{\langle Z_{M},s\rangle\times\mathbb{C}^{*}}(\mathcal{B}_{e})_{q}|_{Z_{M}s\times{\mathbb{C}}^{*}}.

Note that the source of this map is a trivial family of M{\mathbfcal{H}}_{M}-modules, the ρ\rho-multiplicity subspace for the action of π0(ZM(e,s))\pi_{0}(Z_{M^{\vee}}(e,s)) on its fiber equals σ\sigma.

6.4.

For a finite group Γ\Gamma with an irreducible representation ρ\rho and any representation VV, we will denote the multiplicity space HomΓ(ρ,V)\operatorname{Hom}_{\Gamma}(\rho,V) by Vρ.V_{\rho}. We proceed to compare the geometrically defined family 𝒦(ZM,s,q)ρ\mathcal{K}(Z_{M},s,q)_{\rho} with 𝒱(ZM,s,ρ,q)\mathcal{V}(Z_{M},s,\rho,q).

Proposition 6.4.1.

There exists an open set 𝒰ZM\mathcal{U}\subset Z_{M}, containing all non-strictly positive characters, so that over 𝒰\mathcal{U} the natural map

(6.4.1) GMKZM,s×(e,M)q|ZMs𝐊ZM,s×(e)q|ZMs{\mathbfcal{H}}_{G}\otimes_{{\mathbfcal{H}}_{M}}K_{\langle Z_{M},s\rangle\times\mathbb{C}^{*}}(\mathcal{B}_{e,M})_{q}|_{Z_{M}s}\to{\bf{K}}_{\langle Z_{M},s\rangle\times\mathbb{C}^{*}}(\mathcal{B}_{e})_{q}|_{Z_{M}s}

induced by ii_{*}, is an isomorphism.

Proof.

It is enough to prove this statement fiberwise. In this form it is contained in [19, 2.2]. ∎

Corollary 6.4.2.

𝒱(ZM,s,ρ,q)|𝒰𝒦(ZM,s,q)ρ|𝒰.\mathcal{V}(Z_{M},s,\rho,q)|_{\mathcal{U}}\simeq\mathcal{K}(Z_{M},s,q)_{\rho}|_{\mathcal{U}}.

Proof.

It follows from the definitions that 𝒱(ZM,s,ρ,q)\mathcal{V}(Z_{M},s,\rho,q) identifies with the LHS of (6.4.1). Now the claim follows from Proposition 6.4.1. ∎

It follows from Corollary 6.4.2 that 𝒮e\mathcal{S}_{e} embeds into

(6.4.2) M,s,ρEndrat.𝒪(ZM)𝒦(ZM,s,q)ρ.\prod_{M,s,\rho}\operatorname{End}^{\mathrm{rat}.}_{\mathcal{O}(Z_{M})}\mathcal{K}(Z_{M},s,q)_{\rho}.

Moreover, by 2) in Section 6.2, and the fact that any character is conjugate to a non-strictly positive one, LHS can be replaced by M,s,e,ρEnd𝒪(ZM)𝒦(ZM,s,q)ρ\prod_{M,s,e,\rho}\operatorname{End}_{\mathcal{O}(Z_{M})}\mathcal{K}(Z_{M},s,q)_{\rho}.

Moreover, similarly to Theorem 5.3.3, we are reduced to showing that the following holds.

Proposition 6.4.3.

Let \mathcal{E} be the subalgebra of M,sEnd𝒪(ZM)(ZM,s,q)=:E~\prod_{M,s}\operatorname{End}_{\mathcal{O}(Z_{M})}\mathcal{L}(Z_{M},s,q)=:\tilde{E} consisting of elements ϕ=(ϕ(M,s))M,s\phi=(\phi(M,s))_{M,s} satisfying the following property.

For any pair of Levi subgroups M,LGM^{\vee},\,L^{\vee}\subset G^{\vee} whose Lie algebras contain ee and elements sMs\in M^{\vee}, tLt\in L^{\vee}; χZM\chi\in Z_{M}, χZL\chi^{\prime}\in Z_{L} such that sχs\chi is conjugate to tχt\chi^{\prime} we have:

(6.4.3) ϕ(M,s)χ=ϕ(L,t)χ.\phi(M,s)_{\chi}=\phi(L,t)_{\chi^{\prime}}.

Then 𝐉e\mathcal{E}\simeq{\bf{J}}_{e} via the natural action map α:𝐉e\alpha\colon{\bf{J}}_{e}\to\mathcal{E}.

The proof of Proposition 6.4.3 will occupy the rest of this section.

6.5.

First we reduce E~\tilde{E} to a finite product.

It is known (cf. [48], [33]) that there exists a set-theoretic lifting l:π0(ZZe(s))=ΓseZZe(s)l\colon\pi_{0}(Z_{Z_{e}}(s))=\Gamma^{s}_{e}\to Z_{Z_{e}}(s), so that for any γπ0(ZZe(s))\gamma\in\pi_{0}(Z_{Z_{e}}(s)), one has:

1) Adl(γ)\operatorname{Ad}_{l(\gamma)} preserves a pinning of ZZe(s)0Z_{Z_{e}}(s)^{0} (say, B±γB^{\pm}_{\gamma}, TγT_{\gamma});

2) every element of ZZe(s)0γZ_{Z_{e}}(s)^{0}\gamma is conjugate to an element in (Tγγ)0γ(T_{\gamma}^{\gamma})^{0}\gamma (the upper index γ\gamma stands for the invariants of Ad\operatorname{Ad}-action).

We will denote (Tγγ)0(T_{\gamma}^{\gamma})^{0} by C(γ)C(\gamma), and we will denote by L(γ)L(\gamma)^{\vee} the Levi subgroup ZG((Tγγ)0)0GZ_{G^{\vee}}((T_{\gamma}^{\gamma})^{0})^{0}\subset G^{\vee}.

From 2) above it follows that the natural projection

π=πγ:γΓesEnd𝒪(ZL(γ))(ZL(γ),γ,q)\pi=\prod\pi_{\gamma}\colon\mathcal{E}\to\prod_{\gamma\in\Gamma_{e}^{s}}\operatorname{End}_{\mathcal{O}(Z_{L(\gamma)})}\mathcal{L}(Z_{L(\gamma)},\gamma,q)

is injective.

6.6.

We calculate the completion ^s\widehat{\mathcal{E}}^{s} for a semisimple conjugacy class [s]Ze[s]\in Z_{e}. Let T(s)T(s) be a maximal torus inside ZZe(s)Z_{Z_{e}}(s). Let M(s)M(s)^{\vee} be ZG(T(s))0Z_{G^{\vee}}(T(s))^{0} and recall that T(s)=ZM(s)T(s)=Z_{M(s)} is the connected component of 1Z(M(s))1\in Z(M(s)^{\vee}).

As above, the map

πs:End𝒪(T(s))(ZM(s),s,q)×γΓe,C(γ)γ[s]=End𝒪(C(γ))(ZL(γ),γ,q)\pi_{s}\colon\mathcal{E}\to\operatorname{End}_{\mathcal{O}(T(s))}\mathcal{L}(Z_{M(s)},s,q)\times\prod\limits_{\gamma\in\Gamma_{e},\\ C(\gamma)\gamma\cap[s]=\emptyset}\operatorname{End}_{\mathcal{O}(C(\gamma))}\mathcal{L}(Z_{L(\gamma)},\gamma,q)

(which is defined in the same manner as π\pi above) is injective.

Proposition 6.6.1.

For any γΓe\gamma\in\Gamma_{e}, such that ZL(γ)γ[s]=Z_{L(\gamma)}\gamma\cap[s]=\emptyset, we have

End𝒪(C(γ))(ZL(γ),γ,q)^s=0.\widehat{\operatorname{End}_{\mathcal{O}(C(\gamma))}\mathcal{L}(Z_{L(\gamma)},\gamma,q)}^{s}=0.

Here the left hand side is the completion of End𝒪(C(γ))(ZL(γ),γ,q)\operatorname{End}_{\mathcal{O}(C(\gamma))}\mathcal{L}(Z_{L(\gamma)},\gamma,q) at the maximal ideal 𝐊Ze(pt){\bf{K}}_{Z_{e}}(\operatorname{pt}) corresponding to ss.

Proof.

Step 1. It suffices to construct an Ad(Ze)\operatorname{Ad}(Z_{e})-invariant function ff on ZeZ_{e} such that f(s)=0f(s)=0 and f|C(γ)γ=1f|_{C(\gamma)\gamma}=1. Thus it is enough to prove that in the coarse quotient Ze//Ad(Ze)Z_{e}/\!/\operatorname{Ad}(Z_{e}) the image of C(γ)γC(\gamma)\gamma is closed and does not contain the image of ss.

This follows from the following properties of such a quotient.

Step 2. Let KK be a possibly non-connected algebraic group. According to the results of Mohrdieck ([48]; cf. also [33]) there exists a set-theoretic lifting l:ΓK:=π0(K)Kl\colon\Gamma_{K}:=\pi_{0}(K)\to K, so that:

1) there exists a maximal torus TK0T\subset K^{0} normalized by l(γ)l(\gamma) for any γΓK\gamma\in\Gamma_{K}.

2) for a certain finite group 𝒲\mathcal{W} we have: Tl(γ)γ/𝒲G0γ//G0.T^{l(\gamma)}\gamma/\mathcal{W}\simeq G^{0}\gamma/\!/G^{0}.

Combining the above statements we get the following.

Lemma 6.6.2.

The map πs^s:^sEnd𝒪(T(s))(ZM(s),s)^s\widehat{\pi_{s}}^{s}:\widehat{\mathcal{E}}^{s}\to\widehat{\operatorname{End}_{\mathcal{O}(T(s))}\mathcal{L}(Z_{M(s)},s)}^{s} is injective.

6.7.

Recall that T(s)T(s) is a maximal torus inside ZZe(s)0Z_{Z_{e}}(s)^{0}, let WsW_{s} being the Weyl group of ZZe(s)0Z_{Z_{e}}(s)^{0}. Then the image of πs^s\widehat{\pi_{s}}^{s} lies inside

(6.7.1) ((End𝒪(T(s))^(ZM(s),s,q)s)Ws)π0(ZZe(s)).((\widehat{\operatorname{End}_{\mathcal{O}(T(s))}}\mathcal{L}(Z_{M(s)},s,q)^{s})^{W_{s}})^{\pi_{0}(Z_{Z_{e}}(s))}.

Let 𝔪1𝐊ZZe(s)(pt)\mathfrak{m}_{1}\subset{\bf{K}}_{Z_{Z_{e}}(s)}(\operatorname{pt}) be the augmentation ideal of 11. Let 𝔪𝐊ZZe(s)0(pt)\mathfrak{m}\subset{\bf{K}}_{Z_{Z_{e}}(s)^{0}}(\operatorname{pt}) be the ideal generated by the image of 𝔪1\mathfrak{m}_{1} in 𝐊ZZe(s)0(pt){\bf{K}}_{Z_{Z_{e}}(s)^{0}}(\operatorname{pt}) (under the restriction homomorphism 𝐊ZZe(s)(pt)𝐊ZZe(s)0(pt){\bf{K}}_{Z_{Z_{e}}(s)}(\operatorname{pt})\rightarrow{\bf{K}}_{Z_{Z_{e}}(s)^{0}}(\operatorname{pt})). Let us note that:

1) (ZM(s),s,q)\mathcal{L}(Z_{M(s)},s,q) is the vector bundle over T(s)T(s) with fiber 𝐊(e,s){\bf{K}}(\mathcal{B}_{e}^{\mathbb{C}^{*},s}); thus, it can be non-canonically identified with 𝐊(e,s)Rs{\bf{K}}(\mathcal{B}_{e}^{\mathbb{C}^{*},s})\otimes R_{s}, where RsR_{s} is 𝐊T(s)(pt){\bf{K}}_{T(s)}(\operatorname{pt});

2) as was already stated, the completion is in the sense of 𝐊Ze(pt){\bf{K}}_{Z_{e}}(\operatorname{pt})-modules;

3) 𝐊Ze(pt)^s=(𝐊ZZe(s)0(pt)^𝔪)π0(ZZe(s))\widehat{{\bf{K}}_{Z_{e}}(\operatorname{pt})}^{s}=(\widehat{{\bf{K}}_{Z_{Z_{e}}(s)^{0}}(\operatorname{pt})}^{\mathfrak{m}})^{\pi_{0}(Z_{Z_{e}}(s))} (cf. Remark A.1.5 for Y=ptY=\operatorname{pt}).

Now we see that the image of πs^s\widehat{\pi_{s}}^{s} lies inside (End(𝐊(e,s))𝐊ZZe(s)0(pt)^𝔪)π0(ZZe(s))(\operatorname{End}({\bf{K}}(\mathcal{B}_{e}^{\mathbb{C}^{*},s}))\otimes\widehat{{\bf{K}}_{Z_{Z_{e}}(s)^{0}}(\operatorname{pt})}^{\mathfrak{m}})^{\pi_{0}(Z_{Z_{e}}(s))}.

But, similarly to (5.2.2), by the results of Section A.1.5,

(6.7.2) 𝐉e^s=𝐊e^s=(End(𝐊(e,s))𝐊ZZe(s)0(pt)𝔪^)π0(ZZe(s)).\widehat{{\bf{J}}_{e}}^{s}=\widehat{{\bf{K}}_{e}}^{s}=(\operatorname{End}({\bf{K}}(\mathcal{B}_{e}^{\mathbb{C}^{*},s}))\otimes\widehat{{\bf{K}}_{Z_{Z_{e}}(s)^{0}}(\operatorname{pt})^{\mathfrak{m}}})^{\pi_{0}(Z_{Z_{e}}(s))}.

Now it follows that the map α\alpha (introduced in (6.4.3)) induces an identical isomorphism on completions (and, in particular, πs^s:^s(End(𝐊(e,s))𝐊ZZe(s)0(pt)𝔪^)π0(ZZe(s))\widehat{\pi_{s}}^{s}:\widehat{\mathcal{E}}^{s}\to(\operatorname{End}({\bf{K}}(\mathcal{B}_{e}^{\mathbb{C}^{*},s}))\otimes\widehat{{\bf{K}}_{Z_{Z_{e}}(s)^{0}}(\operatorname{pt})^{\mathfrak{m}}})^{\pi_{0}(Z_{Z_{e}}(s))} is surjective). Thus, Proposition 6.4.3 is proven.

Remark 6.7.1.

Isomorphisms in 1) and (6.7.2) in this subsection are non-canonical, cf. Appendix A.1. As follows from loc. cit., both of them are uniquely determined by the choice of a trivialization of the vector bundle 𝐊T(s)(e,s){\bf{K}}_{T(s)}(\mathcal{B}_{e}^{\mathbb{C}^{*},s}) over 𝐊T(s)(pt){\bf{K}}_{T(s)}(\operatorname{pt}). Above we implicitly assumed that the same trivialization is used in both cases.

7. The structure of the cocenter of JJ

7.1.

In this concluding section we will describe C(𝐉)C({\bf{J}}) and C()qC({\mathbfcal{H}})_{q} for generic qq\in\mathbb{C}^{*}, where CC stands for the cocenter, i.e. the 0-th Hochschild homology of an associative ring. This will prove Conjecture 22 from [12]. The main theorem is as follows.

For a reductive algebraic group HH let 𝐂𝐨𝐦𝐦H{\bf{Comm}}_{H}, CommH\operatorname{Comm}_{H} be the corresponding commuting variety equipped with the natural (respectively, reduced) scheme structure.

We will be also interested in the categorical (coarse) quotients of these schemes:

𝐂𝐨𝐦𝐦H//H=:𝐂H,CommH//H=:𝒞H.{\bf{Comm}}_{H}/\!/H=:{\bf C}_{H},\,\operatorname{Comm}_{H}/\!/H=:\mathcal{C}_{H}.
Remark 7.1.1.

By the result of [46] the ring 𝒪(𝐂H)=𝒪(𝐂𝐨𝐦𝐦H)H{\mathcal{O}}({\bf C}_{H})={\mathcal{O}}({\bf{Comm}}_{H})^{H} is reduced if the reductive group HH is connected. One expects that this theorem generalizes to the case of a not necessarily connected reductive group, such a generalization would imply that 𝐂H=𝒞H{\bf C}_{H}=\mathcal{C}_{H}.

Let 𝒪a(𝒞Ze)=𝒪a(e)𝒪(𝒞Ze)\mathcal{O}^{a}(\mathcal{C}_{Z_{e}})=\mathcal{O}^{a}(e)\subset\mathcal{O}\mathcal{(}\mathcal{C}_{Z_{e}}) (where ‘‘aa’’ stands for ‘‘admissible’’) denote the vector space of regular functions ff on 𝒞Ze\mathcal{C}_{Z_{e}} satisfying the following property.

For a semisimple sZes\in Z_{e} let fsf_{s} denote the pull-back of ff under the map x(x,s)x\mapsto(x,s) from the centralizer ZZe(s)Z_{Z_{e}}(s) of ss in ZeZ_{e} to CommZe\operatorname{Comm}_{Z_{e}}. Then f𝒪a(e)f\in\mathcal{O}^{a}(e) iff for any semisimple sZes\in Z_{e} the function fsf_{s} is a linear combination of admissible characters of the group ZZe(s)Z_{Z_{e}}(s). Here by an admissible character we understand the character of an irreducible representation appearing in 𝐊(es){\bf{K}}(\mathcal{B}_{e}^{s}), (or, equivalently, in 𝐊(e,s){\bf{K}}(\mathcal{B}_{e}^{\mathbb{C}^{*},s}); cf. [40, 2.8]).

Notice that an admissible representation factors through the group of components of ZZe(s)Z_{Z_{e}}(s), thus for f𝒪a(e)f\in\mathcal{O}^{a}(e) the function fsf_{s} is locally constant for every sZes\in Z_{e}.

Theorem 7.1.2 (Theorem E).

We have a canonical isomorphism C(𝐊Ze(e×e))𝒪a(e)C({\bf{K}}_{Z_{e}}(\mathcal{B}_{e}^{{\mathbb{C}}^{*}}\times\mathcal{B}_{e}^{{\mathbb{C}}^{*}}))\simeq\mathcal{O}^{a}(e).

Corollary 7.1.3.

We have isomorphisms:

(7.1.1) 𝒪a(e)C(𝐉e),\mathcal{O}^{a}(e)\simeq C({\bf{J}}_{e}),
(7.1.2) e𝒩/𝒪a(e)C(q),\bigoplus\limits_{e\in\mathcal{N}/\sim}\mathcal{O}^{a}(e)\simeq C({\mathbfcal{H}}_{q}),


where isomorphism (7.1.1) depends on a choice of q{±1}q\in\{\pm 1\}; in (7.1.2) qq is assumed not to be the root of unity.

Proof.

The first isomorphism follows from Theorem 7.1.2 and Theorem A. The second one then follows from [12, Theorem 1] which establishes an isomorphism C(Je)C(q)C(J_{e})\simeq C(\mathcal{H}_{q}) when qq is not a root of unity. ∎

Remark 7.1.4.

Another approach to constructing isomorphism (7.1.2), as well as a stronger version describing the unipotent part of the cocenter of the pp-adic group, will be presented in [11] (see also a related result [1]).

In that approach it is realized as a ‘‘decategorification’’ of a result of Ben Zvi-Nadler-Preygel [3] describing the trace of the affine Hecke category as coherent sheaves on commuting pairs of elements in GG.

The rest of this section is devoted to proofs.

7.2.

Let XX be a scheme, {\mathcal{F}} a locally free coherent sheaf on XX and ϕ:\phi\colon{\mathcal{F}}\to{\mathcal{F}} and endomorphism. To this data one assigns a regular function Tr(ϕ)Γ(X,𝒪X)\operatorname{Tr}(\phi)\in\Gamma(X,\mathcal{O}_{X}); this is a special case of the more general Hattori-Stallings trace (cf. [31], [53]). The trace is additive on short exact sequences, thus it extends to the derived category of perfect complexes. The construction is manifestly local in the fppf topology, so it extends to algebraic stacks.

Let now XX be an algebraic stack over a field kk and I(X)=X×X2kXI(X)=X\times_{X^{2}_{k}}X be the inertia stacks. We get two natural isomorphisms between the composed morphisms of stacks I(X)X2pr1XI(X)\to X^{2}\overset{pr_{1}}{\longrightarrow}X and I(X)X2pr2XI(X)\to X^{2}\overset{pr_{2}}{\longrightarrow}X, composing the first one with the inverse of the second we get an automorphism of the first composition (we denote that composition by prpr). Thus for Coh(X){\mathcal{F}}\in\operatorname{Coh}(X) the sheaf pr()pr^{*}({\mathcal{F}}) acquires a canonical automorphism cc_{\mathcal{F}}. For example, if X=S/HX=S/H where SS is a scheme and HH an algebraic group then I(X)=I~(X)/HI(X)=\tilde{I}(X)/H where I~(X)={(x,h)|h(x)=x}S×H\tilde{I}(X)=\{(x,h)\ |\ h(x)=x\}\subset S\times H. In this case the action of cc_{\mathcal{F}} on the fiber of pr()pr^{*}({\mathcal{F}}) at (x,h)(x,h) equals the action of hh on the fiber of {\mathcal{F}} at xx.

Remark 7.2.1.

One expects an isomorphism RΓ(𝒪(I(X)))HH(Coh(X))R\Gamma({\mathcal{O}}(I(X)))\simeq\operatorname{HH}_{*}(\operatorname{Coh}(X)), having such an isomorphism one could define the function cc_{\mathcal{F}} as the image of the class [][{\mathcal{F}}] under the trace map from KK-theory to Hochschild homology.777We thank Jakub Löwit who pointed it out to us. We were not able to find a reference for such an isomorphism, so we resorted to the above less direct construction.

Lemma 7.2.2.

Let XX be a scheme over a characteristic zero field kk and HH an affine algebraic group over kk. Let {\mathcal{F}} be a perfect complex on the stack X/HX/H and f=Tr(c)𝒪(I(X))f=\operatorname{Tr}(c_{\mathcal{F}})\in{\mathcal{O}}(I(X)). For sH(k)s\in H(k) let is:XsI(X)i_{s}\colon X^{s}\to I(X) be given by x(x,s)x\mapsto(x,s). Then is(f)i_{s}^{*}(f) is locally constant.

Proof.

Without loss of generality we can assume that kk is algebraically closed.

Let HsH_{s} be the Zariski closure of the cyclic group s\langle s\rangle in HH. Thus HsH_{s} is abelian algebraic group, so it is a product of a diagonalizable and a vector group. The restriction of an equivariant coherent sheaf {\mathcal{F}} to XsX^{s} carries an action of the diagonalizable group HsdiagH_{s}^{\mathrm{diag}}, thus s:=|Xs{\mathcal{F}}_{s}:={\mathcal{F}}|_{X^{s}} splits as a direct sum of subsheaves sχ{\mathcal{F}}_{s}^{\chi} where χ\chi runs over the characters of HsH_{s}, where HsH_{s} acts on sχ{\mathcal{F}}_{s}^{\chi} via χ\chi. It follows that such a decomposition is also well-defined for a perfect complex. Since the Euler characteristic of a perfect complex is a locally constant function, the claim follows. ∎

We now consider X=H/HX=H/H, the quotient of a reductive algebraic group by the conjugation action. Thus I(X)=𝐂𝐨𝐦𝐦H/HI(X)={\bf Comm}_{H}/H; applying the above construction we get a map τH:𝐊H(H)𝒪(𝐂H)\tau_{H}\colon{\bf{K}}^{H}(H)\to\mathcal{O}({\bf C}_{H}) []Tr(c)[{\mathcal{F}}]\mapsto\operatorname{Tr}(c_{\mathcal{F}}). By Lemma 7.2.2 it lands in the space 𝒪l(𝐂)\mathcal{O}_{l}({\bf C}) of functions satisfying the local constancy condition stated in the Lemma.

7.3.

We now proceed to prove the following reformulation of Theorem 7.1.2.

Theorem 7.3.1.

Set X=eX=\mathcal{B}_{e}^{\mathbb{C}^{*}}, let 𝐚:Ze×XX{\bf{a}}\colon Z_{e}\times X\to X be the action map, and π:Ze×XX\pi\colon Z_{e}\times X\to X the projection. Consider the map

c:𝐉e𝐊Ze(X2)(𝐚×Id)𝐊Ze(Ze×X)π𝐊Ze(Ze)τZe𝒪l(𝐂Ze)𝒪(𝒞Ze),c\colon{\bf{J}}_{e}\simeq{\bf{K}}_{Z_{e}}(X^{2})\xrightarrow{({\bf{a}}\times\operatorname{Id})^{*}}{\bf{K}}_{Z_{e}}(Z_{e}\times X)\xrightarrow{\pi_{*}}{\bf{K}}_{Z_{e}}(Z_{e})\xrightarrow{\tau_{Z_{e}}}\mathcal{O}_{l}({\bf{C}}_{Z_{e}})\to\mathcal{O}(\mathcal{C}_{Z_{e}}),

where the last arrow is the restriction to the reduced subscheme 𝒞Ze𝐂Ze\mathcal{C}_{Z_{e}}\subset{\bf C}_{Z_{e}}. (Compare with the definiton of the map TT in A.4.1.)

Then: 1. [Je,Je][J_{e},J_{e}] lies in the kernel of cc, and cc induces an injective map on Je/[Je,Je]J_{e}/[J_{e},J_{e}].

2. The image of cc lies inside 𝒪a(e)\mathcal{O}^{a}(e).

3. The image of cc equals 𝒪a(e)\mathcal{O}^{a}(e).

We will start with the proof of 22, and then proceed with 11 and 33.

7.4. Admissibility

7.4.1.

Recall that X=eX=\mathcal{B}^{{\mathbb{C}}^{*}}_{e}. For CohZe(X×X)\mathcal{F}\in\operatorname{Coh}_{Z_{e}}(X\times X), we will denote σ\sigma^{*}\mathcal{F} by ¯\bar{\mathcal{F}}, where σ\sigma is the involution swapping factors. For \mathcal{F} as above, and 𝒫=π(𝐚×Id)\mathcal{P}=\pi_{*}({\bf{a}}\times\operatorname{Id})^{*}\mathcal{F}, f𝒫|{s}×ZZe(s)f_{\mathcal{P}}|_{\{s\}\times Z_{Z_{e}}(s)} (by Lemma 7.2.2) can be considered as a function on π0(ZZe(s))=:Γse\pi_{0}(Z_{Z_{e}}(s))=:\Gamma^{s}_{e} invariant under the conjugation. Let us denote this function by fsf_{\mathcal{F}}^{s}. Recall that for γΓse\gamma\in\Gamma^{s}_{e} we have

fs(γ)=Trsiγπ(𝐚×Id),f^{s}_{\mathcal{F}}(\gamma)=\operatorname{Tr}_{s}i_{\gamma}^{*}\pi_{*}({\bf{a}}\times\operatorname{Id})^{*}\mathcal{F},

where iγ:{γ~}Zei_{\gamma}\colon\{\tilde{\gamma}\}\hookrightarrow Z_{e} is an embedding in ZeZ_{e} of some lifting γ~Ze\tilde{\gamma}\in Z_{e} of γ\gamma.

Lemma 7.4.1.

Let ρ\rho be any irreducible representation of Γse\Gamma^{s}_{e} and let χρ\chi_{\rho} be its character. Then the invariant pairing χρ,fs\langle\chi_{\rho},f_{\mathcal{F}}^{s}\rangle is equal to Tr¯(𝐊(e,s)ρ)\operatorname{Tr}_{\bar{\mathcal{F}}}({\bf{K}}(\mathcal{B}_{e}^{\mathbb{C}^{*},s})_{\rho}) (recall that ¯\bar{\mathcal{\mathcal{F}}} acts on 𝐊(e,s)ρ{\bf{K}}(\mathcal{B}_{e}^{\mathbb{C}^{*},s})_{\rho} via the convolution).

Proof.

We start with a piece of notation.

Let us for γΓes\gamma\in\Gamma_{e}^{s} consider the operator γ:=γ\gamma\mathcal{F}:=\gamma\circ\mathcal{F} acting on 𝐊(Xs){\bf{K}}(X^{s}). Let also Eu\operatorname{Eu} stand for the Euler characteristic.

Step 1. Since (all - and -functors below are derived)

χρ,fs=1|Γse|γΓseχρ(γ1)fs(γ)=1|Γse|γΓseχρ(γ1)Trsiγπ(𝐚×Id),\langle\chi_{\rho},f_{\mathcal{F}}^{s}\rangle=\frac{1}{|\Gamma^{s}_{e}|}\sum_{\gamma\in\Gamma^{s}_{e}}\chi_{\rho}(\gamma^{-1})f^{s}_{\mathcal{F}}(\gamma)=\frac{1}{|\Gamma^{s}_{e}|}\sum_{\gamma\in\Gamma^{s}_{e}}\chi_{\rho}(\gamma^{-1})\operatorname{Tr}_{s}i_{\gamma}^{*}\pi_{*}({\bf{a}}\times\operatorname{Id})^{*}\mathcal{F},

(here and below Trs\operatorname{Tr}_{s} is always a graded trace), and

Tr¯(𝐊(e,s)ρ)=1|Γse|γΓseχρ(γ1)Trγ¯(𝐊(e,s)),\operatorname{Tr}_{\bar{\mathcal{F}}}({\bf{K}}(\mathcal{B}_{e}^{\mathbb{C}^{*},s})_{\rho})=\frac{1}{|\Gamma^{s}_{e}|}\sum_{\gamma\in\Gamma^{s}_{e}}\chi_{\rho}(\gamma^{-1})\operatorname{Tr}_{\gamma\bar{\mathcal{F}}}({\bf{K}}(\mathcal{B}_{e}^{\mathbb{C}^{*},s})),

we have to prove that

Trγ¯(𝐊(e,s))=Trs(iγπ(𝐚×Id)).\operatorname{Tr}_{\gamma\bar{\mathcal{F}}}({\bf{K}}(\mathcal{B}_{e}^{\mathbb{C}^{*},s}))=\operatorname{Tr}_{s}(i_{\gamma}^{*}\pi_{*}({\bf{a}}\times\operatorname{Id})^{*}\mathcal{F}).

Equivalently, we are interested in the following equality:

(7.4.1) Trγ¯(𝐊(e,s))=TrsH(X,Γγ),\operatorname{Tr}_{\gamma\bar{\mathcal{F}}}({\bf{K}}(\mathcal{B}_{e}^{\mathbb{C}^{*},s}))=\operatorname{Tr}_{s}H^{*}(X,\Gamma_{\gamma}^{*}\mathcal{F}),

where Γγ\Gamma_{\gamma} stands for the graph embedding γ:XX×X\gamma\colon X\to X\times X and X=eX=\mathcal{B}^{{\mathbb{C}}^{*}}_{e}.

(Let us recall, that we can interchange KK-theory and (co)homology of XX because of [26, Theorem 3.9] and [16, Theorem III.1 (b)].)

Step 2. Now we would like to use the results of Section A.4. Namely, Proposition A.4.1 together with the proof of Proposition A.2.1 say that

(7.4.2) Trγ¯(𝐊(e,s))=χχEu((λs1𝒪Xs)𝕃(Γs(γ)is¯)𝕃𝒪Δ)\operatorname{Tr}_{\gamma\bar{\mathcal{F}}}({\bf{K}}(\mathcal{B}_{e}^{\mathbb{C}^{*},s}))=\sum_{\chi\in{\mathbb{C}}^{*}}\chi\operatorname{Eu}((\lambda_{s}^{-1}\boxtimes\mathcal{O}_{X^{s}})\otimes^{\mathbb{L}}(\Gamma^{s}(\gamma)*i^{*}_{s}\bar{\mathcal{F}})\otimes^{\mathbb{L}}\mathcal{O}_{\Delta})

for Δ\Delta being the diagonal inside Xs×XsX^{s}\times X^{s}, Γs(γ)\Gamma^{s}(\gamma) being a structure sheaf of the graph of γ\gamma inside Xs×XsX^{s}\times X^{s}, and is:Xs×XsX×Xi_{s}\colon X^{s}\times X^{s}\hookrightarrow X\times X being the natural embedding.

But, using a base change from [58, Proposition 1.4], as in the proof of Proposition 4.1.5, for the diagram

Xs×Xs{{X^{s}\times X^{s}}}Xs×Xs×Xs{{X^{s}\times X^{s}\times X^{s}}}Xs{X^{s}}Xs×Xs,{{X^{s}\times X^{s},}}π13\scriptstyle{\pi_{13}}π1\scriptstyle{\pi_{1}}iΔ\scriptstyle{i_{\Delta}}(a,b)(a,b,a)\scriptstyle{(a,b)\mapsto(a,b,a)}j\scriptstyle{j}

one sees that for any sheaves 𝒜\mathcal{A} and \mathcal{B} on Xs×XsX^{s}\times X^{s},

(7.4.3) (𝒜)𝕃𝒪Δ=iΔπ13(π12𝒜𝕃π23)=π1j(π12𝒜𝕃π23)=π1(𝒜𝕃σ)=π1(𝒜𝕃¯),(\mathcal{A}*\mathcal{B})\otimes^{\mathbb{L}}\mathcal{O}_{\Delta}=i_{\Delta}^{*}\pi_{13*}(\pi_{12}^{*}\mathcal{A}\otimes^{\mathbb{L}}\pi_{23}^{*}\mathcal{B})=\pi_{1*}j^{*}(\pi_{12}^{*}\mathcal{A}\otimes^{\mathbb{L}}\pi_{23}^{*}\mathcal{B})=\pi_{1*}(\mathcal{A}\otimes^{\mathbb{L}}\sigma^{*}\mathcal{B})=\pi_{1*}(\mathcal{A}\otimes^{\mathbb{L}}\bar{\mathcal{B}}),

where all of the functors are, as usual, derived (this also holds below); and iΔi_{\Delta} is an embedding of the diagonal.

So, by the proof of Proposition A.2.1, we get (in the notation of loc. cit.)

(7.4.4) Trγ¯(𝐊(e,s))=χχEu((λs1𝒪Xs)𝕃Γs(γ)𝕃isχ)=\operatorname{Tr}_{\gamma\bar{\mathcal{F}}}({\bf{K}}(\mathcal{B}_{e}^{\mathbb{C}^{*},s}))=\sum_{\chi\in{\mathbb{C}}^{*}}\chi\operatorname{Eu}((\lambda_{s}^{-1}\boxtimes\mathcal{O}_{X^{s}})\otimes^{\mathbb{L}}\Gamma^{s}(\gamma)\otimes^{\mathbb{L}}i_{s}^{*}\mathcal{F}_{\chi})=
=χχEu(Γγs)((λs1𝒪Xs)𝕃isχ)=χχEu(λs1𝕃(Γγs)isχ).=\sum_{\chi\in{\mathbb{C}}^{*}}\chi\operatorname{Eu}(\Gamma_{\gamma}^{s})^{*}((\lambda_{s}^{-1}\boxtimes\mathcal{O}_{X^{s}})\otimes^{\mathbb{L}}i_{s}^{*}\mathcal{F}_{\chi})=\sum_{\chi\in{\mathbb{C}}^{*}}\chi\operatorname{Eu}(\lambda_{s}^{-1}\otimes^{\mathbb{L}}(\Gamma_{\gamma}^{s})^{*}i_{s}^{*}\mathcal{F}_{\chi}).

Here Γγs\Gamma_{\gamma}^{s} is an embedding XsXs×XsX^{s}\to X^{s}\times X^{s} of the graph of γ\gamma, and we use the following simple observation: for 𝐚γs:XsXs{\bf{a}}_{\gamma}^{s}\colon X^{s}\to X^{s} being a γ\gamma-action map, (Γγs)(λs1𝒪Xs)=iΔ(1𝐚γs)(λs1𝒪Xs)=λs1𝒪Xs=λs1(\Gamma_{\gamma}^{s})^{*}(\lambda_{s}^{-1}\boxtimes\mathcal{O}_{X^{s}})=i_{\Delta}^{*}(1\boxtimes{\bf{a}}_{\gamma}^{s})^{*}(\lambda_{s}^{-1}\boxtimes\mathcal{O}_{X^{s}})=\lambda_{s}^{-1}\otimes\mathcal{O}_{X^{s}}=\lambda_{s}^{-1}.

Step 3. On the other hand, by a localization theorem (cf. [20, 5.11.8] and the proof of our Proposition A.2.1),

(7.4.5) TrsH(X,Γγ)=TrsH(Xs,λ(s)1𝕃isΓγ),\operatorname{Tr}_{s}H^{*}(X,\Gamma_{\gamma}^{*}\mathcal{F})=\operatorname{Tr}_{s}H^{*}(X^{s},\lambda(s)^{-1}\otimes^{\mathbb{L}}{i^{\prime}_{s}}^{*}\Gamma_{\gamma}^{*}\mathcal{F}),

where is:XsXs×Xsi^{\prime}_{s}\colon X^{s}\hookrightarrow X^{s}\times X^{s} is the natural embedding.

It is clear that (7.4.5) is equal to (7.4.4). ∎

7.4.2.

By the previous lemma, only admissible characters of Γse\Gamma^{s}_{e} may enter fsf_{\mathcal{F}}^{s}. This finishes the proof of the admissibility.

7.5. Injectivity

The fact that [𝐉e,𝐉e][{\bf{J}}_{e},{\bf{J}}_{e}] lies inside the kernel, follows from the proof of Lemma 7.4.1. However, the more geometric proof of the same fact exists (which also does not require our discussion about the reducedness above). For this, see A.4.1.

In view of Lemma 7.4.1, injectivity follows once we check the density of characters for 𝐉{\bf{J}}, i.e., the statement that an element h𝐉h\in{\bf{J}} whose trace in every finite dimensional 𝐉{\bf{J}}-module vanishes lies in [𝐉,𝐉][{\bf{J}},{\bf{J}}]. By [12, Theorem 1] the cocenter of 𝐉{\bf{J}} is isomorphic to the cocenter of q{\mathbfcal{H}}_{q} for almost all qq. Thus the density of characters for 𝐉{\bf{J}} follows from the density of characters for q{\mathbfcal{H}}_{q} together with the fact that pull back of finite dimensional 𝐉{\bf{J}} modules generate the Grothendieck group of finite dimensional q{\mathbfcal{H}}_{q} modules (see [39, Lemma 1.9]). ∎

7.6. Surjectivity

By the results of the previous subsections, we have a morphism c:𝐊Ze(X×X)𝐊Ze(Ze)c\colon{\bf{K}}_{Z_{e}}(X\times X)\to{\bf{K}}_{Z_{e}}(Z_{e}), whose image lies in 𝒪a(e)\mathcal{O}^{a}(e). We proceed to prove that in fact every element of the latter subalgebra lies inside Imc\operatorname{Im}c.

We start with the following statement. Let RR be 𝐊ZZe(s)0(pt){\bf{K}}_{Z_{Z_{e}}(s)^{0}}(\operatorname{pt}), set Y:=SpecRY:=\operatorname{Spec}R.

Consider Y1:=Spec(γΓes𝒪(Yγ))ΓesY_{1}:=\operatorname{Spec}(\oplus_{\gamma\in\Gamma_{e}^{s}}\mathcal{O}(Y^{\gamma}))^{\Gamma_{e}^{s}}, the spectrum of the ring of global functions on the inertia stack of Y/ΓY/\Gamma. Any point of Y1Y_{1} (which is reduced, since YY is smooth and Γes\Gamma_{e}^{s} is finite) can be interpreted as (the conjugacy class of) a pair (x,γ)(x,\gamma), where xYγ.x\in Y^{\gamma}. Recall that we have the natural homomorphisms 𝐊Ze(s)(Ze(s))𝒪(Y)Γse𝒪(Y1){\bf{K}}_{Z_{e}(s)}(Z_{e}(s))\rightarrow\mathcal{O}(Y)^{\Gamma^{s}_{e}}\rightarrow\mathcal{O}(Y_{1}) so 𝒪(Y1)\mathcal{O}(Y_{1}) is a module over 𝐊Ze(s)(Ze(s)){\bf{K}}_{Z_{e}(s)}(Z_{e}(s)).

Proposition 7.6.1.

The natural morphism κ:𝐊ZZe(s)(ZZe(s))𝒪(Y1)\kappa\colon{\bf{K}}_{Z_{Z_{e}}(s)}(Z_{Z_{e}}(s))\to\mathcal{O}(Y_{1}), ϕ\mathcal{F}\mapsto\phi_{\mathcal{F}}; ϕ:(x,γ)Trxγ\phi_{\mathcal{F}}\colon(x,\gamma)\mapsto\operatorname{Tr}_{x}\mathcal{F}_{\gamma}, becomes an isomorphism after completion at 11 (as of 𝐊Ze(s)(Ze(s)){\bf{K}}_{Z_{e}(s)}(Z_{e}(s))-modules).

Proof.

The proof is based on the material of Appendix A.

Step 1. First of all, let us note that, by Section A.1.5, 𝐊ZZe(s)(ZZe(s))^𝔪ZZe(s),1(𝐊ZZe(s)0(ZZe(s))^𝔪)Γes\widehat{{\bf{K}}_{Z_{Z_{e}}(s)}(Z_{Z_{e}}(s))}^{\mathfrak{m}_{Z_{Z_{e}}(s),1}}\simeq(\widehat{{\bf{K}}_{Z_{Z_{e}}(s)^{0}}(Z_{Z_{e}}(s))}^{\mathfrak{m}})^{\Gamma_{e}^{s}}.

(Here we complete the RHS at 𝔪=(𝔪ZZe(s),1)𝐊ZZe(s)0(pt)\mathfrak{m}=(\mathfrak{m}_{Z_{Z_{e}}(s),1})\subset{\bf{K}}_{Z_{Z_{e}}(s)^{0}}(\operatorname{pt})).

Let us note that 𝐊ZZe(s)0(ZZe(s))γΓes𝐊ZZe(s)0(ZZe(s)0γ){\bf{K}}_{Z_{Z_{e}}(s)^{0}}(Z_{Z_{e}}(s))\simeq\bigoplus_{\gamma\in\Gamma_{e}^{s}}{\bf{K}}_{Z_{Z_{e}}(s)^{0}}(Z_{Z_{e}}(s)^{0}\gamma). (Here we’ve chosen some set-theoretic lifting of Γes\Gamma_{e}^{s} to ZZe(s)Z_{Z_{e}}(s).)

Thus, it suffices to see the equality

𝐊ZZe(s)0(ZZe(s)0γ)^𝔪𝒪(Yγ)^𝔪.\widehat{{\bf{K}}_{Z_{Z_{e}}(s)^{0}}(Z_{Z_{e}}(s)^{0}\gamma)}^{\mathfrak{m}}\simeq\widehat{\mathcal{O}(Y^{\gamma})}^{\mathfrak{m}}.

Step 2. Let ZZ be the cover of ZZe(s)0Z_{Z_{e}}(s)^{0}, so that [Z,Z][Z,Z] is simply connected. Functoriality of the construction of such a cover allows us to lift an automorphism Adγ\operatorname{Ad}_{\gamma} of ZZe(s)0Z_{Z_{e}}(s)^{0} to ZZ. Let us denote the latter automorphism by AγA_{\gamma}.

First of all, let UZZ/adZ=Spec𝐊Z(pt)U_{Z}\subset Z/_{\operatorname{ad}}Z=\operatorname{Spec}{\bf{K}}_{Z}(\operatorname{pt}) be a Γse\Gamma^{s}_{e}-invariant open neighbourhood of 11 such that the composition UZZ/adZZZe(s)0/adZZe(s)0U_{Z}\subset Z/_{\operatorname{ad}}Z\to Z_{Z_{e}}(s)^{0}/_{\operatorname{ad}}Z_{Z_{e}}(s)^{0} is an isomorpism onto its image that we denote by UZZe(s)0U_{Z_{Z_{e}}(s)^{0}}. Let 𝔪U[UZZe(s)0]\mathfrak{m}_{U}\subset{\mathbb{C}}[U_{Z_{Z_{e}}(s)^{0}}] be the ideal corresponding to 𝔪\mathfrak{m} and let 𝔫U[UZ]\mathfrak{n}_{U}\subset{\mathbb{C}}[U_{Z}] be its preimage. It follows from Proposition A.1.7 that we have an isomorphism

𝐊ZZe(s)0(ZZe(s)0γ)|UZZe(s)0^𝔪U𝐊Z(ZZe(s)0γ)|UZ^𝔫U.\widehat{{\bf{K}}_{Z_{Z_{e}}(s)^{0}}(Z_{Z_{e}}(s)^{0}\gamma)|_{U_{Z_{Z_{e}}(s)^{0}}}}^{\mathfrak{m}_{U}}\simeq\widehat{{\bf{K}}_{Z}(Z_{Z_{e}}(s)^{0}\gamma)|_{U_{Z}}}^{\mathfrak{n}_{U}}.

Moreover, results of Proposition A.3.2 show us that

𝐊Z(ZZe(s)0γ)𝐊Z2(ZZe(s)0γ)𝐊Z2(pt)𝐊Z(pt).{\bf{K}}_{Z}(Z_{Z_{e}}(s)^{0}\gamma)\simeq{\bf{K}}_{Z^{2}}(Z_{Z_{e}}(s)^{0}\gamma)\otimes_{{\bf{K}}_{Z^{2}}(\operatorname{pt})}{\bf{K}}_{Z}(\operatorname{pt}).

Here Z2Z^{2}-action on ZZe(s)0γZ_{Z_{e}}(s)^{0}\gamma is obtained via the natural projection ZZZe(s)0Z\to Z_{Z_{e}}(s)^{0} and regular right-left translations of the latter group on itself.

Now, let us note that the action of Z2Z^{2} on ZZe(s)γZ_{Z_{e}}(s)\gamma is transitive with a stabilizer of γ\gamma being isomorphic to a finite cover of ZZe(s)0.Z_{Z_{e}}(s)^{0}. Let us denote this group by SS.

Then we get

𝐊Z(ZZe(s)0γ)𝐊S(pt)𝐊Z2(pt)𝐊Z(pt);{\bf{K}}_{Z}(Z_{Z_{e}}(s)^{0}\gamma)\simeq{\bf{K}}_{S}(\operatorname{pt})\otimes_{{\bf{K}}_{Z^{2}}(\operatorname{pt})}{\bf{K}}_{Z}(\operatorname{pt});

here 𝐊Z2(pt){\bf{K}}_{Z^{2}}(\operatorname{pt})-action on 𝐊S(pt){\bf{K}}_{S}(\operatorname{pt}) differs from the natural one (i.e., from the one induced by the diagonal embedding ZZ2Z\to Z^{2}) by the AγA_{\gamma}-twisting.

Step 3.

After restricting to UZU_{Z} and completing at 𝔫U\mathfrak{n}_{U} we obtain:

𝐊Z(ZZe(s)0γ)|UZ^𝔫U𝐊S(pt)𝐊Z2(pt)𝐊Z(pt)|UZ^𝔫U=𝐊S(pt)|UZ×UZ[UZ×UZ]𝐊Z(pt)|UZ^𝔫U.\widehat{{\bf{K}}_{Z}(Z_{Z_{e}}(s)^{0}\gamma)|_{U_{Z}}}^{\mathfrak{n}_{U}}\simeq{\bf{K}}_{S}(\operatorname{pt})\otimes_{{\bf{K}}_{Z^{2}}(\operatorname{pt})}\widehat{{\bf{K}}_{Z}(\operatorname{pt})|_{U_{Z}}}^{\mathfrak{n}_{U}}={\bf{K}}_{S}(\operatorname{pt})|_{U_{Z}\times U_{Z}}\otimes_{{\mathbb{C}}[U_{Z}\times U_{Z}]}\widehat{{\bf{K}}_{Z}(\operatorname{pt})|_{U_{Z}}}^{\mathfrak{n}_{U}}.

Recall now that UZUZZe(s)0U_{Z}\,\vphantom{j^{X^{2}}}\smash{\overset{\sim}{\vphantom{\rule{0.0pt}{1.99997pt}}\smash{\longrightarrow}}}\,U_{Z_{Z_{e}}(s)^{0}} and the composition (UZ×UZ)S(UZZe(s)0×UZZe(s)0)ZZe(s)0(U_{Z}\times U_{Z})\cap S\,\vphantom{j^{X^{2}}}\smash{\overset{\sim}{\vphantom{\rule{0.0pt}{1.99997pt}}\smash{\longrightarrow}}}\,(U_{Z_{Z_{e}}(s)^{0}}\times U_{Z_{Z_{e}}(s)^{0}})\cap Z_{Z_{e}}(s)^{0} is an isomorphism.

We conclude that 𝐊ZZe(s)0(ZZe(s)0γ)^𝔪𝐊Ze(s)0(pt)𝐊(Ze(s)0)2(pt)𝐊Ze(s)0(pt)^𝔪\widehat{{\bf{K}}_{Z_{Z_{e}}(s)^{0}}(Z_{Z_{e}}(s)^{0}\gamma)}^{\mathfrak{m}}\simeq{\bf{K}}_{Z_{e}(s)^{0}}(\operatorname{pt})\otimes_{{\bf{K}}_{(Z_{e}(s)^{0})^{2}}(\operatorname{pt})}\widehat{{\bf{K}}_{Z_{e}(s)^{0}}(\operatorname{pt})}^{\mathfrak{m}}.

And, hence, 𝐊ZZe(s)0(ZZe(s)0γ)^𝔪𝒪(Yγ)^𝔪\widehat{{\bf{K}}_{Z_{Z_{e}}(s)^{0}}(Z_{Z_{e}}(s)^{0}\gamma)}^{\mathfrak{m}}\simeq\widehat{\mathcal{O}(Y^{\gamma})}^{\mathfrak{m}} (since YγY^{\gamma} is an intersection of a diagonal and of a graph of γ\gamma acting on YY).

7.6.1.

Now we can start the proof of the surjectivity.

Let us now, first of all, recall that (cf. Section A.1) for any ZeZ_{e}-variety VV, 𝐊Ze(V)^s𝐊ZZe(s)(V)^1\widehat{{\bf{K}}_{Z_{e}}(V)}^{s}\simeq\widehat{{\bf{K}}_{Z_{Z_{e}}(s)}(V)}^{1}. In particular,

(7.6.1) 𝐊Ze(Ze)^s𝐊ZZe(s)(ZZe(s)^)1\widehat{{\bf{K}}_{Z_{e}}(Z_{e})}^{s}\simeq\widehat{{\bf{K}}_{Z_{Z_{e}}(s)}(Z_{Z_{e}}(s)})^{1}

For a subalgebra 𝒪a(e)\mathcal{O}^{a}(e), let us denote the image of its completion at 11 in the RHS of (7.6.1) by 𝒪^1\widehat{\mathcal{O}}^{1}.

Now, we have to prove that the map

(7.6.2) cs:KZZe(s)(Xs×Xs)^1𝒪^1,c_{s}\colon\widehat{K_{Z_{Z_{e}}(s)}(X^{s}\times X^{s})}^{1}\to\widehat{\mathcal{O}}^{1},

defined analogously to cc, is surjective (recall that X=eX=\mathcal{B}^{{\mathbb{C}}^{*}}_{e}).

7.6.2.

The main idea is to use the Morita equivalence to get another realization of the image of the same map.

To do this, let us denote 𝐊ZZe(s)0(Xs){\bf{K}}_{Z_{Z_{e}}(s)^{0}}(X^{s}) by MM. This MM is a module over R=𝐊ZZe(s)0(pt)R={\bf{K}}_{Z_{Z_{e}}(s)^{0}}(\operatorname{pt}), and (EndR(M))Γes(\operatorname{End}_{R}(M))^{\Gamma_{e}^{s}} can be considered as RΓesR^{\Gamma_{e}^{s}}-algebra.

Then, (cf. Section A.1.5) the LHS of (7.6.2) is isomorphic as an algebra to (EndR^𝔪M^𝔪)Γes(\operatorname{End}_{\widehat{R}^{\mathfrak{m}}}\hat{M}^{\mathfrak{m}})^{\Gamma_{e}^{s}} (indeed, we have a natural Γse\Gamma^{s}_{e}-equivariant homomorphism EndR^𝔪(M^𝔪)𝐊ZZe(s)0(Xs×Xs)^𝔪\operatorname{End}_{\widehat{R}^{\mathfrak{m}}}(\hat{M}^{\mathfrak{m}})\to\widehat{{\bf{K}}_{Z_{Z_{e}}(s)^{0}}(X^{s}\times X^{s})}^{\mathfrak{m}} that becomes an isomorphism at the fiber at 𝔪ZZe(s)0,1\mathfrak{m}_{Z_{Z_{e}}(s)^{0},1} and both of the algebras are free modules over R^𝔪\widehat{R}^{\mathfrak{m}} so our map must be an isomorphism, passing to Γse\Gamma^{s}_{e}-invariants we obtain the desired statement).

Lemma 7.6.2.

The algebra (EndR^𝔪M^𝔪)Γes(\operatorname{End}_{\hat{R}^{\mathfrak{m}}}\hat{M}^{\mathfrak{m}})^{\Gamma_{e}^{s}} is Morita-equivalent to (ϵ(R^𝔪#Γes)ϵ)opp(\epsilon(\hat{R}^{\mathfrak{m}}\#\Gamma_{e}^{s})\epsilon)^{\mathrm{opp}}, where ϵ\epsilon is an idempotent in the group algebra of Γes\Gamma_{e}^{s} corresponding to the set of Γes\Gamma_{e}^{s}-characters appearing in K(Xs)K(X^{s}) (and the completion should be understood in the same sense, as above).

Proof.

Indeed, recall that indecomposable projective modules over R^𝔪#Γes\hat{R}^{\mathfrak{m}}\#\Gamma_{e}^{s} have the form R^𝔪ρ\hat{R}^{\mathfrak{m}}\otimes\rho, where ρ\rho is one of irreducible representations of Γes\Gamma_{e}^{s}. Recall now that M^𝔪R^𝔪𝐊(Xs)\hat{M}^{\mathfrak{m}}\simeq\hat{R}^{\mathfrak{m}}\otimes{\bf{K}}(X^{s}) as R^𝔪#Γes\widehat{R}^{\mathfrak{m}}\#\Gamma_{e}^{s}-module. In particular, M^𝔪\hat{M}^{\mathfrak{m}} is a projective R^𝔪#Γes\hat{R}^{\mathfrak{m}}\#\Gamma_{e}^{s}-module.

Recall also that (EndR^𝔪M^𝔪)Γes=EndR^𝔪#Γes(M^𝔪)(\operatorname{End}_{\hat{R}^{\mathfrak{m}}}\hat{M}^{\mathfrak{m}})^{\Gamma_{e}^{s}}=\operatorname{End}_{\widehat{R}^{\mathfrak{m}}\#\Gamma_{e}^{s}}(\hat{M}^{\mathfrak{m}}), so

HomR^𝔪#Γes(M^𝔪,):R^𝔪#ΓesmodEndR^𝔪#Γes(M^𝔪)oppmod\operatorname{Hom}_{\hat{R}^{\mathfrak{m}}\#\Gamma_{e}^{s}}(\hat{M}^{\mathfrak{m}},-)\colon\widehat{R}^{\mathfrak{m}}\#\Gamma_{e}^{s}-\operatorname{mod}\twoheadrightarrow\operatorname{End}_{\hat{R}^{\mathfrak{m}}\#\Gamma_{e}^{s}}(\hat{M}^{\mathfrak{m}})^{\mathrm{opp}}-\operatorname{mod}

is the quotient functor with the kernel consisting of WR^𝔪#ΓesmodW\in\hat{R}^{\mathfrak{m}}\#\Gamma_{e}^{s}-\operatorname{mod} such that HomR^𝔪#Γes(M,W)=0\operatorname{Hom}_{\widehat{R}^{\mathfrak{m}}\#\Gamma_{e}^{s}}(M,W)=0. We need to check that the kernel consists of WR^𝔪#ΓesmodW\in\hat{R}^{\mathfrak{m}}\#\Gamma_{e}^{s}-\operatorname{mod} such that ϵW=0\epsilon W=0. To see that, it is enough to show that for any representation ρ\rho of Γse\Gamma^{s}_{e} and any R^𝔪#Γes\hat{R}^{\mathfrak{m}}\#\Gamma_{e}^{s}-module WW, we have HomR^𝔪#Γes(R^𝔪ρ,W)=0\operatorname{Hom}_{\hat{R}^{\mathfrak{m}}\#\Gamma_{e}^{s}}(\hat{R}^{\mathfrak{m}}\otimes\rho,W)=0 iff Wρ=0W_{\rho}=0 This is clear since

HomR^𝔪#Γes(R^𝔪ρ,W)=HomΓse(ρ,W).\operatorname{Hom}_{\hat{R}^{\mathfrak{m}}\#\Gamma_{e}^{s}}(\hat{R}^{\mathfrak{m}}\otimes\rho,W)=\operatorname{Hom}_{\Gamma^{s}_{e}}(\rho,W).

It follows from Lemma 7.6.2 that:

(7.6.3) C((EndM^𝔪)Γes))C(ϵ(R^𝔪#Γes)ϵ).C((\operatorname{End}\hat{M}^{\mathfrak{m}})^{\Gamma_{e}^{s}}))\simeq C(\epsilon(\hat{R}^{\mathfrak{m}}\#\Gamma_{e}^{s})\epsilon).

Thus, it follows that for some Γes\Gamma_{e}^{s}-invariant Zariski neighborhood UU of 1SpecR1\in\operatorname{Spec}R, one has

C((End(M|U))Γes)C(ϵ(R#Γes)ϵ|U).C((\operatorname{End}(M|_{U}))^{\Gamma_{e}^{s}})\simeq C(\epsilon(R\#\Gamma_{e}^{s})\epsilon|_{U}).

7.6.3.

It follows from the results of Baranovsky (cf. [2]) that there exists a map R#Γes𝒪(Y1)R\#\Gamma_{e}^{s}\to\mathcal{O}(Y_{1}) which induces an isomorphism C(R#Γes)𝒪(Y1).C(R\#\Gamma_{e}^{s})\,\vphantom{j^{X^{2}}}\smash{\overset{\sim}{\vphantom{\rule{0.0pt}{1.99997pt}}\smash{\longrightarrow}}}\,\mathcal{O}(Y_{1}).

One can construct such an isomorphism (let us call it cs0c_{s}^{0}) as the 0-th graded component of the following chain of equalities (HH\operatorname{HH} corresponds to Hochschild homology):

HHR#Γes(R#Γes)(HHR(R#Γes))Γes(γΓesHHRRγ)Γes.\operatorname{HH}^{R\#\Gamma_{e}^{s}}_{*}(R\#\Gamma_{e}^{s})\simeq(\operatorname{HH}^{R}_{*}(R\#\Gamma_{e}^{s}))_{\Gamma_{e}^{s}}\simeq(\oplus_{\gamma\in\Gamma_{e}^{s}}\operatorname{HH}^{R}_{*}R\gamma)_{\Gamma_{e}^{s}}.

Indeed, we get

HH0R#Γes(R#Γes)(γΓes𝒪(Yγ))Γes(γΓes𝒪(Yγ))Γes.\operatorname{HH}_{0}^{R\#\Gamma_{e}^{s}}(R\#\Gamma_{e}^{s})\simeq(\oplus_{\gamma\in\Gamma_{e}^{s}}\mathcal{O}(Y^{\gamma}))_{\Gamma_{e}^{s}}\simeq(\oplus_{\gamma\in\Gamma_{e}^{s}}\mathcal{O}(Y^{\gamma}))^{\Gamma_{e}^{s}}.

Thus, one sees that the map cs~:R#ΓesC(R#Γes)cs0𝒪(Y1)\widetilde{c_{s}}\colon R\#\Gamma_{e}^{s}\twoheadrightarrow C(R\#\Gamma_{e}^{s})\xrightarrow{c_{s}^{0}}\mathcal{O}(Y_{1}) (where the first arrow is an evident surjection) can be characterized by the following condition: for any rR#Γesr\in R\#\Gamma_{e}^{s}, p(SpecR)/Γesp\in(\operatorname{Spec}R)/\Gamma_{e}^{s},

(7.6.4) cs~(r)|{p}×StabΓes(p),χρ|StabΓes(p)=TrrVpρ,\langle\widetilde{c_{s}}(r)|_{\{p\}\times\operatorname{Stab}_{\Gamma_{e}^{s}}(p)},\chi_{\rho}|_{\operatorname{Stab}_{\Gamma_{e}^{s}}(p)}\rangle=\operatorname{Tr}_{r}{V^{p}_{\rho}},

where Vpρ{V^{p}_{\rho}} is the irreducible R#ΓesR\#\Gamma_{e}^{s}-module corresponding to the pair (p,ρ|StabΓse(p))(p,\rho|_{\operatorname{Stab}_{\Gamma^{s}_{e}}(p)}) (cf. proof of Proposition 7.6.3 below).

We will be interested in cs~|ϵ(R#Γse)ϵ=:cs\widetilde{c_{s}}|_{\epsilon(R\#\Gamma^{s}_{e})\epsilon}=:c_{s}^{\prime}.

Proposition 7.6.3.

The isomorphism of Proposition 7.6.1 induces an isomorphisms between the images of csc_{s} and of cs^1\widehat{c_{s}^{\prime}}^{1}.

Proof.

Irreducible representations of ϵ(R#Γes)ϵ|U\epsilon(R\#\Gamma_{e}^{s})\epsilon|_{U} are easily seen to be in bijection with pairs (p,ρ)(p,\rho), where pUp\in U, ρIrrep(StabΓes(p))\rho\in\operatorname{Irrep}(\operatorname{Stab}_{\Gamma_{e}^{s}}(p)), are such that ρ\rho appears in ϕ|StabΓes(p)\phi|_{\operatorname{Stab}_{\Gamma_{e}^{s}}(p)} for some irreducible representation ϕ\phi of Γes\Gamma_{e}^{s} with ϕ(ϵ)0\phi(\epsilon)\neq 0.

The cocenters of two Morita-equivalent algebras are canonically isomorphic, the isomorphism is compatible with evaluating traces on the corresponding finite-dimensional representations. In view of this observation, Proposition 7.6.3 follows from the previous paragraph together with (7.6.4) and Lemma 7.4.1. ∎

7.6.4.

Let us consider a subring 𝒪𝒪(Y1)\mathcal{O}\subseteq\mathcal{O}(Y_{1}) consisting of functions ff, such that at the every fiber of the projection Y1Y/ΓseY_{1}\to Y/\Gamma^{s}_{e}, they are sums of the restrictions of Γes\Gamma_{e}^{s}-characters lying in the set corresponding to ϵ\epsilon.

Remark 7.6.4.

Note that the notation ‘‘𝒪\mathcal{O}’’ is consistent with the previous one. Namely, our 𝒪^𝔪\hat{\mathcal{O}}^{\mathfrak{m}} is the completion of 𝒪\mathcal{O} at 𝔪\mathfrak{m}.

Clearly, Im(cs)𝒪\operatorname{Im}(c_{s}^{\prime})\subseteq\mathcal{O}. To finish the present subsection, it remains to show that, in fact, equality holds.

Indeed, let us take f𝒪f\in\mathcal{O}. Since 𝒪𝒪(Y1)\mathcal{O}\subseteq\mathcal{O}(Y_{1}), we can rewrite ff as cs~(r)\widetilde{c_{s}}(r) for some rR#Γr\in R\#\Gamma. But then, since f𝒪f\in\mathcal{O}, one has f=cs~(r)=cs~(ϵrϵ)Imcsf=\widetilde{c_{s}}(r)=\widetilde{c_{s}}(\epsilon r\epsilon)\in\operatorname{Im}c_{s}^{\prime}.

Appendix A K-theoretic appendix

Here we summarize general properties of equivariant KK-groups used in the paper. None of this is original; almost everything is taken from [29], from [28], or from [47] (cf. below).

A.1. Completion in equivariant KK-theory

A.1.1.

Let HH be an algebraic group acting on a variety XX (we are not assuming that XX is smooth in this section, but we will actually apply the results of this section only to smooth XX). Pick a semisimple element sHs\in H and set Z:=ZH(s)Z:=Z_{H}(s).

Let γH\gamma\subset H be the conjugacy class of ss. It follows from [28, Theorem 4.3] that there exists the isomorphism of completions

𝐊Z(Xs)^s𝐊H(X)^γ.\widehat{{\bf{K}}_{Z}(X^{s})}^{s}\,\vphantom{j^{X^{2}}}\smash{\overset{\sim}{\vphantom{\rule{0.0pt}{1.99997pt}}\smash{\longrightarrow}}}\,\widehat{{\bf{K}}_{H}(X)}^{\gamma}.

In particular, we have an isomorphism of fibers 𝐊H(X)γ𝐊Z(Xs)s{\bf{K}}_{H}(X)_{\gamma}\simeq{\bf{K}}_{Z}(X^{s})_{s}.

Remark A.1.1.

If XX is smooth, then we have the natural (forgetting the equivariance and restriction) maps:

(A.1.1) 𝐊H(X)𝐊Z(X)𝐊Z(Xs).{\bf{K}}_{H}(X)\to{\bf{K}}_{Z}(X)\to{\bf{K}}_{Z}(X^{s}).

Their composition induces the isomorphism of completions 𝐊H(X)^γ𝐊Z(Xs)^s\widehat{{\bf{K}}_{H}(X)}^{\gamma}\,\vphantom{j^{X^{2}}}\smash{\overset{\sim}{\vphantom{\rule{0.0pt}{1.99997pt}}\smash{\longrightarrow}}}\,\widehat{{\bf{K}}_{Z}(X^{s})}^{s} (see [28, Theorem 4.3 (b)]).

A.1.2.

Now we want to understand 𝐊Z(Xs)s{\bf{K}}_{Z}(X^{s})_{s}. For connected ZZ, it follows from [29, Theorem 1.1] that 𝐊Z(Xs)s𝐊Z(Xs)1𝐊(Xs){\bf{K}}_{Z}(X^{s})_{s}\simeq{\bf{K}}_{Z}(X^{s})_{1}\simeq{\bf{K}}(X^{s}); but this is not true in general, as we have already mentioned in Remark 5.2.1. Nevertheless, Graham explained to us that using the arguments from loc. cit., one can show that 𝐊Z(Xs)s{\bf{K}}_{Z}(X^{s})_{s} surjects onto 𝐊(Xs)π0(Z){\bf{K}}(X^{s})^{\pi_{0}(Z)}; but this would not be enough for us.

A.1.3.

First of all (following [28, Section 5.2]) we identify 𝐊Z(Xs)s{\bf{K}}_{Z}(X^{s})_{s} with 𝐊Z(Xs)1{\bf{K}}_{Z}(X^{s})_{1} as follows. Let us construct the isomorphism 𝐊Z(Xs)s𝐊Z(Xs)1{\bf{K}}_{Z}(X^{s})_{s}\,\vphantom{j^{X^{2}}}\smash{\overset{\sim}{\vphantom{\rule{0.0pt}{1.99997pt}}\smash{\longrightarrow}}}\,{\bf{K}}_{Z}(X^{s})_{1}. Consider a ZZ-equivariant coherent sheaf \mathcal{F} on XsX^{s}. The action of sZs\in Z is trivial on XsX^{s} so the ZZ-equivariant structure on \mathcal{F} induces the decomposition =χχ\mathcal{F}=\bigoplus_{\chi\in\mathbb{C}^{*}}{\mathcal{F}}_{\chi}, where χ{\mathcal{F}}_{\chi}\subset\mathcal{F} is the subsheaf of \mathcal{F} on which ss acts via the multiplication by χ\chi.

The desired isomorphism 𝐊Z(Xs)s𝐊Z(Xs)1{\bf{K}}_{Z}(X^{s})_{s}\,\vphantom{j^{X^{2}}}\smash{\overset{\sim}{\vphantom{\rule{0.0pt}{1.99997pt}}\smash{\longrightarrow}}}\,{\bf{K}}_{Z}(X^{s})_{1} is induced by the automorphism

(A.1.2) []χχ[χ].[\mathcal{F}]\mapsto\sum_{\chi\in\mathbb{C}^{*}}\chi[\mathcal{F}_{\chi}].
Remark A.1.2.

The same argument (cf. loc. cit.) shows that 𝐊Z(Xs)^s𝐊Z(Xs)^1\widehat{{\bf{K}}_{Z}(X^{s})}^{s}\simeq\widehat{{\bf{K}}_{Z}(X^{s})}^{1}.

A.1.4.

Now it remains to describe 𝐊Z(Xs)1{\bf{K}}_{Z}(X^{s})_{1}. We set A:=ZA:=Z, Y:=XsY:=X^{s}. Our goal is to describe 𝐊A(Y)1{\bf{K}}_{A}(Y)_{1}. Let 𝔪A,1𝐊A(pt)\mathfrak{m}_{A,1}\subset{\bf{K}}_{A}(\operatorname{pt}) be the maximal ideal of 11. Recall that we have the restriction homomorphism 𝐊A(pt)𝐊A0(pt){\bf{K}}_{A}(\operatorname{pt})\rightarrow{\bf{K}}_{A^{0}}(\operatorname{pt}) and let (𝔪A,1)𝐊A0(pt)(\mathfrak{m}_{A,1})\subset{\bf{K}}_{A^{0}}(\operatorname{pt}) be the ideal generated by the image of 𝔪A,1\mathfrak{m}_{A,1}.

Remark A.1.3.

Note that (𝔪A,1)π0(A)(\mathfrak{m}_{A,1})^{\pi_{0}(A)} is the maximal ideal in 𝐊A0(pt)π0(A){\bf{K}}_{A^{0}}(\operatorname{pt})^{\pi_{0}(A)} that is equal to the ideal 𝔪A0,1π0(A)\mathfrak{m}_{A^{0},1}^{\pi_{0}(A)}. In particular, (𝔪A,1)(\mathfrak{m}_{A,1}) is equal to the ideal generated by 𝔪A0,1π0(A)\mathfrak{m}_{A^{0},1}^{\pi_{0}(A)} in 𝐊A0(pt){\bf{K}}_{A^{0}}(\operatorname{pt}).

Proposition A.1.4.

We have an isomorphism

𝐊A(Y)/𝔪A,1𝐊A(Y)(𝐊A0(Y)/(𝔪A,1)𝐊A0(Y))π0(A).{\bf{K}}_{A}(Y)/\mathfrak{m}_{A,1}{\bf{K}}_{A}(Y)\simeq({\bf{K}}_{A^{0}}(Y)/(\mathfrak{m}_{A,1}){\bf{K}}_{A^{0}}(Y))^{\pi_{0}(A)}.
Proof.

Consider the induction and restriction functors

res:CohA(Y)CohA0(Y),ind:CohA0(Y)CohA(Y).\mathrm{res}\colon\operatorname{Coh}_{A}(Y)\to\operatorname{Coh}_{A^{0}}(Y),~{}\mathrm{ind}\colon\operatorname{Coh}_{A^{0}}(Y)\to\operatorname{Coh}_{A}(Y).

Functor res\mathrm{res} here just sends CohA(Y)\mathcal{F}\in\operatorname{Coh}_{A}(Y) to itself considered as A0A^{0}-equivariant sheaf.
Functor ind\mathrm{ind} sends 𝒫CohA0(Y)\mathcal{P}\in\operatorname{Coh}_{A^{0}}(Y) to the following sheaf: consider the projection morphism π:A×YY\pi\colon A\times Y\to Y. We have an action of A×A0A\times A^{0} on A×YA\times Y given by (a,a0)(a,y)=(aaa01,a0y)(a,a_{0})\cdot(a^{\prime},y)=(aa^{\prime}a_{0}^{-1},a_{0}y).

Moreover, morphism π\pi is A×A0A\times A^{0}-equivariant (action of AA on YY is trivial); so π(𝒫)\pi^{*}(\mathcal{P}) is A×A0A\times A^{0}-equivariant. The action of A0A^{0} on A×YA\times Y is free, so there exists the unique coherent AA-equivariant sheaf 𝒫~\widetilde{\mathcal{P}} on (A×Y)/A0(A\times Y)/{A^{0}} such that its pullback to A×YA\times Y is isomorphic to π𝒫\pi^{*}\mathcal{P}.

Finally, we define ind(𝒫):=μ𝒫~\mathrm{ind}(\mathcal{P}):=\mu_{*}\widetilde{\mathcal{P}}, where μ:(A×Y)/A0Y\mu\colon(A\times Y)/{A^{0}}\to Y is the natural (AA-equivariant) morphism sending [(a,y)][(a^{\prime},y)] to aya^{\prime}y.

It is clear that, for any \mathcal{F} (resp., any 𝒫\mathcal{P}),

(A.1.3) indres()=[π0(A)],resind(𝒫)=gπ0(A)g𝒫.\mathrm{ind}\circ\mathrm{res}(\mathcal{F})=\mathcal{F}\otimes\mathbb{C}[\pi_{0}(A)],\;\mathrm{res}\circ\mathrm{ind}(\mathcal{P})=\sum_{g\in\pi_{0}(A)}g^{*}\mathcal{P}.

Note that the functors ind,res\mathrm{ind},\,\mathrm{res} are exact so they induce maps [ind],[res][\mathrm{ind}],\,[\mathrm{res}] between 𝐊A(Y){\bf{K}}_{A}(Y) and 𝐊A0(Y)π0(A){\bf{K}}_{A^{0}}(Y)^{\pi_{0}(A)}. Note also, that both [ind][\mathrm{ind}] and [res][\mathrm{res}] are linear for the action of 𝐊A(pt){\bf{K}}_{A}(\operatorname{pt}) (the action of 𝐊A(pt){\bf{K}}_{A}(\operatorname{pt}) on 𝐊A0(Y)π0(A){\bf{K}}_{A^{0}}(Y)^{\pi_{0}(A)} is via the restriction homomorphism 𝐊A(pt)𝐊A0(pt)π0(A){\bf{K}}_{A}(\operatorname{pt})\rightarrow{\bf{K}}_{A^{0}}(\operatorname{pt})^{\pi_{0}(A)}), so (passing to the fiber at 𝔪A,1\mathfrak{m}_{A,1}) we get maps in both directions between 𝐊A(Y)/𝔪1𝐊A(Y){\bf{K}}_{A}(Y)/\mathfrak{m}_{1}{\bf{K}}_{A}(Y) and (𝐊A0(Y)/(𝔪A,1)𝐊A0(Y))π0(A)({\bf{K}}^{A^{0}}(Y)/(\mathfrak{m}_{A,1}){\bf{K}}^{A^{0}}(Y))^{\pi_{0}(A)}. We see that restriction of both [ind][res][\mathrm{ind}]\circ[\mathrm{res}], [res][ind][\mathrm{res}]\circ[\mathrm{ind}] to

𝐊A(Y)/𝔪1𝐊A(Y),(𝐊A0(Y)/(𝔪A,1)𝐊A0(Y))π0(A){\bf{K}}_{A}(Y)/\mathfrak{m}_{1}{\bf{K}}_{A}(Y),\,({\bf{K}}^{A^{0}}(Y)/(\mathfrak{m}_{A,1}){\bf{K}}^{A^{0}}(Y))^{\pi_{0}(A)}

is just the multiplication by |π0(A)||\pi_{0}(A)| so we obtain the desired isomorphism. ∎

A.1.5.

Now let us note that 𝐊A0(Y)/(𝔪A,1)𝐊A0(Y){\bf{K}}_{A^{0}}(Y)/(\mathfrak{m}_{A,1}){\bf{K}}_{A^{0}}(Y) is a finitely generated module over R=𝐊A0(pt)/(𝔪A,1)R={\bf{K}}_{A^{0}}(\operatorname{pt})/({\mathfrak{m}}_{A,1}), and its fiber over one is 𝐊(Y){\bf{K}}(Y) (use [29, Theorem 1.1]).

Moreover, the ring RR is local commutative, so every flat module over it is free.

We conclude that if 𝐊A0(Y)/(𝔪A,1){\bf{K}}_{A^{0}}(Y)/(\mathfrak{m}_{A,1}) is flat over RR (that happens, for example, if 𝐊A0(Y)^1\widehat{{\bf{K}}_{A^{0}}(Y)}^{1} is flat over 𝐊A0(pt)^1\widehat{{\bf{K}}_{A^{0}}(\operatorname{pt})}^{1} ), then we have a (non canonical) π0(A)\pi_{0}(A)-equivariant isomorphism:

(A.1.4) 𝐊A0(Y)/(𝔪A,1)𝐊A0(Y)𝐊(Y)R.{\bf{K}}_{A^{0}}(Y)/(\mathfrak{m}_{A,1}){\bf{K}}_{A^{0}}(Y)\simeq{\bf{K}}(Y)\otimes R.

Passing to π0(A)\pi_{0}(A)-invariants and using Proposition A.1.4 we get the isomorphism:

(A.1.5) 𝐊A(Y)/𝔪A,1𝐊A(Y)(𝐊(Y)R)π0(A).{\bf{K}}_{A}(Y)/\mathfrak{m}_{A,1}{\bf{K}}_{A}(Y)\simeq({\bf{K}}(Y)\otimes R)^{\pi_{0}(A)}.
Remark A.1.5.

The situation is similar to the one in Remark A.1.2: namely, under the same assumptions as above, one gets

(A.1.6) 𝐊A(Y)^𝔪1(𝐊(Y)𝐊A0(pt)^(𝔪A,1))π0(A).\widehat{{\bf{K}}_{A}(Y)}^{\mathfrak{m}_{1}}\simeq({\bf{K}}(Y)\otimes\widehat{{\bf{K}}_{A^{0}}(\operatorname{pt})}^{(\mathfrak{m}_{A,1})})^{\pi_{0}(A)}.
Remark A.1.6.

Our ‘‘local flatness assumption’’ holds for Y=e,sY=\mathcal{B}_{e}^{\mathbb{C}^{*},s} and A=ZZe(s)A=Z_{Z_{e}}(s), to see this recall first that if T(s)ZZe(s)T(s)\subset Z_{Z_{e}}(s) is a maximal torus then 𝐊T(s)(e,s){\bf{K}}_{T(s)}(\mathcal{B}_{e}^{\mathbb{C}^{*},s}) is free over 𝐊T(s)(pt){\bf{K}}_{T(s)}(\operatorname{pt}) (same argument as in [44, Lemma 1.10]). Now, Let ZZ be the cover of ZZe(s)0Z_{Z_{e}}(s)^{0}, so that [Z,Z][Z,Z] is simply connected. Let TZT\subset Z be the preimage of T(s)T(s) in ZZ. It follows from Proposition A.1.7 that we have 𝐊ZZe(s)0(X)^1𝐊Z(X)^1\widehat{{\bf{K}}_{Z_{Z_{e}}(s)^{0}}(X)}^{1}\simeq\widehat{{\bf{K}}_{Z}(X)}^{1} so it is enough to check that 𝐊Z(X)^1\widehat{{\bf{K}}_{Z}(X)}^{1} is flat over 𝐊Z(pt)^1\widehat{{\bf{K}}_{Z}(\operatorname{pt})}^{1}. Recall now that by [47, 7.1] we have the isomorphism 𝐊T(X)𝐊Z(X)𝐊Z(pt)𝐊T(pt){\bf{K}}_{T}(X)\simeq{\bf{K}}_{Z}(X)\otimes_{{\bf{K}}_{Z}(\operatorname{pt})}{\bf{K}}_{T}(\operatorname{pt}) so it remains to check that 𝐊T(X)^1\widehat{{\bf{K}}_{T}(X)}^{1} is flat over 𝐊T(pt)^1\widehat{{\bf{K}}_{T}(\operatorname{pt})}^{1} (we use fpqc descent here). This again follows from the identification 𝐊T(X)^1𝐊T(s)(X)^1\widehat{{\bf{K}}_{T}(X)}^{1}\simeq\widehat{{\bf{K}}_{T(s)}(X)}^{1}.

A.1.6. Passing to a covering

Suppose that GG is a connected reductive group. Let π:HG\pi\colon H\rightarrow G be the covering of GG so that [H,H][H,H] is simply connected. Let Γ\Gamma be the (finite central) kernel of π\pi.

Suppose also that HH acts on some (smooth) algebraic variety XX, and that Γ\Gamma lies in the kernel of this action. Recall that Spec(𝐊H(pt))=H//adH\operatorname{Spec}({\bf{K}}_{H}(\operatorname{pt}))=H/\!/_{\operatorname{ad}}H, Spec(𝐊𝐆(pt))=𝐆//ad𝐆\operatorname{Spec}(\bf{}K_{G}(\operatorname{pt}))=G/\!/_{\operatorname{ad}}G, and the the natural map H//adHG//adGH/\!/_{\operatorname{ad}}H\to G/\!/_{\operatorname{ad}}G is an étale morphism. Let UHH//adHU_{H}\subset H/\!/_{\operatorname{ad}}H be an open neighbourhood of 1H//adH1\in H/\!/_{\operatorname{ad}}H such that the composition UH//adHG//adGU\subset H/\!/_{\operatorname{ad}}H\to G/\!/_{\operatorname{ad}}G is an isomorphism onto the image that we denote by UGG//adGU_{G}\subset G/\!/_{\operatorname{ad}}G.

Proposition A.1.7.
(A.1.7) 𝐊H(X)|UH𝐊G(X)|UG.{\bf{K}}_{H}(X)|_{U_{H}}\simeq{\bf{K}}_{G}(X)|_{U_{G}}.
Proof.

Let us note that the formula (A.1.2) (for ss being some γΓ\gamma\in\Gamma) gives a well-defined action of Γ\Gamma on 𝐊H(X){\bf{K}}_{H}(X).

Step 1. We claim that 𝐊G(X)=𝐊H(X)Γ{\bf{K}}_{G}(X)={\bf{K}}_{H}(X)^{\Gamma}. Indeed, since any HH-equivariant sheaf \mathcal{F} has a direct sum decomposition =χCharΓχ\mathcal{F}=\bigoplus_{\chi\in\operatorname{Char}\Gamma}\mathcal{F}_{\chi}, there is a direct sum decomposition CohH(X)χCharΓCohH(X)χ\operatorname{Coh}_{H}(X)\simeq\bigoplus_{\chi\in\operatorname{Char}\Gamma}\operatorname{Coh}_{H}(X)_{\operatorname{\chi}}; – and 𝐊G(X)K0(CohH(X)triv)𝐊H(X)Γ{\bf{K}}_{G}(X)\simeq K_{0}(\operatorname{Coh}_{H}(X)_{\operatorname{triv}})\otimes_{\mathbb{Z}}\mathbb{C}\simeq{\bf{K}}_{H}(X)^{\Gamma}. We now need to compare the restrictions KH(X)|UHK_{H}(X)|_{U_{H}}, (KH(X)Γ)|UG(K_{H}(X)^{\Gamma})|_{U_{G}}.

Step 2. Set K:=𝐊H(X)K:={\bf{K}}_{H}(X) and L:=KΓL:=K^{\Gamma}. Let us consider the restriction KΓUHK_{\Gamma U_{H}} of KK to the open set γΓγU\bigsqcup_{\gamma\in\Gamma}\gamma U. Note that L|UG=(KΓUH)ΓL|_{U_{G}}=(K_{\Gamma U_{H}})^{\Gamma}, so our goal is to identify K|UHK|_{U_{H}} with (KΓUH)Γ(K_{\Gamma U_{H}})^{\Gamma}.

The identification of K|γUHK|_{\gamma U_{H}} with K|UHK|_{U_{H}} from (A.1.2) gives a natural Γ\Gamma-equivariant isomorphism:

KΓU=γΓK|γU=K|UH|Γ|,K_{\Gamma U}=\bigoplus_{\gamma\in\Gamma}K|_{\gamma U}=K|_{U_{H}}^{\oplus|\Gamma|},

where Γ\Gamma acts on the RHS via permuting the factors. It follows that K|UHK|_{U_{H}} is naturally isomorphic to (KΓUH)Γ(K_{\Gamma U_{H}})^{\Gamma}.

Remark A.1.8.

As a corollary of Proposition A.1.7 we conclude that 𝐊H(X)^1𝐊G(X)^1\widehat{{\bf{K}}_{H}(X)}^{1}\simeq\widehat{{\bf{K}}_{G}(X)}^{1}.

A.2. Modules over convolution algebras

Let GG be a (possibly, disconnected) algebraic group acting on a smooth variety XX. Let us now assume that GG is reductive, and that the cycle morphisms A(Xs)H(Xs,)A_{*}(X^{s})\otimes_{\mathbb{Z}}\mathbb{C}\to H_{*}(X^{s},{\mathbb{C}}), A(Xs×Xs)H(Xs×Xs,)A_{*}(X^{s}\times X^{s})\otimes_{\mathbb{Z}}\mathbb{C}\to H_{*}(X^{s}\times X^{s},\mathbb{C}), are isomorphisms for every semisimple sGs\in G.

(This is true, for example, for X=eX=\mathcal{B}_{e}^{\mathbb{C}^{*}} (see our discussion around the formula (5.1.4).)

We also assume that 𝐊G0(X){\bf{K}}_{G^{0}}(X) is flat over 𝐊G0(pt){\bf{K}}_{G^{0}}(\operatorname{pt}) in some neighbourhood of 11. We claim that the following corollary of the above results holds.

Proposition A.2.1.

(a)(a) Every simple module over the algebra 𝐊G(X×X){\bf{K}}_{G}(X\times X) is of the form 𝐊(Xs)ρ{\bf{K}}(X^{s})_{\rho} for some semisimple sGs\in G and an irreducible ρIrrep(ZG(s))\rho\in\operatorname{Irrep}(Z_{G}(s)).

(b)(b) Modules K(Xs)ρK(X^{s})_{\rho}, K(Xs)ρK(X^{s^{\prime}})_{{\rho^{\prime}}} are isomorphic iff (s,ρ)(s,\rho), (s,ρ)(s^{\prime},\rho^{\prime}) lie in the same conjugacy class.

Note that similar results for GG being a finite group go back to Lusztig.

Proof.

Step 1. Set Z:=ZG(s)Z:=Z_{G}(s). Note, first of all, that there exists a natural morphism ϕ:𝐊Z(Xs)𝐊Z(Xs)\phi\colon{\bf{K}}_{Z}(X^{s})\to{\bf{K}}_{Z}(X^{s}) defined analogously to the formula (A.1.2). One can now consider a composition of ϕ\phi and the forgetful morphism; we will call this map Φ\Phi.

One has Φ:𝐊Z(Xs)𝐊(Xs)\Phi\colon{\bf{K}}_{Z}(X^{s})\to{\bf{K}}(X^{s}); it is easy to see that it induces a morphism 𝐊Z(Xs)s𝐊(Xs){\bf{K}}_{Z}(X^{s})_{s}\to{\bf{K}}(X^{s}).

Let λ(s)\lambda(s) be a ZZ-equivariant Thom class of a normal bundle to ss-fixed points inside XX; let λs\lambda_{s} be Φ(λ(s)1)\Phi(\lambda(s)^{-1}).

Let γsG\gamma_{s}\subset G be the conjugacy class of ss. One can consider a morphism

ϕ:𝐊G(X×X)γs𝐊Z(Xs×Xs)sϕ12𝐊Z(Xs×Xs)1𝐊(Xs×Xs)1λs1𝐊(Xs×Xs).\phi^{\prime}\colon{\bf{K}}_{G}(X\times X)_{\gamma_{s}}\to{\bf{K}}_{Z}(X^{s}\times X^{s})_{s}\xrightarrow{\phi_{12}}{\bf{K}}_{Z}(X^{s}\times X^{s})_{1}\to{\bf{K}}(X^{s}\times X^{s})\xrightarrow{1\boxtimes\lambda_{s}^{-1}}{\bf{K}}(X^{s}\times X^{s}).

Here the first arrow is a natural restriction morphism ii^{*} for i:Xs×XsX×Xi\colon X^{s}\times X^{s}\to X\times X, the second one is analogous to ϕ\phi, and the third one is induced by λs\lambda_{s} via the Kunneth formula.

We claim that ϕ\phi^{\prime} is well-defined and is a homomorphism of algebras.

Step 2. To prove this, it would be enough to show that the morphism

ϕ~:𝐊G(X×X)i𝐊Z(Xs×Xs)1λ(s)1𝐊Z(Xs×Xs)𝐊Z(Xs×Xs)s\tilde{\phi}\colon{\bf{K}}_{G}(X\times X)\xrightarrow{i^{*}}{\bf{K}}_{Z}(X^{s}\times X^{s})\xrightarrow{1\boxtimes\lambda(s)^{-1}}{\bf{K}}_{Z}(X^{s}\times X^{s})\to{\bf{K}}_{Z}(X^{s}\times X^{s})_{s}

is a well-defined homomorphism of algebras.

This follows, analogously to the proofs of [20, 5.11.7] and [20, 5.11.10], from [28, Theorem 4.3].

Step 3. Moreover, by (A.1.5), one can see that the image of ϕ\phi^{\prime} is a semisimple subalgebra 𝐊(Xs×Xs)Γ{\bf{K}}(X^{s}\times X^{s})^{\Gamma} (where Γ\Gamma is the component group of ZZ), and that ϕ\phi^{\prime} can be identified with the morphism 𝐊G(X×X)γs=:KK/Rad(K){\bf{K}}_{G}(X\times X)_{\gamma_{s}}=:K\twoheadrightarrow K/\operatorname{Rad}(K). ∎

A.3. Restriction of equivariance in KK-theory.

We start with the following standard lemma.

Lemma A.3.1.

Let ι:TC\iota\colon T\hookrightarrow C be an embedding of algebraic tori (over \mathbb{C}). Then there exists a collection of characters χi:C()\chi_{i}\colon C\rightarrow(\mathbb{C}^{*}), i=1,,dimCdimTi=1,\ldots,\operatorname{dim}C-\operatorname{dim}T such that

(A.3.1) T=i=1dimCdimTkerχiT=\bigcap_{i=1}^{\operatorname{dim}C-\operatorname{dim}T}\operatorname{ker}\chi_{i}

and i=1kkerχi\bigcap_{i=1}^{k}\operatorname{ker}\chi_{i} is a torus of dimension dimCk\operatorname{dim}C-k for every k=1,,dimCdimTk=1,\ldots,\operatorname{dim}C-\operatorname{dim}T.

Proof.

Let ΩC\Omega_{C}, ΩT\Omega_{T} be the character lattices of CC, TT. The embedding ι\iota induces the surjection ι:ΩCΩT\iota^{*}\colon\Omega_{C}\twoheadrightarrow\Omega_{T} (see, for example, [57, Corollary 22.5.4 (iii)]). The kernel of ι\iota^{*} is free of rank dimCdimT\operatorname{dim}C-\operatorname{dim}T. Let χ1,,χdimCdimT\chi_{1},\ldots,\chi_{\operatorname{dim}C-\operatorname{dim}T} be the generators of this kernel. It follows from the definitions that the equality (A.3.1) holds. It remains to note that for every k=1,,dimCdimTk=1,\ldots,\operatorname{dim}C-\operatorname{dim}T, the quotient ΩC/Span(χ1,,χk)\Omega_{C}/\operatorname{Span}_{\mathbb{Z}}(\chi_{1},\ldots,\chi_{k}) is torsion-free of rank dimCk\operatorname{dim}C-k. Indeed, let ν¯1,,ν¯dimT\bar{\nu}_{1},\ldots,\bar{\nu}_{\operatorname{dim}T} be any generators of the lattice ΩT\Omega_{T}. Let ν1,,νdimTΩC\nu_{1},\ldots,\nu_{\operatorname{dim}T}\in\Omega_{C} be such that ι(νi)=ν¯i\iota^{*}(\nu_{i})=\bar{\nu}_{i}. It follows from the definitions that classes of ν1,,νdimT,χk+1,,χdimCdimT\nu_{1},\ldots,\nu_{\operatorname{dim}T},\chi_{k+1},\ldots,\chi_{\operatorname{dim}C-\operatorname{dim}T} freely generate ΩC/Span(χ1,,χk)\Omega_{C}/\operatorname{Span}_{\mathbb{Z}}(\chi_{1},\ldots,\chi_{k}). ∎

Let ι:HG\iota\colon H\hookrightarrow G be an embedding of connected reductive algebraic groups with simply connected derived subgroups.

We would like to discuss the following proposition. (The similar statement was proved in [47] under different assumptions.)

Proposition A.3.2.

Assume that GG acts on a smooth variety XX. Then, the natural restriction morphism provides an isomorphism:

(A.3.2) 𝐊G(X)𝐊G(pt)𝐊H(pt)𝐊H(X).{\bf{K}}_{G}(X)\otimes_{{\bf{K}}_{G}(\operatorname{pt})}{\bf{K}}_{H}(\operatorname{pt})\,\vphantom{j^{X^{2}}}\smash{\overset{\sim}{\vphantom{\rule{0.0pt}{1.99997pt}}\smash{\longrightarrow}}}\,{\bf{K}}_{H}(X).
Remark A.3.3.

Note that we do not make any assumptions concerning properness of XX, nor existence of an equivariant affine paving.

Proof.

Step 1. Let us reduce the situation to the case of HH and GG being tori.

Namely, let TCT\hookrightarrow C be a pair of compatible maximal tori of HH, GG.

Note that, by [47, Proposition 31], one has the natural restriction isomorphism

(A.3.3) 𝐊C(X)𝐊G(X)𝐊G(pt)𝐊C(pt).{\bf{K}}_{C}(X)\simeq{\bf{K}}_{G}(X)\otimes_{{\bf{K}}_{G}(\operatorname{pt})}{\bf{K}}_{C}(\operatorname{pt}).

In particular, for WGW_{G} being a Weyl group of GG, 𝐊G(X)=𝐊C(X)WG{\bf{K}}_{G}(X)={\bf{K}}_{C}(X)^{W_{G}} (here we use that 𝐊G(pt)=[G]G[T]W{\bf{K}}_{G}(\operatorname{pt})=\mathbb{C}[G]^{G}\,\vphantom{j^{X^{2}}}\smash{\overset{\sim}{\vphantom{\rule{0.0pt}{1.99997pt}}\smash{\longrightarrow}}}\,\mathbb{C}[T]^{W}, see, for example, [52, Theorem 4]) and similarly 𝐊H(X)=𝐊T(X)WH{\bf{K}}_{H}(X)={\bf{K}}_{T}(X)^{W_{H}}.

The assertion of the Proposition in the toric case reads:

𝐊T(X)𝐊C(X)𝐊C(pt)𝐊T(pt);{\bf{K}}_{T}(X)\simeq{\bf{K}}_{C}(X)\otimes_{{\bf{K}}_{C}(\operatorname{pt})}{\bf{K}}_{T}(\operatorname{pt});

now we deduce:

𝐊H(X)𝐊T(X)WH(𝐊C(X)𝐊C(pt)𝐊T(pt))WH(𝐊G(X)𝐊G(pt)𝐊C(pt)𝐊C(pt)𝐊T(pt))WH==(𝐊G(X)𝐊G(pt)𝐊T(pt))WH𝐊G(X)𝐊G(pt)𝐊H(pt).{\bf{K}}_{H}(X)\simeq{\bf{K}}_{T}(X)^{W_{H}}\simeq({\bf{K}}_{C}(X)\otimes_{{\bf{K}}_{C}(\operatorname{pt})}{\bf{K}}_{T}(\operatorname{pt}))^{W_{H}}\simeq({\bf{K}}_{G}(X)\otimes_{{\bf{K}}_{G}(\operatorname{pt})}{\bf{K}}_{C}(\operatorname{pt})\otimes_{{\bf{K}}_{C}(\operatorname{pt})}{\bf{K}}_{T}(\operatorname{pt}))^{W_{H}}=\\ =({\bf{K}}_{G}(X)\otimes_{{\bf{K}}_{G}(\operatorname{pt})}{\bf{K}}_{T}(\operatorname{pt}))^{W_{H}}\simeq{\bf{K}}_{G}(X)\otimes_{{\bf{K}}_{G}(\operatorname{pt})}{\bf{K}}_{H}(\operatorname{pt}).

Step 2. Now we can assume that HH is a subtorus inside some torus GG.

Let us change the notation. T:=HT:=H, C:=GC:=G. Using Lemma A.3.1 and the induction on dimCdimT\operatorname{dim}C-\operatorname{dim}T, we reduce to the case T=kerχT=\operatorname{ker}\chi for some (primitive) nonzero character χ:C\chi\colon C\rightarrow\mathbb{C}^{*}.

Lemma A.3.4.

The natural restriction map res:𝐊C(X)𝐊C(pt)𝐊T(pt)𝐊T(X)\operatorname{res}\colon{\bf{K}}_{C}(X)\otimes_{{\bf{K}}_{C}(\operatorname{pt})}{\bf{K}}_{T}(\operatorname{pt})\to{\bf{K}}_{T}(X) is surjective.

Proof.

The more powerful statement is proven in [47, Proposition 26]. The proof is as follows.

In fact, we have to prove that the restriction map r:𝐊C(X)𝐊T(X)r\colon{\bf{K}}_{C}(X)\to{\bf{K}}_{T}(X) is surjective.

Let us rewrite it. Note that the character χ\chi, above turns 𝔸1\mathbb{A}^{1} and \mathbb{C}^{*} into CC-varieties. Thus, by [47, Corollaries 5, 12 and Theorem 8], we deduce that rr can be rewritten as:

𝐊C(X)𝐊C(X×𝔸1)𝐊C(X×)=𝐊C(X×C/T)𝐊T(X).{\bf{K}}_{C}(X)\simeq{\bf{K}}_{C}(X\times\mathbb{A}^{1})\twoheadrightarrow{\bf{K}}_{C}(X\times\mathbb{C}^{*})={\bf{K}}_{C}(X\times C/T)\simeq{\bf{K}}_{T}(X).

The second arrow is a surjective restriction to an open subset (cf. [47, Corollary 27]), hence the lemma follows. ∎

Step 3. What remains is to prove injectivity of res\operatorname{res}. This is contained in [47, Corollary 27].

More precisely, we can write down the following piece of the long exact sequence in KK-theory:

𝐊C(X)𝐊C(X×𝔸1)𝐊C(X×)0.{\bf{K}}_{C}(X)\to{\bf{K}}_{C}(X\times\mathbb{A}^{1})\to{\bf{K}}_{C}(X\times\mathbb{C}^{*})\to 0.

One has to prove that the leftmost map can be identified with the multiplication by 1χKT(pt)1-\chi\in K_{T}(\operatorname{pt}).

But the latter is evident from the projection formula: cf. loc. cit.

A.4. Traces in KK-theory

Suppose that XX is a smooth proper variety with the same assumptions as in Section A.2.

Then, by the Kunneth formula and Poincaré duality,

(A.4.1) 𝐊(X×X)𝐊(X)𝐊(X)𝐊(X)𝐊(X)End𝐊(X).{\bf{K}}(X\times X)\simeq{\bf{K}}(X)\otimes{\bf{K}}(X)\simeq{\bf{K}}(X)\otimes{\bf{K}}(X)^{*}\simeq\operatorname{End}{\bf{K}}(X).

This gives a natural action of 𝐊(X×X){\bf{K}}(X\times X) on 𝐊(X){\bf{K}}(X). Moreover, it is well-known that the resulting algebra structure on 𝐊(X×X){\bf{K}}(X\times X) coincides with the one defined via the convolution product.

Let Eu\operatorname{Eu} stand for the Euler characteristic.

Proposition A.4.1.

For any 𝐊(X×X)\mathcal{F}\in{\bf{K}}(X\times X), Tr(𝐊(X))=Eu(𝕃𝒪Δ)\operatorname{Tr}_{\mathcal{F}}({\bf{K}}(X))=\operatorname{Eu}(\mathcal{F}\otimes^{\mathbb{L}}\mathcal{O}_{\Delta}) where 𝒪Δ\mathcal{O}_{\Delta} stands for the structure sheaf of diagonal.

Proof.

Let us denote the linear functional Eu(𝕃𝒪Δ)\mathcal{F}\mapsto\operatorname{Eu}(\mathcal{F}\otimes^{\mathbb{L}}\mathcal{O}_{\Delta}) by τ\tau.

Since 𝒪Δ\mathcal{O}_{\Delta} corresponds to the identity operator under (A.4.1), to show that τ=Tr(𝐊(X))\tau=\operatorname{Tr}_{-}({\bf{K}}(X)), one has to establish that (cf. [32]):

1) τ(𝒢)=τ(𝒢)\tau(\mathcal{F}*\mathcal{G})=\tau(\mathcal{G}*\mathcal{F}) where * stands for the convolution;

2) τ(𝒪Δ)=dim𝐊(X)\tau(\mathcal{O}_{\Delta})=\operatorname{dim}{\bf{K}}(X).

Step 1. 1) is checked by diagram chase as follows.

Let p12p_{12}, p23p_{23}, and p13p_{13} be the projections to the corresponding pairs of the arguments from X×X×XX\times X\times X. Let iΔi_{\Delta} be the embedding of the diagonal into X×XX\times X. Let pp denote a projection to a point.

Then one has:

(A.4.2) τ(𝒢)=piΔp13(p12𝕃p23𝒢)=p(𝕃σ𝒢),\tau(\mathcal{F}*\mathcal{G})=p_{*}i_{\Delta}^{*}p_{13*}(p_{12}^{*}\mathcal{F}\otimes^{\mathbb{L}}p_{23}^{*}\mathcal{G})=p_{*}(\mathcal{F}\otimes^{\mathbb{L}}\sigma^{*}\mathcal{G}),

where all of the functors are derived and σ\sigma is the involution permuting the factors in X×XX\times X (the second equality is explained in more detail in the analogous calculation (7.4.3)).

Equality 1) is now clear.

Step 2. We have to show that the Euler characteristic of the derived self-intersection of the diagonal copy of XX inside X×XX\times X is equal to the dimension of H(X)H^{*}(X).

This follows easily from the Hodge theorem and the fact that Tori(𝒪Δ,𝒪Δ)=iΔΩXi\operatorname{Tor}^{i}(\mathcal{O}_{\Delta},\mathcal{O}_{\Delta})=i_{\Delta*}\Omega_{X}^{i}, since XX is smooth (for more details see [17] and references therein). ∎

A.4.1. The map TT

Here we introduce a trace functor TT that plays a key role in the last section of the main text, and prove that it is actually a trace (commutator) functor.

We consider an action of an arbitrary algebraic group GG on a smooth variety XX.

Definition A.4.2.

Consider the diagram X×XG×XGX\times X\leftarrow G\times X\twoheadrightarrow G, where the first map is the action map, the second one is the projection.

The functor T:Db(QCohG(X×X))Db(QCohG(G))T\colon D^{b}(\operatorname{QCoh}^{G}(X\times X))\to D^{b}(\operatorname{QCoh}^{G}(G)) is defined to be the map given by this correspondence.

Lemma A.4.3.

For any 𝒜,Db(CohG(X×X)){\mathcal{A}},{\mathcal{B}}\in D^{b}(\operatorname{Coh}_{G}(X\times X)) we have a canonical isomorphism T(𝒜)T(𝒜)T(\mathcal{A}*\mathcal{B})\simeq T(\mathcal{B}*\mathcal{A}) where * stands for the convolution product.

Proof.

To fix ideas we first produce an isomorphism of fibers.

The fiber of T(𝒜)T(\mathcal{A}*\mathcal{B}) at gGg\in G is precisely RΓ(Rιg(π12𝒜π23))R\Gamma(R\iota_{g}^{*}(\pi_{12}^{*}\mathcal{A}\otimes\pi_{23}^{*}\mathcal{B})) for ιg\iota_{g} being the embedding of the twisted diagonal {(gx,y,x)|gG}\{(gx,y,x)\ |\ g\in G\} into X3X^{3}. One has:

T(𝒜)g=RΓ({(gx,y,x)},𝒜)=RΓ({(gx,y,y,x)},𝒜)=RΓ({(y,x,gx,y)},𝒜)=T(\mathcal{A}*\mathcal{B})_{g}=R\Gamma(\{(gx,y,x)\},\mathcal{A}*\mathcal{B})=R\Gamma(\{(gx,y,y,x)\},\mathcal{A}\boxtimes\mathcal{B})=R\Gamma(\{(y,x,gx,y)\},\mathcal{B}\boxtimes\mathcal{A})=
=RΓ({(gy,gx,gx,y)},𝒜)=RΓ({(gy,t,t,y)},𝒜)=T(𝒜)g.=R\Gamma(\{(gy,gx,gx,y)\},\mathcal{B}\boxtimes\mathcal{A})=R\Gamma(\{(gy,t,t,y)\},\mathcal{B}\boxtimes\mathcal{A})=T(\mathcal{B}*\mathcal{A})_{g}.

Here we have performed a change of the variable, and have used the GG-equivariance of \mathcal{B}. It is easy to see that the identification of fibers νg:T(𝒜)gT(𝒜)g\nu_{g}\colon T(\mathcal{A}*\mathcal{B})_{g}\,\vphantom{j^{X^{2}}}\smash{\overset{\sim}{\vphantom{\rule{0.0pt}{1.99997pt}}\smash{\longrightarrow}}}\,T(\mathcal{B}*\mathcal{A})_{g} as above is GG-equivariant.

We now present a modification of the above argument that works in families and yields the isomorphism of objects in Db(QCohG(G))D^{b}(\operatorname{QCoh}^{G}(G)).

Step 1. Consider the universal twisted diagonal

ΔG:G×X×XG×X3,(g,x,y)(g,gx,y,x).\Delta_{G}\colon G\times X\times X\to G\times X^{3},(g,x,y)\mapsto(g,gx,y,x).

Variety G×X2G\times X^{2} is equipped with a natural morphism πG:G×X2G\pi_{G}\colon G\times X^{2}\twoheadrightarrow G. By the definition,

T(𝒜)=πGΔG(π12𝒜π23)=πGΔG(γπ12𝒜σπ23),T(\mathcal{A}*\mathcal{B})=\pi_{G*}\Delta_{G}^{*}(\pi_{12}^{*}\mathcal{A}\otimes\pi_{23}^{*}\mathcal{B})=\pi_{G*}\Delta_{G}^{*}(\gamma^{*}\pi_{12}^{*}\mathcal{A}\otimes\sigma^{*}\pi_{23}^{*}\mathcal{B}),

where, as usual, all of the functors are derived, 𝒜\mathcal{A} and \mathcal{B} stand (by a slight abuse of notation) for the corresponding pull-backs under the second projection G×X2X2G\times X^{2}\to X^{2}, πij:G×X××XX×X\pi_{ij}\colon G\times X\times\ldots\times X\to X\times X is the ijij-th projection, γ:(g,x,y)(g,gx,y)\gamma\colon(g,x,y)\to(g,gx,y), σ:(g,x,y)(g,y,x)\sigma\colon(g,x,y)\to(g,y,x). (Note the similarity between this calculation and the formula (A.4.2).)

We can rewrite the last isomorphism using that \otimes is the restriction of \boxtimes to the diagonal: for Δ:G×X×XG×X4\Delta\colon G\times X\times X\hookrightarrow G\times X^{4} being the diagonal embedding,

T(𝒜)=πGΔ(γπ12𝒜σπ23)=πGΔ(σπ23γπ12𝒜).T(\mathcal{A}*\mathcal{B})=\pi_{G*}\Delta^{*}(\gamma^{*}\pi_{12}^{*}\mathcal{A}\boxtimes\sigma^{*}\pi_{23}^{*}\mathcal{B})=\pi_{G*}\Delta^{*}(\sigma^{*}\pi_{23}^{*}\mathcal{B}\boxtimes\gamma^{*}\pi_{12}^{*}\mathcal{A}).

Step 2. In other words, for ΔG,4:G×X2X4,\Delta_{G,4}:G\times X^{2}\to X^{4}, (g,x,y)(g,gx,y,y,x)(g,x,y)\mapsto(g,gx,y,y,x),

T(𝒜)=πGΔG,4(𝒜)=πGΔG,4~(𝒜).T(\mathcal{A}*\mathcal{B})=\pi_{G*}\Delta_{G,4}^{*}(\mathcal{A}\boxtimes\mathcal{B})=\pi_{G*}\widetilde{\Delta_{G,4}}^{*}(\mathcal{B}\boxtimes\mathcal{A}).

for ΔG,4~\widetilde{\Delta_{G,4}} being {(g,y,x,gx,y)}\{(g,y,x,gx,y)\} as in the Step 1.

(We one more time abuse notation by writing 𝒜\mathcal{A}\boxtimes\mathcal{B} for the corresponding pullback from X4X^{4} to G×X4G\times X^{4}.)

We now proceed as in the fiberwise argument above.

Namely, let δG,4\delta_{G,4} be the embedding G×X2X4G\times X^{2}\to X^{4}, (g,x,y)(g,gy,gx,gx,y)(g,x,y)\mapsto(g,gy,gx,gx,y). By the GG-equivariance of \mathcal{B}, we get that T(𝒜)=πGδG,4(𝒜)T(\mathcal{A}*\mathcal{B})=\pi_{G*}\delta_{G,4}^{*}(\mathcal{B}\boxtimes\mathcal{A}).

Moreover, let θ\theta be the automorphism (g,a,b,c,d)(g,a,g1b,g1c,d)(g,a,b,c,d)\to(g,a,g^{-1}b,g^{-1}c,d). Since θ\theta is the GG-equivariant automorphism (GG acts on itself via conjugation), and πGθ=πG\pi_{G}\theta=\pi_{G},

T(𝒜)=πGθδG,4(𝒜)=πG(θ1)δG,4(𝒜).T(\mathcal{A}*\mathcal{B})=\pi_{G*}\theta_{*}\delta_{G,4}^{*}(\mathcal{B}\boxtimes\mathcal{A})=\pi_{G*}(\theta^{-1})^{*}\delta_{G,4}^{*}(\mathcal{B}\boxtimes\mathcal{A}).

Since the composition θ1δG,4:(g,x,y)(g,gy,x,x,y)\theta^{-1}\delta_{G,4}:(g,x,y)\to(g,gy,x,x,y) is just ΔG,4σ\Delta_{G,4}\sigma, and σ\sigma is a GG-equivariant automorphism, so that πσ=π\pi\sigma=\pi, we are done. ∎

References

  • [1] A.-M. Aubert, D. Ciubotaru, B. Romano, A nonabelian Fourier transform for tempered unipotent representations, arXiv:2106.13969, (2021).
  • [2] V. Baranovsky, Orbifold Cohomology as Periodic Cyclic Homology, International Journal of Mathematics, 14 (2003), no. 8, 791-812.
  • [3] D. Ben-Zvi, D. Nadler, A. Preygel, A spectral incarnation of affine character sheaves, Compositio Math., 153 (2017), no. 9, 1908-1944.
  • [4] J. Bernstein, Draft of: Representations of pp-adic groups (written by K. Rummelhart), (1992).
  • [5] J. Bernstein, P. Deligne Le ‘‘centre’’ de Bernstein, ‘‘Representations des groups reductifs sur un corps local, Traveaux en cours’’ (P. Deligne ed.), Hermann, Paris, (1984), 1-32.
  • [6] R. Bezrukavnikov, On tensor categories attached to cells in affine Weyl groups, Representation theory of algebraic groups and quantum groups, Adv. Stud. Pure Math., Math. Soc. Japan, Tokyo, 40 (2004), 69-90.
  • [7] R. Bezrukavnikov, Noncommutative counterparts of the Springer resolution, Proceeding of the International Congress of Mathematicians, Madrid, Spain, 2 (2006), 1119-1144.
  • [8] R. Bezrukavnikov, Cohomology of tilting modules over quantum groups and t-structures on derived categories of coherent sheaves. Inventiones mathematicae, 166 (2006), no. 2, 327–357.
  • [9] R. Bezrukavnikov, Perverse sheaves on affine flags and nilpotent cone of the Langlands dual group, Israel Journal of Mathematics, 170 (2009), 185-206.
  • [10] R. Bezrukavnikov, On two geometric realizations of an affine Hecke algebra, Publ. IHES, 123 (2016), no. 1, 1–67.
  • [11] R. Bezrukavnikov, D. Ciubotaru, D. Kazhdan, Y. Varshavsky, in preparation.
  • [12] R. Bezrukavnikov, G. Dobrovolska, S. Dawydiak, On the structure of the affine asymptotic Hecke algebras (with an Appendix by Roman Bezrukavnikov, Alexander Braverman, and David Kazhdan), Transformation Groups, (2022), 1-21.
  • [13] R. Bezrukavnikov, I. Losev, Dimensions of modular irreducible representations of semisimple Lie algebras, Journal of American Math. Soc., (2023).
  • [14] R. Bezrukavnikov, I. Mirkovic, Representations of semisimple Lie algebras in prime characteristic and the noncommutative Springer resolution (with an Appendix by Eric Sommers), Ann. of math., 178 (2013), 835-919.
  • [15] R. Bezrukavnikov, V. Ostrik, On tensor categories attached to cells in affine Weyl groups II, Adv. Stud. Pure Math., (2004), 101-119.
  • [16] P. Baum, W. Fulton, R. MacPherson, Riemann-Roch for singular varieties, Publ. Math. IHES., 45 (1975), 101-145.
  • [17] B. Bhatt et al., Answer at Mathoverfow.
  • [18] A. Bialynicki-Birula, Some theorems on actions of algebraic groups, Ann. of Math., 98 (1973), no. 3, 480-497.
  • [19] A. Braverman, D. Kazhdan, Remarks on the asymptotic Hecke algebra, Lie Groups, Geometry, and Representation Theory: A Tribute to the Life and Work of Bertram Kostant, (2018), 91-108.
  • [20] N. Chriss, V. Ginzburg, Introduction to the geometric representation theory, Birkhäuser Boston, Inc., Boston, MA, (1997).
  • [21] D. Ciubotaru, X. He, Cocenters and representations of affine Hecke algebras, J. Eur. Math. Soc. (JEMS), 19 (2014), no. 10.
  • [22] S. Dawydiak, A coherent categorification of the based ring of the lowest two-sided cell, arXiv:2111.15648, (2021).
  • [23] S. Dawydiak, Three pictures of Lusztig’s asymptotic Hecke algebra, PhD thesis at University of Toronto, (2022).
  • [24] S. Dawydiak, The asymptotic Hecke algebra and rigidity, arXiv:2312.11092, (2023).
  • [25] V. Drinfeld, D. Gaitsgory, On a theorem of Braden, Transform. Groups 19, No. 2, 313-358 (2014).
  • [26] C. De Concini, G. Lusztig, C. Procesi, Homology of the zero set of a nilpotent vector field on a flag manifold, Journal of the American Mathematical Society, 1 (1988), no. 1, pp. 15-34.
  • [27] W. Fulton, Intersection Theory, Springer-Verlag, (1984).
  • [28] D. Edidin, W. Graham, Algebraic cycles and completions of equivariant KK-theory, Duke Math. J., 144 (2008), no. 3, 489-524.
  • [29] W. Graham. The forgetful map in rational KK-theory. Pacific Journal of Mathematics, 236 (2008), no. 1, 45-55.
  • [30] D. Halpern-Leinster, The derived category of GIT quotient, Journal of the American Mathematical Society, 28 (2015), no. 3, 871-912.
  • [31] A. Hattori, Rank element of a projective module, Nagoya Math. J., 25 (1965), 113-120.
  • [32] J. Ilmavirta et al., Answer at MathStackexchange.
  • [33] F. Knop, Answer at Mathoverflow.
  • [34] D. Kazhdan, Cuspidal geometry of pp-adic groups, Journal d’Analyse Mathématique, 47 (1986), 1-36
  • [35] D. Kazhdan, G. Lusztig, Representations of Coxeter groups and Hecke algebras, Invent. Math., 53 (1979), 165-184.
  • [36] D. Kazhdan, G. Lusztig, Proof of the Deligne-Langlands conjectures for affine Hecke algebras, Invent. Math., 87 (1987), 153-215.
  • [37] G. Lusztig, Cells in affine Weyl groups, Algebraic groups and related topics, 6 (1985), 255-287.
  • [38] G. Lusztig, Cells in affine Weyl groups, II, Journal of Algebra, 109 (1987), no. 2, 536-548.
  • [39] G. Lusztig, Cells in affine Weyl groups, III, J. Fac. Sci. Univ. Tokyo Sect. IA Math., 34 (1987), no. 2, 223-243.
  • [40] G. Lusztig, Cells in affine Weyl groups, IV, J. Fac. Sci. Univ. Tokyo Sect. IA Math., 36 (1989), no. 2, 297-328.
  • [41] G. Lusztig, Leading coefficients of character values of Hecke algebras, Proc. Symp. Pure Math. Amer. Math. Soc., 47 (1987), no. 2, 235-262.
  • [42] G. Lusztig, Cells in affine Weyl groups and tensor categories, Adv. Math., 129 (1997), 85-98.
  • [43] G. Lusztig, Bases in equivariant KK-theory, Representation Theory of the American Mathematical Society, 2 (1998), no. 9, 298-369.
  • [44] G. Lusztig, Bases in equivariant KK-theory-II, Representation Theory of the American Mathematical Society, 3 (1999), no. 11, 281-353.
  • [45] G. Lusztig, Unipotent elements in small characteristic, Transformation groups, 10 (2005), 449-487.
  • [46] P. Li, D. Nadler, Z. Yun, Functions on the commuting stack via Langlands duality, arXiv:2301.02618, (2023).
  • [47] A. Merkurjev, Equivariant KK-theory, Handbook of KK-theory, (2005), 925-954.
  • [48] S. Mohrdieck, Conjugacy classes of non-connected semisimple algebraic groups, Transformation Groups, 8 (2003), 377–395.
  • [49] S. Nie, The convolution algebra structure on KG(×)K^{G}(\mathcal{B}\times\mathcal{B}), arXiv:1111.1868, (2011).
  • [50] O. Propp, A Coherent Categorification of the Asymptotic Affine Hecke Algebra, PhD Thesis at MIT, (2023).
  • [51] Y. Qiu, N. Xi, The based rings of two-sided cells in an affine Weyl group of type B3~\tilde{B_{3}}, I, Science China Mathematics, 66 (2023), no. 2, 221-236.
  • [52] J. Serre, Groupes de Grothendieck des schémas en groupes réductifs déployés, Inst. Hautes ´ Etudes Sci. Publ. Math., (1968), no. 34, 37-52.
  • [53] J. Stallings, Centerless groups - an algebraic formulation of Gottlieb’s theorem, Topology, 4 (1965), no. 2, 129-134.
  • [54] H. Sumihiro, Equivariant completion, J. Math. Kyoto Univ., 14 (1974), no. 1, 1-28
  • [55] K. Suzuki, Trace Paley-Wiener theorem for Braverman-Kazhdan’s asymptotic Hecke algebra, arXiv:2407.02752, (2024).
  • [56] T. Tanisaki, N. Xi, Kazhdan-Lusztig basis and a geometric filtration of an affine Hecke algebra, Nagoya Math. J., 182 (2006), 285-311.
  • [57] P. Tauvel and R. W. T. Yu, Lie algebras and algebraic groups, Springer Monographs in Mathematics, (2005).
  • [58] B.  Toën, Proper local complete intersection morphisms preserve perfect complexes, arXiv:1210.2827, (2012).
  • [59] N. Xi, The based ring of the lowest two-sided cell of an affine Weyl group, J. Algebra, 134 (1990), 356-368.
  • [60] N. Xi, The Based Ring of Two-Sided Cells of Affine Weyl Groups of Type A~n1\tilde{A}_{n-1}, Amer. Math. Soc., (2002), no. 749.
  • [61] N. Xi, Representations of affine Hecke algebras and based rings of affine Weyl groups, J. Amer. Math. Soc., 20 (2007), no. 1, 211-217.
  • [62] N. Xi, Kazhdan-Lusztig Basis and A Geometric Filtration of an affine Hecke Algebra, II, J. Eur. Math. Soc., 13 (2011), 207-217.
  • [63] N. Xi, The based ring of the lowest two-sided cell of an affine Weyl group, III, Algebras and Representation Theory, 19 (2016), 1467-1475.