Hostname: page-component-6bf8c574d5-2jptb Total loading time: 0 Render date: 2025-02-19T22:11:45.549Z Has data issue: false hasContentIssue false

THE COHOMOLOGY OF p-ADIC DELIGNE-LUSZTIG SCHEMES OF COXETER TYPE

Published online by Cambridge University Press:  10 February 2025

Alexander B. Ivanov*
Affiliation:
Fakultät für Mathematik, Ruhr-Universität Bochum, D-44780 Bochum, Germany
Sian Nie
Affiliation:
Academy of Mathematics and Systems Science, Chinese Academy of Sciences, Beijing 100190, China School of Mathematical Sciences, University of Chinese Academy of Sciences, Chinese Academy of Sciences, Beijing 100049, China [email protected]
Rights & Permissions [Opens in a new window]

Abstract

We determine the cohomology of the closed Drinfeld stratum of p-adic Deligne–Lusztig schemes of Coxeter type attached to arbitrary inner forms of unramified groups over a local non-archimedean field. We prove that the corresponding torus weight spaces are supported in exactly one cohomological degree and are pairwise non-isomorphic irreducible representations of the pro-unipotent radical of the corresponding parahoric subgroup. We also prove that all Moy–Prasad quotients of this stratum are maximal varieties, and we investigate the relation between the resulting representations and Kirillov’s orbit method.

Type
Research Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2025. Published by Cambridge University Press

1 Introduction

Let k be a non-archimedean local field with residue characteristic $p>0$ and residue field $\mathbb {F}_q$ . Let $\breve {k}$ be the completion of the maximal unramified extension of k and let F denote the Frobenius automorphism of $\breve {k}$ over k. Let G be a reductive group over k, which splits over $\breve {k}$ . Let $T \subseteq B$ be a maximal torus and a Borel subgroup of G, such that T splits and B becomes rational over $\breve {k}$ . Let U resp. $\overline {U}$ denote the unipotent radical of B resp. of the opposed Borel subgroup. To $G,T,U$ one can attach an ind-(perfect scheme) over $\overline {\mathbb {F}}_q$ ,

(1.1) $$ \begin{align} X_{T,U} = \{g \in G(\breve{k}) \colon g^{-1}F(g) \in \overline{U} \cap FU \}, \end{align} $$

which is a variant of the p-adic Deligne–Lusztig spaces from [Reference Ivanov19]. Moreover, $X_{T,U}$ is endowed with an action of the locally compact group $G(k) \times T(k)$ , so that its $\ell $ -adic cohomology realizes smooth $G(k)$ -representations, parametrized by smooth characters of $T(k)$ , very much in the style of Deligne–Lusztig theory [Reference Deligne and Lusztig12]. Recently, the $\ell $ -adic cohomology of these and closely related spaces was extensively studied (especially when T is elliptic) and related with the local Langlands correspondences. See, for example, [Reference Chan and Ivanov9, Reference Chan and Oi10] for the relation with the type-theoretic construction of J.-K. Yu [Reference Yu31] and the related work of Kaletha and others (see, for example, [Reference Kaletha21]). However, see [Reference Chan and Ivanov9, §9], [Reference Feng14] for relations with Fargues–Scholze’s and Zhu’s geometric local Langlands [Reference Fargues and Scholze15, Reference Zhu34]. In this article, we continue the study of geometry and cohomology of $X_{T,U}$ .

Assume that $(T,U)$ is a Coxeter pair (see §2.5). In particular, T is elliptic and the apartment of T in the reduced affine building of G over k consists of one point. Bruhat–Tits theory attaches to this point a parahoric model $\mathcal {G}$ of G over the integers $\mathcal {O}_k \subseteq k$ with connected special fiber. Let $\mathcal {O}$ denote the integers of $\breve {k}$ . It was shown in [Reference Ivanov20, Reference Nie25] that $X_{T,U} \cong \coprod _{G(k)/\mathcal {G}(\mathcal {O}_k)} g X$ , where

(1.2) $$ \begin{align} X = \{ g \in \mathcal{G}(\mathcal{O}) \colon g^{-1}F(g) \in (\overline{\mathcal{U}} \cap F\mathcal{U})(\mathcal{O})\} \end{align} $$

is a perfect affine $\overline {\mathbb {F}}_q$ -scheme with $\mathcal {G}(\mathcal {O}_k) \times \mathcal {T}(\mathcal {O}_k)$ -action, and where we denote by $\mathcal {T},\mathcal {U} \subseteq \mathcal {G}$ the closures of $T,U$ . Cohomology of $X_{T,U}$ is then obtained by compactly inducing that of X.

There is a fibration $X \rightarrow X_{0+}$ over a Deligne–Lusztig variety $X_{0+}$ of the reductive quotient $\mathbb {G}_{0+} = (\mathcal {G} \otimes _{\mathcal {O}_k} \mathbb {F}_q)_{\mathrm {red}}$ of the special fiber of $\mathcal {G}$ . The variety $X_{0+}$ admits a natural stratification by locally closed subschemes. The stratification of X obtained by pulling it back was first considered in [Reference Chan and Ivanov8] (for $\mathrm {GL}_n$ and inner forms) resp. in [Reference Chan and Oi10, §6.2] (in general) and called the Drinfeld stratification there. There is a (in full generality only conjectural) relation between the cohomologies of X and of the strata; see [Reference Chan and Ivanov9, Theorem 5.1], [Reference Chan and Ivanov8, Conjecture 7.2.1], [Reference Chan and Oi10, Conjecture 6.5]

The cohomology of the unique closed stratum is very interesting and seems to be the most accessible one. When G is an inner form of $\mathrm {GL}_n$ , its cohomology as a $\mathcal {G}(\mathcal {O}_k) \times \mathcal {T}(\mathcal {O}_k)$ -representation was determined in [Reference Chan and Ivanov8, Theorem 6.1.1], the case of division algebras (where the closed stratum coincides with the whole scheme X) being already handled in [Reference Chan6]. The main goal of the present article is to extend these results to all G, thus giving a full account of the cohomology of the closed stratum. As a consequence, we also produce a rich supply of maximal varieties in the sense of [Reference Boyarchenko and Weinstein5] associated with groups other than $\mathrm {GL}_n$ . Our second goal is to investigate how this cohomology relates to representations obtained via Kirillov’s orbit method; see below.

To state our main result, let $\mathcal {G}^+$ be the pro-unipotent radical of $\mathcal {G}$ and let $\mathcal {T}^+,\ \mathcal {U}^+$ be the closures of $T,U$ in $\mathcal {G}^+$ . Then the closed stratum is a disjoint union of finitely many copies of the affine perfect scheme

(1.3) $$ \begin{align} Y = \{ g \in \mathcal{G}^{+}(\mathcal{O}) \colon g^{-1}F(g) \in (\overline{\mathcal{U}} \cap F\mathcal{U}^+)(\mathcal{O})\} \end{align} $$

with $\mathcal {G}^{+}(\mathcal {O}_k) \times \mathcal {T}^{+}(\mathcal {O}_k)$ -action. As Y is infinite-dimensional, it has no reasonable cohomology with compact support. We could remedy this by working with quotients of Y attached to Moy–Prasad quotients of $\mathcal {G}^{+}$ (and on the technical level, we will do precisely this). However, it seems most natural to state our results in terms of the homology functor $f_\natural $ , which is the left adjoint of $f^\ast $ , introduced in [Reference Ivanov and Mann18] in the schematic context following the approach of [Reference Fargues and Scholze15, VII.3] (see §2.7 for more details). Let therefore $H_i(Y,\overline {\mathbb {Q}}_\ell )$ denote the homology groups of the complex $f_\natural \overline {\mathbb {Q}}_\ell $ , where $f \colon Y \rightarrow \operatorname {\mathrm {Spec}} \overline {\mathbb {F}}_q$ is the structure map. If $\chi $ is a smooth character $\mathcal {T}^+(\mathcal {O}_k) \rightarrow \overline {\mathbb {Q}}_\ell ^\times $ , we also have the $\chi $ -weight part $f_\natural \overline {\mathbb {Q}}_\ell [\chi ]$ of $f_\natural \overline {\mathbb {Q}}_\ell $ . Let $N \geq 1$ be the smallest positive integer with $F^N U = U$ . Then Y has an obvious $\mathbb {F}_{q^N}$ -rational structure. A variety over a finite field is called maximal in [Reference Boyarchenko and Weinstein5] if its number of rational points attains the Weil–Deligne bound given by its Betti numbers.

Theorem 1.1. Suppose that $(T, U)$ is a Coxeter pair. For a smooth character $\chi \colon \mathcal {T}^{+}(\mathcal {O}_k) \rightarrow \overline {\mathbb {Q}}_\ell ^\times $ , the following hold.

  1. (1) Assume that p satisfies Condition 2.1.Footnote 1 The homology of $f_\natural \overline {\mathbb {Q}}_\ell [\chi ]$ is non-vanishing in precisely one degree $s_{\chi } \geq 0$ .

  2. (2) Assume that p satisfies Condition 2.1. The Frobenius $F^N$ acts in the space $H_{s_\chi }(Y,\overline {\mathbb {Q}}_\ell )[\chi ]$ as multiplication by the scalar $(-1)^{s_\chi } q^{s_\chi N/2}$ . In particular, all Moy–Prasad quotients of Y are $\mathbb {F}_{q^N}$ -maximal varieties.

  3. (3) For varying $\chi $ , $H_{s_\chi }(Y,\overline {\mathbb {Q}}_\ell )[\chi ]$ runs through pairwise non-isomorphic irreducible smooth $\mathcal {G}^{+}(\mathcal {O}_k)$ -representations.

This theorem follows from Theorems 5.5, 7.1 and Corollary 6.2 (where for part (1) the discussion of §2.7 and Corollary 5.10 apply). We determine the integer $s_\chi $ explicitly in terms of the Howe factorization of $\chi $ ; see Corollary 5.19.

Recently, Gordon obtained in his PhD thesis [Reference Gordon16] a result closely related to Theorem 1.1(2).

The same proof of Theorem 7.1, combined with Remark 3.2, shows that the statement (3) of Theorem 1.1 is true if $(T, U)$ is a minimal elliptic pair; see §2.5. This partially motivates us to propose the following conjecture.

Conjecture 1.2. Theorem 1.1 holds for all minimal elliptic pairs $(T, U)$ .

Using parts (1),(2) of the theorem along with a fixed point formula of Boyarchenko [Reference Boyarchenko3, Lemma 2.12], we give the following representation-theoretic interpretation of the integer $s_\chi $ , generalizing [Reference Chan and Ivanov9, Lemma 8.1].

Corollary 1.3. If Condition 2.1 holds for p, then $\dim _{\overline {\mathbb {Q}}_\ell } H_{s_\chi }(Y,\overline {\mathbb {Q}}_\ell )[\chi ] = q^{s_\chi N/2}$ .

This corollary is proven in §6. More generally, we obtain a trace formula for any element of $\mathcal {G}^+(\mathcal {O}_k)$ on $H_{s_\chi }(Y,\overline {\mathbb {Q}}_\ell )[\chi ]$ in terms of geometric points of (a Moy–Prasad quotient of) Y; see Proposition 6.1.

To apply our main result to the cohomology of $X_{T,U}$ (in the style of [Reference Chan and Ivanov9]), it is necessary to study the relation between the cohomology of X and of the closed stratum ([Reference Chan and Oi10, Conjecture 6.5]); this will be considered in a follow-up work. Once this is done, our results, combined with the main results of [Reference Chan and Oi10] and [Reference Dudas and Ivanov13] (see [Reference Dudas and Ivanov13, Corollary 1.0.2]), would give geometric approaches to some representation-theoretic questions. For example, Corollary 1.3 allows a purely geometric proof of the formal degree formulas for many supercuspidal representations (note that an algebraic computation is given in the recent work of Schwein [Reference Schwein26]).

The second goal of this article is to formulate and verify in a special case a conjecture about the relation of the homology of Y with Kirillov’s orbit method for the pro-p-group $\mathcal {G}^+(\mathcal {O}_k)$ , whenever the latter applies. Namely, by a theory of Lazard, a uniform pro-p-group (resp. a p-group of nilpotence class $<p$ ) $\Gamma $ is completely described by its $\mathbb {Z}_p$ -Lie algebra (resp. finite Lie ring) $\mathfrak {g}$ via an exponential map; see [Reference Boyarchenko and Drinfeld1, §2]. Kirillov’s orbit method establishes a natural bijection between smooth irreducible representations of $\Gamma $ and adjoint $\Gamma $ -orbits in the dual $\mathfrak {g}^\ast = \mathrm {Hom}_{\mathrm {cont}}(\mathfrak {g},\overline {\mathbb {Q}}_\ell ^\times )$ – see [Reference Boyarchenko and Sabitova4] – characterized by a trace formula. Often it happens that $\mathcal {G}^+(\mathcal {O}_k)$ (resp. its Moy–Prasad quotient) is a uniform pro-p-group (resp. p-group of nilpotence class $<p$ ). In this case, the natural question to determine the adjoint orbit corresponding to $H_{s_\chi }(Y,\overline {\mathbb {Q}}_\ell )[\chi ]$ arises. In Conjecture 8.4, we make this precise. We verify this conjecture for the finite p-group $\{ g \in \mathrm {GL}_2(\mathbb {F}_q[\varpi ]/\varpi ^3) \colon g \equiv 1\ \mod \varpi \}$ if q is odd.

Finally, we complete the task of comparing the spaces $X_{T,U}$ from (1.1) with the p-adic Deligne–Lusztig spaces from [Reference Ivanov19], when $(T,U)$ is a Coxeter pair. This was done for classical groups in [Reference Ivanov20, Proposition 5.12], and in §4.1, we prove it for general G. To achieve this, we need to extend the loop version of twisted Steinberg’s cross-section (see [Reference He and Lusztig17, 3.6], [Reference Ivanov20, Proposition 5.3] and [Reference Malten23]) to non-classical groups; see Proposition 3.1. Note that this result is also used in the proof of Theorem 1.1(3).

2 Notation and setup

2.1 General notation

Throughout the article, we let $\breve {k}/k$ with integers $\mathcal {O}_k \subseteq \mathcal {O}$ , residue field extension $\overline {\mathbb {F}}_q/\mathbb {F}_q$ , and Frobenius F be as in the introduction. We denote by $\varpi $ a uniformizer of k.

Given a $\mathbb {F}_q$ -algebra R, let $\mathrm {Perf}_R$ be the category of perfect R-algebras. For $R \in \mathrm {Perf}_{\mathbb {F}_q}$ , let $W(R)$ be the ring of p-typical Witt vectors of R, and put $\mathbb {W}(R) = W(R) \otimes _{\mathbb {Z}_p} \mathcal {O}_k$ if $\mathrm {char}(k) = 0$ , resp. $\mathbb {W}(R) = R[\![\varpi ]\!]$ otherwise. In particular, $\mathbb {W}(\mathbb {F}_q) = \mathcal {O}_k$ and $\mathbb {W}(\overline {\mathbb {F}}_q) = \mathcal {O}$ . Let $[\cdot ] \colon R \rightarrow \mathbb {W}(R)$ be the Teichmüller lift if $\mathrm {char}(k) = 0$ , resp. $[x] = x$ if $\mathrm {char}(k)> 0$ .

Let $\mathcal {X}$ be any $\mathcal {O}$ -scheme and let X be any $\breve {k}$ -scheme. We will abbreviate

$$\begin{align*}\breve{\mathcal{X}} := \mathcal{X}(\mathcal{O}) \quad \text{ and } \quad \breve{X} = X(\breve{k}). \end{align*}$$

Suppose that $\mathcal {X}$ is affine and of finite type over $\mathcal {O}$ . We regard the set $\breve {\mathcal {X}}$ as a perfect affine scheme $\mathbb {X}$ over $\overline {\mathbb {F}}_q$ , so that $\mathbb {X}(\overline {\mathbb {F}}_q) = \breve {\mathcal {X}}$ . More precisely, one puts $\mathbb {X} = L^+\mathcal {X}$ , where $L^+\mathcal {X} \colon \mathrm {Perf}_{\overline {\mathbb {F}}_q} \rightarrow \mathrm {Sets}$ , $L^+\mathcal {X}(R) = \mathcal {X}(\mathbb {W}(R))$ is the functor of positive loops; see, for example, [Reference Chan and Ivanov7, §2.5] for details. We always will identify the scheme $\mathbb {X}$ with the set $\breve {\mathcal {X}}$ of its geometric points. If $\mathcal {X}$ is defined over $\mathcal {O}_k$ , $\mathbb {X}$ has a natural $\mathbb {F}_q$ -structure, corresponding to the F-action on $\breve {\mathcal {X}}$ . Moreover, the set

$$\begin{align*}\mathbb{X}(\mathbb{F}_q) = \breve{\mathcal{X}}^F = \mathcal{X}(\mathcal{O}_k) \end{align*}$$

has a natural structure of a profinite set. Similarly, if X is affine of finite type over $\breve {k}$ , then we regard $\breve {X}$ as an ind-(perfect affine scheme) over $\overline {\mathbb {F}}_q$ via the loop functor $LX \colon \mathrm {Perf}_{\overline {\mathbb {F}}_q} \ni R \mapsto X(\mathbb {W}(R)[p^{-1}])$ , and the same claim about $\mathbb {F}_q$ -structure holds, except that now $\breve {X}^F$ is only locally profinite.

2.2 Group-theoretic setup

We fix a reductive group G defined over k and split over $\breve {k}$ . We fix a k-rational, $\breve {k}$ -split maximal torus T of G, and we denote by $N_G(T)$ its normalizer. Its Weyl group $W= N_G(T)/T$ is a finite étale group scheme over k becoming constant over $\breve {k}$ . We identify W with the set of its $\breve {k}$ -points, endowed with the action of F. We denote by $X_\ast (T)$ , $X^\ast (T)$ the groups of (co)characters of $T_{\breve {k}}$ , equipped with natural F-actions, and by $\langle ,\rangle \colon X^\ast (T) \times X_\ast (T) \rightarrow \mathbb {Z}$ the natural W- and F-equivariant pairing. We denote by N the order of F as an automorphism of $X_\ast (T)$ .

We fix a Borel subgroup $T \subseteq B \subseteq G$ defined over $\breve {k}$ , and we denote by U the unipotent radical of B. Denote by $\Phi \subseteq X^\ast (T)$ the set of roots of T in G, and by $\Phi ^+$ resp. $\Phi ^-$ the subset of positive roots corresponding to U resp. $\overline {U}$ . For each $\alpha \in \Phi $ , let $U_\alpha \cong \mathbb {G}_{a,\breve {k}}$ denote the corresponding root subgroup.

2.3 Filtration of the torus and affine roots

Let $\mathcal {T}$ denote the connected Néron model of T. Let $\breve T^0$ be the maximal bounded subgroup of $\breve T$ . Then $\mathcal {T}(\mathcal {O}) = \breve T^0$ . Moreover, for $r \in \mathbb {Z}_{\geq 0}$ ,

$$\begin{align*}\breve T^r = \{t \in \breve T^0 \colon \operatorname{\mathrm{ord}}_{\varpi}(\chi(t) - 1) \geq r \, \forall \chi \in X^\ast(T) \} \end{align*}$$

defines a descending separated filtration on $\breve T$ . For each r, one has an isomorphism

$$\begin{align*}V := X_\ast(T) \otimes \overline{\mathbb{F}}_q \stackrel{\sim}{\longrightarrow} \breve T^r /\breve T^{r+1}, \quad \lambda \otimes x \mapsto \lambda(1 + [x]\varpi^r). \end{align*}$$

Fix some (e.g., hyperspecial) point $\textbf {x}_0$ in the apartment $\mathcal {A}_{T,\breve {k}}$ of T in the reduced affine building of G over $\breve {k}$ . Let

$$\begin{align*}\widetilde \Phi_{\mathrm{aff}} = \{\alpha+m: x \mapsto -\alpha(x - {\mathbf x}_0) + m; \alpha \in \Phi, m \in \mathbb{Z} \} \cong \Phi \times \mathbb{Z} \end{align*}$$

be the set of affine roots. Let $\widetilde \Phi = \Phi _{\mathrm {aff}} \sqcup \mathbb {Z}_{\geq 0}$ be the (enlarged) set of affine roots of T in G. For an affine root $\alpha +m$ , we have the corresponding subgroup $\breve U_{\alpha +m} \subseteq \breve U$ . For $m \in \mathbb {Z}_{\geq 0}$ , the corresponding root subgroup is $\breve T^m$ . There is an action of F on $\widetilde \Phi $ , such that $F\breve U_{\alpha +m} = \breve U_{F(\alpha +m)}$ .

2.4 Parahoric model and Moy–Prasad quotients

Assume that T is elliptic. Then the apartment of T in the reduced affine building of G over k consists of precisely one point ${\mathbf x}$ . We denote by $\mathcal {G}$ the parahoric $\mathcal {O}_k$ -model of G with connected special fiber attached to ${\mathbf x}$ , and by $\mathcal {G}^+$ its pro-unipotent radical.

If $H \subseteq G$ is a closed subgroup, then we denote by $\mathcal {H}$ the closure of H in $\mathcal {G}$ .Footnote 2 Similarly, we denote by $\mathcal {T}^+$ the closure of T in $\mathcal {G}^+$ .

Note that $\breve {\mathcal {G}}$ (resp. $\breve {\mathcal {G}}^+$ ) is generated by all $\breve U_{f}$ with $f \in \widetilde \Phi $ satisfying $f({\mathbf x}) \geq 0$ (resp. $f({\mathbf x})> 0$ ), and that $\breve {\mathcal {G}}/\breve {\mathcal {G}}^+$ is naturally isomorphic to the reductive quotient of the special fiber of $\mathcal {G}$ .

For any $h = r$ or $h = r+$ with $r \in \mathbb {Z}_{\geq 0}$ , Moy–Prasad have defined in [Reference Moy and Prasad24] the normal F-stable subgroup $\breve {\mathcal {G}}^h \subseteq \breve {\mathcal {G}}$ generated by all $\breve U_f$ with $f \in \widetilde \Phi $ satisfying $f({\mathbf x}) \geq h$ . Note that $\breve {\mathcal {G}} = \mathcal {G}(\mathcal {O})^0$ and $\breve {\mathcal {G}}^+ = \mathcal {G}(\mathcal {O})^{0+}$ . There is a smooth $\mathbb {F}_q$ -group scheme $\mathbb {G}_r$ with

$$\begin{align*}\mathbb{G}_r(\overline{\mathbb{F}}_q) = \breve{\mathcal{G}}/\breve{\mathcal{G}}^r. \end{align*}$$

It has the subgroup $\mathbb {G}_r^+ = \breve {\mathcal {G}}^+/\breve {\mathcal {G}}^r$ , and the set of affine roots appearing in $\mathbb {G}_r^+$ is

$$\begin{align*}\widetilde \Phi_r^+ = \{f \in \widetilde\Phi \colon 0 < f({\mathbf x}) < r \}. \end{align*}$$

According with §2.1, we have also the $\mathbb {F}_q$ -groups $\mathbb {G}$ and $\mathbb {G}^+$ such that $\mathbb {G}(\overline {\mathbb {F}}_q) = \breve {\mathcal {G}}$ and $\mathbb {G}^+(\overline {\mathbb {F}}_q) = \breve {\mathcal {G}}^+$ . Note that and .

Note that any of the subgroups $H = T,B,U, \dots $ of G defines a closed subgroup $\mathbb {H}_r \subseteq \mathbb {G}_r$ (resp. $\mathbb {H} \subseteq \mathbb {G}$ ) by first taking the closure $\mathcal {H} \subseteq \mathcal {G}$ of H, and then letting $\mathbb {H}_r(\overline {\mathbb {F}}_q)$ be the image of the map $\mathcal {H}(\mathcal {O}) \rightarrow \mathcal {G}(\mathcal {O}) \rightarrow \mathbb {G}_r(\overline {\mathbb {F}}_q)$ . Similarly, H defines a closed subgroup $\mathbb {H}_r^+ \subseteq \mathbb {G}_r^+$ (and $\mathbb {H}^+ \subseteq \mathbb {G}^+$ ). Note that if $F^sH = H$ for some $s\geq 1$ , then $\mathbb {H}_r,\mathbb {H}_r^+$ are defined over $\mathbb {F}_{q^s}$ .

2.5 Coxeter pairs

Let $c \in W$ be the unique element such that $FB = {}^c B$ . Then for any lift $\dot c$ of c, ${\mathrm {Ad}}({\dot c})^{-1} \circ F: \breve G \to \breve G$ fixes the pinning $(T,B)$ , and hence defines an automorphism $\sigma _W$ of the Coxeter system $(W,S)$ . We call $(T,B)$ (and $(T,U)$ ) a Coxeter pair if c is a Coxeter element in the Coxeter triple $(W,S,\sigma _W)$ – that is, if a(ny) reduced expression of c contains precisely one element from each $\sigma _W$ -orbit on S. More generally, $(T, U)$ is called a minimal elliptic pair if c is of minimal length in its $\sigma _W$ -twisted conjugacy class. We have implications $(T,B)$ Coxeter $\Rightarrow $ $(T,B)$ minimal elliptic $\Rightarrow $ T is elliptic.

We define

$$\begin{align*}\Delta := \Phi^- \cap F\Phi^+. \end{align*}$$

Note that if $(T,B)$ is Coxeter, then each F-orbit in $\Phi $ has length exactly N and intersects the set in precisely one element; see, for example, [Reference Steinberg28, §7]. In particular, $\#\Delta $ is equal to the semisimple rank of G, $\Phi / \langle c \sigma _W \rangle \cong \Delta $ and $\#\Phi = N \cdot \#\Delta $ .

2.6 A condition on p

Assume that T is elliptic. We will prove Theorem 5.5 under the following condition on the characteristic p of $\mathbb {F}_q$ , which is satisfied if p does not divide the order of the Weyl group of G.

Condition 2.1. The characteristic p of $\mathbb {F}_q$ is not a torsion prime for $\Phi $ (see [Reference Steinberg29, Definition 1.3]), and p does not divide $\#\pi _1(M_{\mathrm {der}})$ for any F-stable Levi subgroup M containing T. Here, $M_{\mathrm {der}}$ denotes the derived subgroup of M.

Note the all torsion primes for $\Phi $ are $\leq 5$ . Note that the second part of this condition holds for all p when $G_{\mathrm {der}}$ is simply connected. Let $P = P(G,T)$ denote the set of primes, for which this condition does not hold. If G is of type $A_n$ , then $P \subseteq \{\ell \text { prime } \colon \ell \text { divides } n \}$ . If G is of type $B_n$ or $C_n$ with n even, then $P \subseteq \{2\}$ . If G is of type $B_n$ or $C_n$ with n odd, then $P \subseteq \{2\} \cup \{\ell \text { prime } \colon \ell \text { divides } n\}$ . If G is of type $D_n$ , then $P \subseteq \{\ell \text { prime } \colon \ell < n\}$ .

We will use this condition in the proof of Theorem 5.5 by applying the following lemma to derived subgroups of various F-stable unramified twisted Levi subgroups of G containing T. Recall $V = X_\ast (T) \otimes \overline {\mathbb {F}}_q$ from §2.3 and consider the following norm map:

$$\begin{align*}\mathrm{Nm}_N: V \to V, ~ v \mapsto v + F(v) + \cdots + F^{N-1}(v). \end{align*}$$

Lemma 2.2. Suppose that G is semisimple and p does not divide $\#\pi _1(G)$ . Then $V^F = \mathrm {Nm}_N(\mathbb {Z}\Phi ^\vee \otimes \mathbb {F}_{q^N})$ , where $\Phi ^\vee $ is the set of coroots.

Proof. By assumption, we have $\mathbb {Z}\Phi ^\vee \otimes \mathbb {F}_{q^N} = X_*(T) \otimes \mathbb {F}_{q^N}$ . Hence,

$$\begin{align*}V^F = \mathrm{Nm}_N(V^{F^N}) = \mathrm{ Nm}_N(X_*(T) \otimes \mathbb{F}_{q^N}) = \mathrm{Nm}_N(\mathbb{Z}\Phi^\vee \otimes \mathbb{F}_{q^N}),\end{align*}$$

as desired.

2.7 Homology

For a morphism $Y \rightarrow Z$ of perfect $\mathbb {F}_p$ -schemes and a coefficient ring $\Lambda $ , which we assume to be either $\overline {\mathbb {Q}}_\ell $ or $\overline {\mathbb {F}}_\ell $ here, in [Reference Ivanov and Mann18] the left adjoint $f_\natural \colon D_{\blacksquare }(Y,\Lambda ) \rightarrow D_{\blacksquare }(Z, \Lambda )$ of $f^\ast $ on unramified solid sheaves is constructed. Readers feeling uncomfortable with the use $f_\natural $ , could just regard (2.2) as a definition (which is well-behaved because of (2.1)). Assume that $Z = \operatorname {\mathrm {Spec}} \overline {\mathbb {F}}_q$ , in which case we get the $\Lambda $ -module

$$\begin{align*}H_i(Y,\Lambda) := H^{-i}f_\natural\Lambda. \end{align*}$$

Assume now that with all $f_r \colon Y_r \rightarrow \operatorname {\mathrm {Spec}} \overline {\mathbb {F}}_q$ perfections of smooth morphisms of dimension $d_r$ . Assume that there are compatible actions of finite groups $\Gamma _r$ on $Y_r$ , inducing an action of on Y. Let $\chi \colon \Gamma \rightarrow \Lambda ^\times $ be a smooth character. There is some $r_\chi \geq 0$ such that for each $r\geq r_\chi $ , $\chi $ factors through a character of $\Gamma _r$ again denoted $\chi $ . Assume that for all $r \geq r_\chi $ , the map

(2.1) $$ \begin{align} f_{r !}\Lambda[\chi][2(d_r - d_{r_\chi})] \rightarrow f_{r_\chi !}\Lambda[\chi]. \end{align} $$

is an isomorphism. It then follows from the properties of $f_\natural $ – see [Reference Ivanov and Mann18] – that

where the second equality holds because $f_r$ is smooth (and hence $f_{r \natural } = f_{r!}[2d_r]$ ) and the last equality is by (2.1). With other words,

(2.2) $$ \begin{align} H_i(Y,\Lambda)[\chi] = H^{2d_r - i}_c(Y_{r_\chi},\Lambda)[\chi] \quad \text{for all } r \geq r_\chi. \end{align} $$

3 Steinberg’s cross-section

The following proposition is a variant of [Reference He and Lusztig17, 3.6, 3.14], generalizing [Reference Ivanov20, Proposition 5.3].

Proposition 3.1. Suppose $(T,U)$ is a Coxeter pair. Then the map $(x, y) \mapsto x^{-1} y F(x)$ induces isomorphisms:

(1) $(\mathbb {U}_r \cap F\mathbb {U}_r) \times (\overline {\mathbb {U}}_r \cap F\mathbb {U}_r) \cong F\mathbb {U}_r$ ;

(2) $\mathbb {U}_r \times (\overline {\mathbb {U}}_r \cap F\mathbb {U}_r) \cong \mathbb {U}_r F\mathbb {U}_r = \mathbb {U}_r (\overline {\mathbb {U}}_r \cap F\mathbb {U}_r) \cong \mathbb {U}_r \times (\overline {\mathbb {U}}_r \cap F\mathbb {U}_r)$ .

Moreover, the analogous statements also hold with $\mathbb {U}_r$ replaced by $\mathbb {U}_r^+$ or by $\breve {\mathcal {U}}$ or by $\breve {\mathcal {U}}^+$ or by $\breve U$ .

Remark 3.2. Using a different approach, Malten [Reference Malten23] shows that Proposition 3.1 holds for all minimal elliptic pairs $(T, U)$ . We will not use this result in the paper.

We use this result for $\mathbb {U}_r^+$ in §7 to deduce the irreducibility of $H_{s_\chi }(Y,\overline {\mathbb {Q}}_\ell )[\chi ]$ , and for $\breve U$ in §4 to prove the isomorphism of $X_{T,U}$ with the p-adic Deligne–Lusztig space from [Reference Ivanov19].

Proof. In any of the setups ( $\mathbb {U}_r,\mathbb {U}_r^+,\breve {\mathcal {U}},\breve {\mathcal {U}}^+,\breve U$ ), (1) is equivalent to (2) as in [Reference He and Lusztig17, 3.14], so it suffices to prove (1). By [Reference He and Lusztig17, 3.6], the map in (1) is always injective.

In the setup with $\mathbb {U}_r$ resp. $\mathbb {U}_r^+$ the proposition follows from injectivity and [Reference He and Lusztig17, Proposition 1.2(ii)], as the source and the target of the map in (1) are isomorphic to the (perfect) affine space over $\overline {\mathbb {F}}_q$ of the same finite dimension. By passing to the inverse limit over r, the proposition also follows in the setup with $\breve {\mathcal {U}},\breve {\mathcal {U}}^+$ .

It remains to handle the setup with $\breve U$ , where we argue as in [Reference Ivanov20, Proposition 5.3]. By [Reference He and Lusztig17, §3.5], it suffices to prove (1) for a single Coxeter element. By [Reference Ivanov20, Lemma 5.5], it suffices to assume that the Dynkin diagram of G is connected. The cases when G is classical were handled in [Reference Ivanov20, Proposition 5.3], so it suffices to verify [Reference Ivanov20, Lemma 5.7] for the remaining types ( $G_2$ , $F_4$ , $E_6$ , $E_7$ , $E_8$ , ${}^3D_4$ , ${}^2E_6$ ). That is, we must provide a filtration

$$\begin{align*}\Phi^+ = \Psi_r \supseteq \Psi_{r-1} \supseteq \dots \Psi_2 \supseteq \Psi_1 = \Phi^+ \cap F^{-1}(\Phi^-),\end{align*}$$

such that for each i, $\Psi _i$ and $\Psi _i {\,\setminus \,} \Psi _1$ are closed under addition; the implication $\alpha ,\beta \in \Psi _i$ , $\alpha +\beta \in \Phi \Rightarrow \alpha +\beta \in \Psi _{i-1}$ holds for all $i>1$ ; and for all i, $F(\Psi _i {\,\setminus \,} \Psi _1) \subseteq \Psi _i$ . We do this using an algorithm implemented in SAGE [30]. It even turns out that it is always possible to arrange that $\#(\Psi _{i+1} {\,\setminus \,} \Psi _i) = 1$ . Our algorithm is explained in Appendix A.

4 Review of some properties of $X_{T,U}$

Let $X_{T,U}$ be as in (1.1). Here, we recall/prove some facts about it. As we explain below, if T is elliptic, there is an equivariant map $X_{T,U} \rightarrow \dot X_{\dot w}(b)$ into a certain p-adic Deligne–Lusztig space from [Reference Ivanov19, §8]. If $(T,U)$ is Coxeter and if G is classical in the sense of [Reference Ivanov20, Definition 5.1], it was shown in [Reference Ivanov20, Proposition 5.12] that this map is an isomorphism. We prove in this section that this holds for all G.

4.1 Comparison with the definition in [Reference Ivanov19]

Assume that T is elliptic. Assume that G admits a (necessarily unique) unramified inner form $G_0$ over k (the general case easily reduces to this by using a derived embedding of G into a group with connected center). Then one can choose

  • a k-rational pinning $(T_0, B_0 = T_0 U_0)$ of $G_0$ with Weyl group $(W_0,S_0)$ ,

  • an elliptic element $w \in W_0$ ,

  • a lift $\dot w \in N(T_0)(\breve {k})$ ,

such that there is a $\breve {k}$ -rational isomorphism $G \stackrel {\sim }{ \rightarrow } G_0$ , identifying $T,B,U,W$ with $T_0,B_0,U_0,W_0$ , and F with $Ad(\dot w) \circ \sigma $ as an automorphism of $\breve G \cong \breve G_0$ . Let $b \in \breve G$ . In [Reference Ivanov19, §8], the p-adic Deligne–Lusztig space attached to the datum $(G_0, \dot w, b)$ is defined as the arc-sheaf on perfect $\overline {\mathbb {F}}_q$ -schemes,

$$\begin{align*}\dot X_{\dot w}(b) = \{x \in L(G_0/U_0) \colon x^{-1}b\sigma(x) \in L(U_0 \dot w U_0) \}, \end{align*}$$

where $L(\cdot )$ is the perfect loop functor as in §2.1. Note that $(g,t) \colon x \mapsto gxt$ defines an action of the locally profinite group $G(k) \times T(k)$ on this arc-sheaf; see [Reference Ivanov19, §8] for details. This seems to be a natural definition, most similar to classical Deligne–Lusztig varieties.

Note that w equals the relative position of U with $FU$ . Thus, $(T,U)$ is a Coxeter (resp. minimal elliptic) pair if and only if w is a Coxeter (resp. minimal elliptic) element.

Identifying G with $G_0$ via the given isomorphism, we have the composition

(4.1) $$ \begin{align} \nonumber X_{T,U} & \rightarrow \{g \in \breve G_0 \colon g^{-1}\dot w\sigma(g) \in \dot w \breve U_0 \}/(\breve U_0 \cap {}^{w}\breve U_0)\\ &\stackrel{\sim}{ \rightarrow } \dot X_{\dot w}(\dot w), \end{align} $$

given by $g \mapsto g (\breve U_0 \cap {}^{w}\breve U_0) \mapsto g\breve U_0$ . Just as was done in [Reference Ivanov19, Proposition 5.12] for classical groups, we deduce from Proposition 3.1:

Corollary 4.1. Assume $(T,U)$ is a Coxeter pair. Then the map (4.1) is a $G(k)\times T(k)$ -equivariant isomorphism.

4.2 Integral decomposition of $X_{T,U}$

A priori, $X_{T,U}$ is a huge ind-scheme, which is hard to control. However, in the Coxeter case, it has the following decomposition. Let X be as in (1.2) and note that $X \subseteq X_{T,U}$ is a closed subscheme. Surprisingly, it is also open and the following holds.

Theorem 4.2 [Reference Ivanov20],[Reference Nie25].

Suppose $(T,U)$ is a Coxeter pair and let X be as in (1.2). Then there is a decomposition

$$\begin{align*}X_{T,U} = \bigsqcup_{g \in G(k)/\mathcal{G}(\mathcal{O}_k)} g X. \end{align*}$$

In particular, $X_{T,U}$ is a disjoint union of affine perfect $\overline {\mathbb {F}}_q$ -schemes.

This reduces the study of the cohomology of $X_{T,U}$ to that of X.

4.3 Drinfeld stratification

In this subsection, the ellipticity assumption on T can be dropped. Note that the projection $\mathbb {G} \rightarrow \mathbb {G}_{0+}$ restricts to a projection

$$\begin{align*}X \rightarrow X_{0+} = \{g \in \mathbb{G}_{0+} \colon g^{-1}F(g) \in \overline{\mathbb{U}}_{0+} \cap F\mathbb{U}_{0+}\}\end{align*}$$

over (a variant of) a classical Deligne–Lusztig variety. Let $\mathfrak {L}$ denote the set of all twisted Levi subgroups of $\mathbb {G}_{0+}$ containing $\mathbb {T}_{0+}$ . For any $\mathbb {L}_{0+} \in \mathfrak {L}$ , we have the locally closed $\mathbb {G}_{0+}^F \times \mathbb {T}_{0+}^F$ -stable closed subscheme

$$\begin{align*}X_{0+}^{(\mathbb{L}_{0+})} = \{ g \in \mathbb{G}_{0+} \colon g^{-1}F(g) \in \mathbb{L}_{0+} \cap \overline{\mathbb{U}}_{0+} \cap F\mathbb{U}_{0+} \}. \end{align*}$$

Pulling back to X, we obtain a closed subscheme $X^{(\mathbb {L}_{0+})} \subseteq X$ . Following [Reference Chan and Ivanov8] and [Reference Chan and Oi10, §6.2], we then call

$$\begin{align*}X^{(\mathbb{L}_{0+})} {\,\setminus\,} \bigcup_{\mathbb{L}_{0+}' \subseteq \mathbb{L}_{0+} \in \mathfrak{L}} X^{(\mathbb{L}'_{0+})} \end{align*}$$

a Drinfeld stratum of X. This defines a finite and locally closed stratification of X. Its has a unique minimal/closed stratum $X^{(\mathbb {T}_{0+})}$ .

Lemma 4.3. With Y as in (1.3), we have $X^{(\mathbb {T}_{0+})} = \bigsqcup _{g \in \mathbb {G}_{0+}^F/\mathbb {T}_{0+}^F} g (X \cap \mathbb {T} \mathbb {G}^+) = \bigsqcup _{g \in \mathbb {G}_{0+}^F} g Y$ .

Proof. The first equality is [Reference Chan and Ivanov8, Lemma 3.3.3]. As $\mathbb {T}_{0+} \cap \overline {\mathbb {U}}_{0+} \cap F\mathbb {U}_{0+} = 1$ , the image of $X^{(\mathbb {T}_{0+})}$ under $X \rightarrow X_{0+}$ is contained in the finite subset $\mathbb {G}_{0+}^F \subseteq X_{0+}$ . By exploiting the $\mathbb {G}^F$ -action on X and the surjectivity of $\mathbb {G}^F \rightarrow \mathbb {G}_{0+}^F$ , each fiber is a translate of Y.

In the rest of the article, we consider Y and its cohomology. To approximate Y, consider for any $r \in \mathbb {Z}_{>0}$ the affine perfect $\overline {\mathbb {F}}_q$ -scheme

$$\begin{align*}Y_r = \{ g \in \mathbb{G}_r^+ \colon g^{-1}F(g) \in \overline{\mathbb{U}}_r\cap F\mathbb{U}_r^+ \}, \end{align*}$$

equipped with $(\mathbb {G}_r^+)^F \times (\mathbb {T}_r^+)^F$ -action, so that . Similarly, we have the schemes $X_r^{(\mathbb {L}_{0+})} \subseteq \mathbb {G}_r$ approximating $X^{(\mathbb {L}_{0+})}$ .

Recall the set $\Delta $ from §2.5. Let $\Phi ^{\mathrm {red}}$ denote the set of those $\alpha \in \Phi $ for which $\alpha ({\mathbf x}) \in \mathbb {Z}$ , and let $\Delta ^{\mathrm {red}} = \Phi ^{\mathrm {red}} \cap \Delta $ .

Lemma 4.4. The scheme $Y_r$ is the perfection of an affine smooth scheme of dimension $r \cdot \#\Delta - \#\Delta ^{\mathrm {red}} = \frac {1}{N} (r \cdot \#\Phi - \Phi ^{\mathrm {red}})$ .

Proof. The last equality follows from the last sentense of §2.5. Note that if $\alpha \in \Delta ^{\mathrm {red}}$ (resp. $\alpha \in \Delta {\,\setminus \,} \Delta ^{\mathrm {red}}$ ), then there are precisely $r-1$ (resp. precisely r) affine roots with vector part $\alpha $ appearing in $\overline {\mathbb {U}}_r \cap F\mathbb {U}_r^+$ . Thus, $\overline {\mathbb {U}}_r \cap F\mathbb {U}_r^+$ is isomorphic to the perfection of $\mathbb {A}^{r\cdot \#\Delta - \#\Delta ^{\mathrm {red}}}_{\overline {\mathbb {F}}_q}$ . Note that $Y_r$ is the pullback of $\overline {\mathbb {U}}_r \cap F\mathbb {U}_r^+$ under the Lang map $g\mapsto g^{-1}F(g)$ of $\mathbb {G}_r^+$ . By [Reference Zhu33, Lemma A.26], there is a smooth algebraic group $\mathbb {H}$ over $\mathbb {F}_q$ with perfection $\mathbb {G}_r^+$ . Let $\mathbb {W} \subseteq \mathbb {H}$ be the (reduced) closed subgroup whose perfection is $\overline {\mathbb {U}}_r \cap F\mathbb {U}_r^+$ . In particular, $\mathbb {W}$ is necessarily isomorphic to $\mathbb {A}^{r\cdot \#\Delta - \#\Delta ^{\mathrm {red}}}_{\overline {\mathbb {F}}_q}$ . Let $Y_r'$ be the pullback of $\mathbb {W}$ under the Lang map of $\mathbb {H}$ . As perfection commutes with limits, $Y_r$ is the perfection of $Y_r'$ . As the Lang map is étale, the claim follows.

5 The minimal Drinfeld stratum

In this section, we assume that $\Phi / \langle c \sigma _W\rangle \cong \Delta $ and $\#\Phi = N \cdot \#\Delta $ , where c and $\sigma _W$ are as in §2.5. This condition is satisfied if $(T, U)$ is a Coxeter pair.

We will study the geometric and cohomological properties of $Y_r$ for $r \in \mathbb {Z}_{>0}$ . To this end, we will study Deligne-Lusztig type constructions for various subquotient groups of $\mathbb {G}_r^+$ .

5.1 A total order on affine roots

For $f \in \widetilde \Phi $ , we write $\alpha _f \in \Phi \sqcup \{0\}$ and $m_f \in \mathbb {Z}$ such that $f = \alpha _f + m_f$ . Let $\mathcal {O}_f$ be the F-orbit of f.

Let $\Phi _{\mathrm {aff}}^+$ (resp. $\widetilde \Phi ^+$ ) be the set of affine roots $f \in \Phi _{\mathrm {aff}}$ (resp. $f \in \widetilde \Phi $ ) such that $f({\mathbf x})> 0$ . Note that $\widetilde \Phi ^+ = \Phi _{\mathrm {aff}}^+ \sqcup \mathbb {Z}_{\geqslant 1}$ and $\widetilde \Phi = \Phi _{\mathrm {aff}}^+ \sqcup \widetilde \Phi ^0 \sqcup -\widetilde \Phi _{\mathrm {aff}}^+$ . Here, $\widetilde \Phi ^0 = \{f \in \widetilde \Phi; f({\mathbf x}) = 0\}$ .

Recall that $\Delta = \Phi ^- \cap F\Phi ^+$ . Set $\Delta _{\mathrm {aff}}^+ = (\Delta \times \mathbb {Z}) \cap \Phi _{\mathrm {aff}}^+$ and $\widetilde \Delta ^+ = \Delta _{\mathrm {aff}}^+ \sqcup \mathbb {Z}_{\geqslant 1}$ .

Lemma 5.1. The map $f \mapsto \mathcal {O}_f$ induces a bijection $\widetilde \Delta ^+ \cong \widetilde \Phi ^+ / \langle F\rangle $ .

Proof. This follows from our assumption on $(T, U)$ in this section.

Definition 5.2. We define a linear order $\leqslant $ on $\widetilde \Phi ^+$ such that

  • $f < f'$ if either (1) $f({\mathbf x}) < f'({\mathbf x})$ or (2) $f({\mathbf x}) = f'({\mathbf x})$ , $f \in \mathbb {Z}_{\geqslant 1}$ and $f' \in \Delta _{\mathrm {aff}}^+$ ;

  • if $f_1, f_2 \in \widetilde \Delta ^+$ such that $f_1 < f_2$ , then $f_1' < f_2'$ for any $f_1' \in \mathcal {O}_{f_1}$ and any $f_2' \in \mathcal {O}_{f_2}$ .

  • $f < F(f) < \cdots < F^{N-1}(f)$ for $f \in \Delta _{\mathrm {aff}}^+$ .

Let $f \in \widetilde \Delta ^+$ . We denote by $f+$ and $f-$ the descendant and the ascendant of f in $\widetilde \Delta ^+ \sqcup \{0\}$ , respectively, such that $0 = f-$ if $f = \min \widetilde \Delta ^+$ . Set $\widetilde \Phi ^f = \{f' \in \widetilde \Phi ^+; f' \geqslant f\}$ .

5.2 The variety $Y^A_B$

We fix an integer $r \in \mathbb {Z}_{\geqslant 1}$ . Let $\mathbb {G}_r^+ = \breve {\mathcal {G}}^{0+} / \breve {\mathcal {G}}^r$ and let $\widetilde \Phi _r^+ = \{f \in \widetilde \Phi; 0 < f({\mathbf x}) < r\}$ be the set of affine roots appearing in $\mathbb {G}_r^+$ .

Let $f \in \widetilde \Phi ^+$ . If $f \in \Phi _{\mathrm {aff}}^+$ , we take $\mathbb {A}_f = \mathbb {G}_a$ and define $u_f: \mathbb {A}^1 \to \mathbb {G}_r^+$ by $x \mapsto U_{\alpha _f}([x]\varpi ^{m_f})$ for $x \in \overline {\mathbb {F}}_q$ . If $f \in \mathbb {Z}_{\geqslant 1}$ , we take $\mathbb {A}_f = V := X_*(T) \otimes \overline {\mathbb {F}}_q$ and define $u_f: \mathbb {A}_f \to \mathbb {G}_r^+$ by $\lambda \otimes x \mapsto \lambda (1 + [x] \varpi ^{n_f})$ for $\lambda \in X_*(T)$ and $x \in \overline {\mathbb {F}}_q$ .

We define an abelian group $\mathbb {A}[r] = \prod _{f \in \widetilde \Phi _r^+} \mathbb {A}_f$ . Then we have an isomorphism of varieties

$$\begin{align*}u: \mathbb{A}[r] \overset \sim \to \mathbb{G}_r^+, \quad (x_f)_f \mapsto \prod_f u_f(x_f),\end{align*}$$

where the product is taking with respect to the linear order $\leqslant $ on $\widetilde \Phi ^+$ .

Let $E \subseteq \widetilde \Phi _r^+$ . We set $\mathbb {A}_E = \prod _{f \in E} \mathbb {A}_f$ , which is viewed as a subgroup group of $\mathbb {A}[r]$ in the natural way. Denote by $p_E: \mathbb {A}[r] \to \mathbb {A}_E$ the natural projection. Using the identification $u: \mathbb {A}[r] \cong \mathbb {G}_r^+$ , we define

$$\begin{align*}\mathrm{pr}_E = u \circ p_E \circ u^{-1}: \mathbb{G}_r^+ \to u(\mathbb{A}_E).\end{align*}$$

For $f \in \widetilde \Phi _r^+$ , we put $p_f = p_{\{f\}}$ and $\mathrm {pr}_f = \mathrm {pr}_{\{f\}}$ . By abuse of notation, we will identify $\mathrm {pr}_f: \mathbb {G}_r^+ \to u(\mathbb {A}_f)$ with $u^{-1} \circ \mathrm { pr}_f: \mathbb {G}_r^+ \to \mathbb {A}_f$ freely according to the context.

Let $A, B \subseteq \widetilde \Phi ^+$ be two subsets. We set

$$\begin{align*}A + B = \{f + f' \in \widetilde \Phi; f \in A, f' \in B\}.\end{align*}$$

We say A is closed if $A + A \subseteq A$ and $A + \mathbb {Z}_{\geqslant 0} = A$ . In this case, we denote by $\mathbb {G}_r^A \subseteq \mathbb {G}_r^+$ the subgroup generated by $u(\mathbb {A}_f)$ for $f \in A$ .

Suppose that $\widetilde \Phi ^r \subseteq A, B \subseteq \widetilde \Phi ^+$ are F-stable and closed such that $B \subseteq A$ and $A + B \subseteq B$ . Then $\mathbb {G}_r^B$ is a normal subgroup of $\mathbb {G}_r^A$ . The isomorphism $u: \mathbb {A}[r] \overset \sim \to \mathbb {G}_r^+$ restricts to an isomorphism $u_{A:B}: \mathbb {A}_{A \setminus B} \overset \sim \to \mathbb {G}_r^A / \mathbb {G}_r^B$ . So we get an embedding

$$\begin{align*}s_{A:B} = u \circ u_{A:B}^{-1} : \mathbb{G}_r^A/\mathbb{G}_r^B \to \mathbb{G}_r^+.\end{align*}$$

We define

$$\begin{align*}Y_B^A = \{g \in \mathbb{G}_r^A; g^{-1} F(g) \in (\overline{\mathbb{U}}_r \cap F\mathbb{U}_r) \mathbb{G}_r^B\} / \mathbb{G}_r^B \subseteq \mathbb{G}_r^A / \mathbb{G}_r^B,\end{align*}$$

which admits a natural action by $(\mathbb {G}_r^A)^F \times (\mathbb {T}_r^+ \cap \mathbb {G}_r^A)^F$ . Let $\chi : (\mathbb {T}_r^+ \cap \mathbb {G}_r^A)^F \to \overline {\mathbb {Q}}_\ell ^\times $ be a character. We denote by $H_c^i(Y_B^A, \overline {\mathbb {Q}}_\ell )[\chi ]$ the $\chi $ -weight space of the $(\mathbb {T}_r^+)^F$ -action on $H_c^i(Y_B^A, \overline {\mathbb {Q}}_\ell )$ . For $f \in \widetilde \Phi _r^+$ , we define

$$\begin{align*}\pi^{A:B}_f = u^{-1} \circ \mathrm{ pr}_{\mathcal{O}_f} \circ L \circ s_{A:B}: \mathbb{G}_r^A/\mathbb{G}_r^B \to \mathbb{A}_{\mathcal{O}_f}.\end{align*}$$

Here, for any F-stable sub-quotient group of $\mathbb {G}_r^+$ , we always denote by L the Lang’s self-map given by $g \mapsto g^{-1} F(g)$ .

Proposition 5.3. Let $\widetilde \Phi ^r \subseteq A, B \subseteq \widetilde \Phi ^+$ be F-stable and closed. Let $f \in B$ and $C = B \setminus \mathcal {O}_f$ . Suppose that $\widetilde \Phi ^r \subseteq C$ is closed, $C + A \subseteq C$ and $\mathcal {O}_f + A \subseteq C$ . Then

(1) if $f \in \Delta _{\mathrm {aff}}^+$ , then the map $\psi = (q_f, \mathrm {pr}_f): Y_C^A \cong Y_B^A \times \mathbb {A}_f$ is an isomorphism;

(2) if $f \in \mathbb {Z}_{\geqslant 1}$ (in which case $\mathbb {A}_{\mathcal {O}_f} = \mathbb {A}_f = V$ ), then there is a Cartesian diagram

Here, $q_f$ denotes the natural projection.

Proof. By assumption, the map u induces an identification $\mathbb {A}_{\mathcal {O}_f} \cong \mathbb {G}_r^B/\mathbb {G}_r^C$ as abelian groups. Moreover,

(a) $\mathbb {G}_r^B/\mathbb {G}_r^C$ lies in the center of $\mathbb {G}_r^A/\mathbb {G}_r^C$ .

Assume that $f \in \Delta _{\mathrm {aff}}^+$ . We define a morphism $\phi : Y_B^A \times \mathbb {A}_f \to Y_C^A$ as follows. Let $(g, y) \in Y_B^A \times \mathbb {A}_f$ . Write $\pi ^{A:B}_f(g) = (z_i)_{1 \leqslant i \leqslant N} \in \mathbb {A}_{\mathcal {O}_f}$ with each $z_i \in \mathbb {A}_{F^i(f)}$ . We define

$$\begin{align*}\phi(g, y) = s_{A:B}(g) u(y) F(u(y)) \cdots F^{N-1}(u(y)) \prod_{1 \leqslant i \leqslant N-1} u(z_i) F(u(z_i)) \cdots F^{N-i-1}(u(z_i)).\end{align*}$$

By (a), one checks that

$$\begin{align*}\phi(Y_B^A \times \mathbb{A}_f) \subseteq Y_C^A\end{align*}$$

and $\psi \circ \phi = \mathrm {id}$ . Let $g \in Y_C^A$ and set $g' = \phi (\psi (g)) \in Y_C^A$ . Then $\psi (g) = \psi (g')$ ; that is, $g^{-1} g' \in \mathbb {A}_{\mathcal {O}_f \setminus \{f\}} \subseteq \mathbb {G}_r^B / \mathbb {G}_r^C$ . As $g, g' \in Y_C^A$ , it follows by (a) that

$$\begin{align*}L(g^{-1} g') = L(g)^{-1} L(g') \in \mathbb{A}_f \subseteq \mathbb{G}_r^B / \mathbb{G}_r^C.\end{align*}$$

Hence, $g = g'$ by Lemma 5.4. So $\phi \circ \psi = \mathrm {id}$ and (1) is proved.

Assume that $f \in \mathbb {Z}_{\geqslant 1}$ . As both vertical maps in the diagram are finite étale $V^F$ -torsors, it suffices to show that the square commutes. Let $g \in Y_C^A$ . Write $s_{A:C}(g) = u(x) u(y)$ with $x \in \mathbb {A}_{A \setminus B}$ and $y \in \mathbb {A}_f$ . Then $\mathrm {pr}_f(g) = y$ and $q_f(g) = u(x)$ . As $f \in \mathbb {Z}_{\geqslant 1}$ and $g \in Y_C^A$ , we have $\mathrm {pr}_f (L(s_{A:C}(g)) = 0 \in \mathbb {A}_f$ . Using (a), one computes that

$$ \begin{align*} \pi_f^{A:B}(q_f(g)) &= \mathrm{pr}_f(L(u(x))) \\ &= \mathrm{pr}_f(L(s_{A:C}(g) u(y)^{-1})) \\ &= \mathrm{pr}_f(L(s_{A:C}(g)) L(u(y)^{-1})) \\ &= \mathrm{pr}_f(L(s_{A:C}(g))) \mathrm{pr}_f(L(u(y)^{-1})) \\ &= L(y^{-1}) \\ &= - L(y). \end{align*} $$

So (2) is proved.

Lemma 5.4. Let $f \in \Delta _{\mathrm {aff}}^+$ and let $x = (x_i)_{0 \leqslant i \leqslant N-1} \in \mathbb {A}_{\mathcal {O}_f}$ with each $x_i \in \mathbb {A}_{F^i(f)}$ such that $L(x) \in \mathbb {A}_f$ . Then $x_i = F^i(x_0)$ for $1 \leqslant i \leqslant N-1$ . In particular, (1) $L(x) = F^N(x_0) - x_0$ and (2) $x = 0$ if and only if $x_0 = 0$ .

Proof. By definition, we have

$$\begin{align*}L(x) = F(x) - x = \sum_{i=0}^{N-1}F(x_{i-1}) - x_i \in \mathbb{A}_{\mathcal{O}_f},\end{align*}$$

from which the lemma follows.

5.3 Main result

For $f' \leqslant f \in \widetilde \Delta ^+$ , we set $\mathbb {G}_f^+ = \mathbb {G}_r^+ / \mathbb {G}_r^{\widetilde \Phi ^f}$ , $Y_f = Y^{\widetilde \Phi ^+}_{\widetilde \Phi ^f}$ , $\mathbb {T}_f = \mathbb {T}_{\lceil f \rceil }$ and $\mathbb {T}_f^{f'} = \ker (\mathbb {T}_f \to \mathbb {T}_{f'})$ , where $\widetilde \Phi ^f = \{f' \in \widetilde \Phi ^+; f' \geqslant f\}$ and $\lceil f \rceil = \min \{n \in \mathbb {Z}_{\geqslant 1}, n \geqslant f\}$ . Note that $\mathbb {T}_{f+}^{f}$ is nontrivial if and only if $f \in \mathbb {Z}_{\geqslant 1}$ , in which case $\mathbb {T}_{f+}^f \cong V = X_*(T) \otimes \overline {\mathbb {F}}_q$ .

Theorem 5.5. Assume that p satisfies Condition 2.1. Let $f \in \widetilde \Delta ^+$ and let $\chi : (\mathbb {T}_f^+)^F \to \overline {\mathbb {Q}}_\ell ^\times $ be a character. Then there exists $s = s_{f, \chi } \in \mathbb {Z}_{\geqslant 0}$ such that

$$\begin{align*}H_c^i(Y_f, \overline{\mathbb{Q}}_\ell)[\chi] \neq 0 \Longleftrightarrow i = s,\end{align*}$$

on which $F^N$ acts by multiplication by $(-1)^s q^{sN/2}$ .

After necessary preparations, we prove Theorem 5.5 in §5.7. We compute the cohomological degree $s_{\chi ,r}$ explicitly in terms of the Howe factorization of $\chi $ in §5.8. We will make use of the following well-known result.

Theorem 5.6. Let $f: Z \to Y$ be an étale $\Gamma $ -torsor, where $\Gamma $ is a finite group. Let $\Lambda $ be a ring. Assume that either $\Lambda $ is finite, or $Z, Y$ are irreducible and geometrically unibranch. Then

$$\begin{align*}f_! (\Lambda) = \bigoplus_\rho \rho \otimes \mathcal{E}_\rho,\end{align*}$$

where $\rho $ ranges over irreducible representations of $\Gamma $ and $\mathcal {E}_\rho $ is a local system on Y.

Proof. The category of locally constant $\Lambda $ -sheaves on $Z_{\mathrm {et}}$ is equivalent to the category of continuous $\pi _1(Z)$ -representations on finite $\Lambda $ -modules [27, 0GIY, 0DV5]. The same holds for Y and the functor $f_! = f_\ast $ correspond to induction of representations. Thus, $f_!(\Lambda )$ corresponds to the $\pi _1(Y)$ -representation $\mathrm {ind}_{\pi _1(Z)}^{\pi _1(Y)} 1_{\pi _1(Z)}$ , which is equal to the inflation along $\pi _1(Y) \twoheadrightarrow \Gamma $ of the regular $\Gamma $ -representation. The latter decomposes as $\bigoplus _{\rho \in \mathrm { Irr}(\Gamma )} \rho ^{\oplus \dim (\rho )}$ . Thus, if $\mathcal {E}_\rho $ denotes the local system on Y corresponding to the inflation of $\rho $ , we deduce $f_!(\Lambda ) \cong \oplus _{\rho } \mathcal {E}_\rho ^{\oplus \dim (\rho )}$ .

Proposition 5.7. Let $\Gamma $ be a finite group. Suppose that Z and $Y = Z \times \mathbb {A}^1$ are $\Gamma $ -varieties, and the natural projection $\pi : Y \to Z$ is $\Gamma $ -equivariant. Then we have $H_c^i(Y, \overline {\mathbb {Q}}_\ell ) \cong H_c^{i-2}(Z, \overline {\mathbb {Q}}_\ell )$ as $\Gamma $ -modules.

Proof. It suffices to show $\pi _!(\overline {\mathbb {Q}}_\ell ) \cong \overline {\mathbb {Q}}_\ell [-2]$ as $\Gamma $ -equivariant sheaves. Indeed, the adjunction map gives an isomorphism

$$\begin{align*}\pi_!(\overline{\mathbb{Q}}_\ell) \cong \pi_!\pi^*(\overline{\mathbb{Q}}_\ell) \cong \pi_!\pi^!(\overline{\mathbb{Q}}_\ell [-2]) \cong \overline{\mathbb{Q}}_\ell [-2]\end{align*}$$

as $\Gamma $ -equivariant sheaves.

5.4 Multiplicative local systems

Let P be a commutative unipotent algebraic group defined over $\mathbb {F}_q$ . Then the map $\mathcal {L} \mapsto t_{\mathcal {L}}$ induces a bijection from the isomorphism classes of multiplicative local systems on P to the set $\operatorname {\mathrm {Hom}}(H(\mathbb {F}_q), \overline {\mathbb {Q}}_\ell ^\times )$ of characters of $P(\mathbb {F}_q)$ . Here, $t_{\mathcal {L}}: P(\mathbb {F}_q) \to \overline {\mathbb {Q}}_\ell ^\times $ is the trace-of-Frobenius function for $\mathcal {L}$ . See [Reference Boyarchenko and Drinfeld1, §1.8] for details. For $\theta \in \operatorname {\mathrm {Hom}}(P(\mathbb {F}_q), \overline {\mathbb {Q}}_\ell ^\times )$ , we denote by $\mathcal {L}_\theta $ the multiplicative local system corresponding to $\theta $ .

Lemma 5.8. Let $\mathcal {L}$ be a multiplicative local system on P. Then the base change of $\mathcal {L}$ to $P_{\mathbb {F}_{q^n}}$ (with $n \in \mathbb {Z}_{\geq 1}$ ) corresponds to the character $t_{\mathcal {L}} \circ \mathrm {Nm}_n$ , where $\mathrm {Nm}_n(x) = x F_P(x) \cdots F_P^{n-1}(x)$ and $F_P$ denotes the Frobenius automorphism of P.

For a character $\chi $ of $(\mathbb {T}_{f+}^+)^F$ , we denote by $\chi _{f+}^f$ the restriction of $\chi $ to $(\mathbb {T}_{f+}^f)^F$ . Proposition 5.3 has the following consequence:

Corollary 5.9. Let $f \in \widetilde \Delta ^+$ and let $\chi $ be a character of $(\mathbb {T}_{f+}^+)^F$ .

(1) if $f \in \Delta _{\mathrm {aff}}^+$ , then $H_c^i(Y_{f+}, \overline {\mathbb {Q}}_\ell )[\chi ] \cong H_c^{i-2}(Y_f, \overline {\mathbb {Q}}_\ell )[\chi ]$ ;

(2) if $f \in \mathbb {Z}_{\geqslant 1}$ , then $H_c^i(Y_{f+}, \overline {\mathbb {Q}}_\ell )[\chi _{f+}^f] \cong H_c^i(Y_f, \pi ^*(\mathcal {L}_{\chi _{f+}^f}))$ , and hence,

$$\begin{align*}H_c^i(Y_{f+}, \overline{\mathbb{Q}}_\ell)[\chi] \cong H_c^i(Y_f, \pi^*(\mathcal{L}_{\chi_{f+}^f}))[\chi].\end{align*}$$

Here, $\pi = \pi ^{\widetilde \Phi ^+:\widetilde \Phi ^f}_f$ and $H_c^i(Y_{f+}, \overline {\mathbb {Q}}_\ell )[\chi _{f+}^f]$ is the $\chi _{f+}^f$ -weight space of $(\mathbb {T}_{f+}^f)^F$ .

Proof. If $f \in \Delta _{\mathrm {aff}}^+$ , by Lemma 5.3 (1), we have $Y_{f+} \cong Y_f \times \mathbb {G}_a$ , and the natural projection $q_f: Y_{f+} \to Y_f$ respects the right actions of $(\mathbb {T}_{f+}^+)^F = (\mathbb {T}_f^+)^F$ on $Y_{f+}$ and $Y_f$ . So the statement (1) follows from Proposition 5.7.

Now assume that $f \in \mathbb {Z}_{\geqslant 1}$ . Note that the Lang’s map $L: \mathbb {T}_{f+}^f \to \mathbb {T}_{f+}^f$ is an étale $(\mathbb {T}_{f+}^f)^F$ -torsor. It follows from Theorem 5.6 that

$$\begin{align*}L_! (\overline{\mathbb{Q}}_\ell) = \bigoplus_\theta \mathcal{L}_\theta,\end{align*}$$

where $\theta $ ranges over characters of $(\mathbb {T}_{f+}^f)^F$ , and $\mathcal {L}_\theta $ is the multiplicative local system corresponding to $\theta $ . By the Cartesian diagram in Proposition 5.3 (2), it follows from the base change theorem that

$$\begin{align*}(q_f)_!(\overline{\mathbb{Q}}_\ell) = (q_f)_! \mathrm{pr}_f^* (\overline{\mathbb{Q}}_\ell) \cong \pi_f^* L_!(\overline{\mathbb{Q}}_\ell) = \bigoplus_\theta \pi_f^* \mathcal{L}_\theta.\end{align*}$$

So the statement (2) follows by noticing that $(\mathbb {T}_{f+}^f)^F$ acts on the sheaf $\mathcal {L}_\theta $ via the character  $\theta $ .

From this we deduce the following:

Corollary 5.10. Let $\chi \colon \mathcal {T}(\mathcal {O}_k) \rightarrow \overline {\mathbb {Q}}_\ell ^\times $ be a smooth character, which factors through $(\mathbb {T}_r^+)^F$ . Then for any $r_2 \geq r_1 \geq r$ , the map $f_{r_2,!}\overline {\mathbb {Q}}_\ell [\chi ][2(\dim Y_{r_2} - \dim Y_{r_1})] \rightarrow f_{r_1,!}\overline {\mathbb {Q}}_\ell [\chi ]$ is an isomorphism, where $f_{r_i} \colon Y_{r_i} \rightarrow \operatorname {\mathrm {Spec}}\overline {\mathbb {F}}_q$ is the structure map. With other words, (2.1) holds for the schemes $Y_r$ .

5.5 Reduction to the semisimple case

Let $G' \subseteq G$ be the derived subgroup. Let $T'$ be a maximal torus of $G'$ contained in T. One can define the objects $Y_f' = Y_f$ for $G'$ in a similar way.

Lemma 5.11. For $f \in \widetilde \Delta ^+$ , we have

$$\begin{align*}Y_f = \bigsqcup_{x \in (\mathbb{T}_f^+)^F/ (\mathbb{T}_f^{\prime +})^F} x Y_f' = \bigsqcup_{x \in (\mathbb{T}_f^+)^F / (\mathbb{T}_f^{\prime +})^F} Y_f' x^{-1}.\end{align*}$$

In particular, $H_c^i(Y_f, \overline {\mathbb {Q}}_\ell ) \cong \mathrm { ind}_{(\mathbb {T}_f^{\prime +})^F}^{(\mathbb {T}_f^+)^F} H_c^i(Y_f', \overline {\mathbb {Q}}_\ell )$ as $(\mathbb {T}_f^+)^F$ -modules.

Proof. Let $g \in Y_f$ . Then $g^{-1} F(g) \in \overline {\mathbb {U}}_f^+ \cap F\mathbb {U}_f^+ \subseteq \mathbb {G}_f^{\prime +}$ . By Lang’s theorem, there exists $g' \in \mathbb {G}_f^{\prime +}$ such that ${g'}^{-1} F(g') = g^{-1} F(g)$ . So $g = (g {g'}^{-1}) g' \in (\mathbb {G}_f^+)^F Y_f'$ , and hence, $Y_f = (\mathbb {G}_f^+)^F Y_f'$ .

However, there is a natural isomorphism

$$\begin{align*}(\mathbb{T}_f^+)^F / (\mathbb{T}_f^{\prime +})^F \cong (\mathbb{G}_f^+)^F / (\mathbb{G}_f^{\prime +})^F.\end{align*}$$

Now it follows that

$$\begin{align*}Y_f^+ = (\mathbb{G}_f^+)^F Y_f' = \bigsqcup_{x \in (\mathbb{T}_f^+)^F/ (\mathbb{T}_f^{\prime +})^F} x Y_f^{\prime +} = \bigsqcup_{x \in (\mathbb{T}_f^+)^F / (\mathbb{T}_f^{\prime +})^F} Y_f^{\prime +} x^{-1},\end{align*}$$

where the last equality follows from the observation that $(\mathbb {T}_f^+)^F$ normalizes $Y_f'$ .

5.6 Handling jumps in the Howe factorization of $\chi $

We fix a positive integer $h \leqslant r$ and a character $\chi $ of $(\mathbb {T}_{h+}^+)^F $ . Recall that $\mathbb {T}_{h+}^h \cong \mathbb {A}_h = V = X_*(T) \otimes \overline {\mathbb {F}}_q$ , and recall from §2.6 the norm map

$$\begin{align*}\mathrm{Nm}_N: V \to V, ~ v \mapsto v + F(v) + \cdots + F^{N-1}(v). \end{align*}$$

Using the character $\chi $ , we define a root system

$$\begin{align*}\Phi_\chi = \{\alpha \in \Phi; \chi \circ \mathrm{Nm}_N (\alpha^\vee \otimes \mathbb{F}_{q^N}) = \{1\}\}.\end{align*}$$

Note that $\Phi _\chi $ is F-stable. By [Reference Kaletha21, Lemma 3.6.1], it is a Levi subsystem of $\Phi $ (note that by Condition 2.1, p is not a torsion prime for $\Phi $ ).

Let $M = M_\chi \subseteq G$ be the twisted Levi subgroup generated by T and $U_\alpha $ for $\alpha \in \Phi _\chi $ . Let $\widetilde \Phi _M$ be the set of affine roots of M. We set

$$\begin{align*}D = (\Delta_{\mathrm{aff}}^+ \cap \Phi_h^+) \setminus \widetilde \Phi_M = \{f \in \Delta_{\mathrm{aff}}^+ \setminus \widetilde \Phi_M; f < h\}.\end{align*}$$

By Lemma 5.1, the map $f \mapsto \mathcal {O}_f$ gives a natural bijection

$$\begin{align*}D \overset \sim \to (\widetilde \Phi_h^+ \setminus \widetilde \Phi_M) / \langle F \rangle.\end{align*}$$

Let $f \in D$ . As $f < h$ , it follows by Definition 5.2 that $0 < f({\mathbf x}) < h$ , and hence, $h-f \in \widetilde \Phi _h^+ \setminus \widetilde \Phi _M$ . Hence, there exists a unique affine root $f^\flat \in \Delta _{\mathrm {aff}}^+$ such that $-f+h \in \mathcal {O}_{f^\flat }$ . In particular, $f^\flat \in D$ and $f({\mathbf x}) + f^\flat ({\mathbf x}) = h$ . We label all the affine roots in D by

$$\begin{align*}f_1, \dots, f_{m-1}, f_m = f_m^\flat, \dots, f_n = f_n^\flat, f_{m-1}^\flat, \dots, f_1^\flat\end{align*}$$

such that

$$\begin{align*}f_1({\mathbf x}) \leqslant \cdots \leqslant f_{m-1}(x) \leqslant \frac{h}{2} = f_m({\mathbf x}) = \cdots = f_n({\mathbf x}) = \frac{h}{2} \leqslant f_{m-1}^\flat({\mathbf x}) \leqslant \cdots \leqslant f_1^\flat({\mathbf x}),\end{align*}$$

$f_i < f_i^\flat $ for $1 \leqslant i \leqslant m-1$ and $f_{m-1}^\flat < \cdots < f_1^\flat $ .

Let $1 \leqslant i \leqslant m$ . We set $D_i^\flat = \{f_j^\flat \in D; 1 \leqslant j \leqslant i\}$ if $1 \leqslant i \leqslant m-1$ and $D_i^\flat = \{f_j^\flat; 1 \leqslant j \leqslant n\}$ if $i=m$ . Define

$$\begin{align*}A_i = \widetilde \Phi^+ \setminus \cup_{j=1}^{i-1} \mathcal{O}_{f_j}, \quad B_i = \widetilde \Phi^h \cup \bigcup_{f \in D_i^\flat} \mathcal{O}_f, \quad C_{i-1} = B_{i-1} \setminus \{h\}.\end{align*}$$

Moreover, we set $A_0 = A_1 = \widetilde \Phi ^+$ , $B_0 = \widetilde \Phi ^h$ and $C_0 = B_0 \setminus \{h\}$ . Note that $A_m = B_m \cup \widetilde \Phi _M^+$ with $\widetilde \Phi _M^+ = \widetilde \Phi _M \cap \widetilde \Phi ^+$ .

Lemma 5.12. Let $1 \leqslant i \leqslant m$ . Then $A_{i-1} + A_{i-1} \subseteq A_i$ , $A_i + B_i \subseteq B_{i-1}$ , $\widetilde \Phi _M^+ + B_i \subseteq C_{i-1}$ , $C_{i-1} + C_{i-1} \subseteq C_{i-1}$ and $A_{i+1} + B_i \subseteq C_{i-1}$ , where $A_{m+1} = B_{m-1} \cup \widetilde \Phi _M^+$ . In particular, $A_i$ , $B_i$ and $C_{i-1}$ are F-stable and closed.

Proof. We only show the second and the third inclusions. The others can be proved similarly. Let $f \in A_i$ and $f' \in B_i$ such that $f + f' \in \widetilde \Phi $ .

First, we assume that $f \in \widetilde \Phi _M^+$ . Then $f+f' \notin \widetilde \Phi _M^+$ since $f' \notin \widetilde \Phi _M^+$ . As

$$\begin{align*}(f+f')({\mathbf x})> f'({\mathbf x}) \geqslant f_i^\flat({\mathbf x}) \geqslant h/2,\end{align*}$$

we have $f + f' \in \cup _{f" \in D_{i-1}^\flat } \mathcal {O}_{f"} \subseteq C_{i-1} \subseteq B_{i-1}$ , as desired.

Now we assume that $f \notin \widetilde \Phi _M^+$ . Then $f({\mathbf x}) \geqslant f_i({\mathbf x})$ and

$$\begin{align*}(f+f')({\mathbf x}) \geqslant f_i({\mathbf x}) + f_i^\flat({\mathbf x}) = h.\end{align*}$$

By Definition 5.2, we have $f + f' \in \widetilde \Phi ^h \subseteq B_{i-1}$ , as desired.

Let $g \in \mathbb {G}_r^+$ , $x \in \mathbb {A}[r]$ and $E \subseteq \widetilde \Phi _r^+$ . We set $g_E = \mathrm {pr}_E(g) \in u(\mathbb {A}_E)$ , $x_E = p_E(x) \in \mathbb {A}_E$ and $\hat x = u(x) \in \mathbb {G}_r^+$ . For $f \in \widetilde \Phi _r^+$ , we will set $x_f = x_{\{f\}}$ and $x_{\geqslant f} = x_{\widetilde \Phi ^f}$ . We can define $g_f$ and $g_{\geqslant f} \in \mathbb {G}_r^+$ in a similar way. By abuse of notation, we will identify $g_f \in u(\mathbb {A}_f)$ with $u^{-1}(g_f) \in \mathbb {A}_f$ according to the context.

Lemma 5.13. Let $A, B \subseteq \widetilde \Phi ^+$ such that $A + B \subseteq \widetilde \Phi ^h$ . Let $x \in \mathbb {A}_A$ and $y \in \mathbb {A}_B$ . Then

$$\begin{align*}(\hat y \hat x)_h = -\sum_f \alpha_f^\vee \otimes y_{h-f} x_f + y_h + x_h \in \mathbb{A}_h = V,\end{align*}$$

where f ranges over A such that $f < h-f$ .

Proof. As $A + B \subseteq \widetilde \Phi ^h$ , for any $f \in A$ and $f' \in B$ , we have $[\hat y_{f'}, \hat x_f] = \hat y_{f'} \hat x_f \hat y_{f'}^{-1} \hat x_f^{-1} \in \mathbb {G}_r^h$ . Moreover, one computes that

(a) $[\hat y_{f'}, \hat x_f]_h = \alpha _f^\vee \otimes y_{f'} x_f$ if $f + f' = h$ , and $[\hat y_{f'}, \hat x_f]_h= 0$ otherwise.

Assume that $y = y_{\leqslant f'}$ and $x = x_{\geqslant f}$ for some $f' \in B$ and $f \in A$ . We argue by induction on f. If $f \geqslant f'$ , the statement is trivial. Suppose that $f < f'$ . Then we have

$$ \begin{align*} (\hat y \hat x)_h &= (\hat y_{< f'} \hat y_{f'} \hat x_f \hat x_{> f})_h \\ &= (\hat y_{< f'} \hat x_f \hat y_{f'} [\hat y_{f'}^{-1}, \hat x_f^{-1}] \hat x_{> f})_h \\ &=(\hat y_{< f'} \hat x_f \hat y_{f'} \hat x_{>f} )_h + [\hat y_{f'}^{-1}, \hat x_f^{-1}]_h \\ &~\vdots \\ &= (\hat y_{\leqslant f} \hat x_f \hat y_{[f+, f']} \hat x_{>f})_h + \sum_{f" \in [f+, f']} [\hat y_{f"}^{-1}, \hat x_f^{-1}]_h \\ &= (\hat y_{\leqslant f} \hat x_f)_h + (\hat y_{[f+, f']} \hat x_{>f})_h + \sum_{f" \in [f+, f']} [\hat y_{f"}^{-1}, \hat x_f^{-1}]_h, \end{align*} $$

where $[f+, f'] = \{f" \in \widetilde \Phi ^+; f+ \leqslant f" \leqslant f'\}$ . Now the statement follows from (a) and induction hypothesis.

Lemma 5.14. Let $1 \leqslant i \leqslant m$ . Let $x \in \mathbb {A}_{A_i}$ and $y \in \mathbb {A}_{B_i}$ . Assume that $x \in \mathbb {A}_{A_i \cap \widetilde \Phi _M}$ or $1 \leqslant i \leqslant m-1$ . Then $(\hat x \hat y)_h = x_h + y_h \in \mathbb {A}_h$ .

Proof. Assume that $x = x_{\leqslant f}$ and $y = y_{\geqslant f'}$ for some $f \in A_i$ and $f' \in B_i$ . We argue by induction on $f'$ . If $f \leqslant f'$ , the statement is trivial. Assume that $f> f'$ . We claim that

(a) $f + f'> h$ if $f + f' \in \widetilde \Phi $ .

First, note that $f({\mathbf x}) + f'({\mathbf x}) \geqslant 2f'({\mathbf x}) \geqslant 2f_i^\flat ({\mathbf x}) \geqslant h$ . Suppose that (a) does not hold. Then $f + f' = h$ . Assume $x \in \mathbb {A}_{A_i \cap \widetilde \Phi _M}$ . Then $f \in \widetilde \Phi _M^+$ and $f + f' \in C_{i-1}$ by Lemma 5.12, which is a contradiction. Assume $1 \leqslant i \leqslant m-1$ . If $f \in \mathcal {O}_{f_i}$ , then $f' \in \mathcal {O}_{f_i^\flat }$ and hence $f < f'$ by our choice that $f_i < f_i^\flat $ , which is a contradiction. So $f \in A_{i+1}$ and $f + f' \in A_{i+1} + B_i \subseteq C_i$ by Lemma 5.12, which is also a contradiction. So (a) is proved.

By (a), we have $[\hat x_f^{-1}, \hat y_{f'}^{-1}] \in \mathbb {G}_r^{h+}$ . Hence,

$$ \begin{align*} (\hat x \hat y)_h &= ((\hat x_{< f} \hat y_{f'} \hat x_f [\hat x_f^{-1}, \hat y_{f'}^{-1}]) \hat y_{>f'})_h \\ &= (\hat x_{<f} \hat y_{f'} \hat x_f \hat y_{>f'})_h \\ &~\vdots \\ &= (\hat x_{\leqslant f'} \hat y_{f'} \hat x_{[f'+, f]} \hat y_{>f'})_h \\ &= (\hat x_{\leqslant f'} \hat y_{f'})_h + (\hat x_{[f'+, f]} \hat y_{>f'})_h \\ &= (\hat x_{\leqslant f'})_h + (\hat y_{f'})_h + (\hat x_{[f'+, f]} \hat y_{>f'})_h. \end{align*} $$

Now the statement follows by induction hypothesis.

We set $\pi = \pi _h^{\widetilde \Phi ^+: \widetilde \Phi ^h}: \mathbb {G}_h^+ = \mathbb {G}_r^+/\mathbb {G}_r^h \to \mathbb {A}_h \cong V$ .

Proposition 5.15. Let $1 \leqslant i \leqslant m$ . Then there is an isomorphism

$$\begin{align*}\psi_i: Y^{A_i}_h \cong Y^{A_i}_{B_i} \times \mathbb{A}_{D_i^\flat}.\end{align*}$$

Moreover, for $(\hat x, y) \in Y^{A_i}_{B_i} \times \mathbb {A}_{D_i^\flat }$ with $x \in \mathbb {A}_{A_i \setminus B_i}$ we have

(1) if $1 \leqslant i \leqslant m-1$ , then

$$\begin{align*}\pi(\psi_i^{-1}(\hat x, y)) = \alpha_{f_i}^\vee \otimes (x_{f_i}^{q^N} - x_{f_i}) y_{f_i^\flat}^{q^{n_i}} + \pi(\psi_i^{-1}(\hat x, 0)) \in V,\end{align*}$$

where $0 \leqslant n_i \leqslant N-1$ such that $F^{n_i}(f_i^\flat ) = -f_i + h$ ;

(2) if $i=m$ , then $\pi (\psi _i^{-1}(\hat x, y)) = -\sum _{j=m}^n \alpha _{f_j}^\vee \otimes y_{f_j}^{q^{N/2}+1} + \pi (\psi _i^{-1}(\hat x, 0))$ .

Proof. Without loss of generality, we may assume that $m = n$ . In particular, $B_i = \widetilde \Phi ^h \cup \mathcal {O}_{f_1^\flat } \cup \cdots \cup \mathcal {O}_{f_i^\flat }$ for $0 \leqslant i \leqslant m$ .

By Lemma 5.12, we have $A_i + \mathcal {O}_{f_j^\flat } \subseteq A_j + B_j \subseteq B_{j-1}$ for $1 \leqslant j \leqslant i \leqslant m$ . Thus, by applying Proposition 5.3 (1) repeatedly, we obtain an isomorphism

$$\begin{align*}\psi_i: Y^{A_i}_h = Y^{A_i}_{B_0} \cong Y^{A_i}_{B_1} \times \mathbb{A}_{f_1^\flat} \cong \cdots \cong Y^{A_i}_{B_i} \times \mathbb{A}_{D_i^\flat}.\end{align*}$$

Let $z = s_{\widetilde \Phi ^+: \widetilde \Phi ^h} \circ \psi _i^{-1}(\hat x, y)$ . We claim that

(a) $z = \hat x \hat w$ for some $w \in \mathbb {A}_{B_i \setminus \widetilde \Phi ^h}$ such that $w_{F^j(f_i^\flat )} = y_{f_i^\flat }^{q^j} + P_j(x)$ for $0 \leqslant j \leqslant N-1$ , where each $P_j$ is a polynomial function on $\mathbb {A}_{A_i \setminus B_i}$ . Moreover, $P_j = 0$ if $i = m$ .

Indeed, the first claim follows from the Proposition 5.3. Moreover, if $i = m$ , then $A_i \setminus B_i \subseteq \widetilde \Phi _M$ and $L(\hat x) \in \mathbb {M}_r^+$ . Hence, $P_j = 0$ for $1 \leqslant j \leqslant N-1$ by Proposition 5.3. So (a) is proved.

Then we claim that

(b) $x_{F^j(f_i)} = x_{f_i}^{q^j}$ for $1 \leqslant i \leqslant m-1$ and $0 \leqslant j \leqslant N-1$ .

Indeed, Let $v = x_{\mathcal {O}_{f_i}} \in \mathbb {A}_{\mathcal {O}_{f_i}}$ . As $\hat x \in Y^{A_i}_{B_i}$ , we have $\hat v \in Y^{A_i}_{A_{i+1}} \subseteq \mathbb {G}_r^{A_i} / \mathbb {G}_r^{A_{i+1}} \cong \mathbb {A}_{\mathcal {O}_{f_i}}$ ; that is, $L(\hat v) \in \mathbb {A}_{f_i} \subseteq \mathbb {A}_{\mathcal {O}_{f_i}}$ . Now (b) follows from Lemma 5.4.

Assume that $1 \leqslant i \leqslant m-1$ . By (b), we have

(c) $L(\hat x)_f = 0 $ if $f \in \mathcal {O}_{f_i} \setminus \{f_i\}$ and $L(\hat x)_f = x_{f_i}^{q^N} - x_{f_i}$ if $f = f_i$ .

Note that $\hat w, F(\hat w) \in \mathbb {G}_r^{B_i}$ , $\mathcal {O}_{f_i} < \mathcal {O}_{f_i^\flat }$ and $B_i + B_i \subseteq \widetilde \Phi ^{h+}$ . Moreover, $L(\hat x) \in \mathbb {G}_r^{A_i}$ and $[\hat w, (L(\hat x)_{<f_i})^{-1}] \in [\mathbb {G}_r^{B_i}, \mathbb {G}_r^{A_{i+1}}] \subseteq \mathbb {G}_r^{C_{i-1}}$ . It follows from Lemma 5.13 and Lemma 5.14 that $(\hat w^{-1})_h = 0$ and

$$ \begin{gather*} (L(\hat x)_{\geqslant f_i} F(\hat w))_h = (L(\hat x)_{\geqslant f_i})_h + F(\hat w)_h = L(\hat x)_h; \\ (\hat w^{-1} [\hat w, (L(\hat x)_{< f_i})^{-1}])_h = (\hat w^{-1})_h + ([\hat w, (L(\hat x)_{< f_i})^{-1}])_h = 0.\end{gather*} $$

Then one computes that

$$ \begin{align*} \pi(\psi_i^{-1}(\hat x, y)) &= (\hat w^{-1} L(\hat x) F(\hat w))_h \\ &=(L(\hat x)_{<f_i} \hat w^{-1} [\hat w, (L(\hat x)_{<f_i})^{-1}] L(\hat x)_{\geqslant f_i} F(\hat w))_h \\ &= (\hat w^{-1} [\hat w, (L(\hat x)_{<f_i})^{-1}] L(\hat x)_{\geqslant f_i} F(\hat w))_h \\ &= \sum_{f \in \mathcal{O}_{f_i}} ((\hat w^{-1})_{h-f} L(\hat x)_f)_h + (\hat w^{-1} [\hat w, (L(\hat x)_{< f_i})^{-1}])_h + (L(\hat x)_{\geqslant f_i} F(\hat w))_h \\ &= ((\hat w^{-1})_{h-f_i} L(\hat x)_{f_i})_h + L(\hat x)_h \\ &= \alpha_{f_i}^\vee \otimes ((x_{f_i}^{q^N} - x_{f_i}) (y_{f_i^\flat}^{q^{n_i}} + P_{n_i}(x))) + L(\hat x)_h \\ &= \alpha_{f_i}^\vee \otimes (x_{f_i}^{q^N} - x_{f_i}) y_{f_i^\flat}^{q^{n_i}} + \pi(\psi_i^{-1}(\hat x, 0), \end{align*} $$

where the third equality follows from that $f_i < B_i$ , the fourth equality follows from Lemma 5.13, and the fifth equality follows from (c).

Assume $i = m$ . Then $\hat x, L(\hat x) \in \mathbb {M}_r^+$ and $F^{N/2}(f_i) = h - f_i$ . Moreover, by the second statement of (a), we have

$$ \begin{align*} \hat w_{\mathcal{O}_{f_i}}^{-1} F(\hat w_{\mathcal{O}_{f_i}}) &= (\hat w_{f_i} \cdots F^{N-1}(\hat w_{f_i}))^{-1} F(\hat w_{f_i}) \cdots F^{N-1}(\hat w_{f_i}) \\ &\equiv \hat w_{f_i}^{-1} F^N(\hat w_{f_i}) [\hat w_{f_i}, F^{N/2}(\hat w_{f_i})^{-1}] \quad\mod \mathbb{G}_r^{h+}. \end{align*} $$

Thus,

$$\begin{align*}(\hat w^{-1} F(\hat w))_h = (\hat w_{\mathcal{O}_{f_i}}^{-1} F(\hat w_{\mathcal{O}_{f_i}}))_h = [\hat w_{f_i}, F^{N/2}(\hat w_{f_i})^{-1}]_h = -\alpha_{f_i}^\vee \otimes y_{f_i}^{q^{N/2} + 1}.\end{align*}$$

As $L(\hat x) \in \mathbb {M}_r^+$ , we have $[\hat w, L(\hat x)^{-1}] \in \mathbb {G}_r^{C_{i-1}}$ . In particular, $[\hat w, L(\hat x)^{-1}]_h = 0$ and $[[\hat w, L(\hat x)^{-1}], F(\hat w)] \in \mathbb {G}_r^{h+}$ . Now we have

$$ \begin{align*} \pi(\psi_i^{-1}(\hat x, y)) &= (\hat w^{-1} L(\hat x) F(\hat w))_h \\ &= (L(\hat x) \hat w^{-1} [\hat w, L(\hat x)^{-1}] F(\hat w))_h \\ &= (\hat w^{-1} [\hat w, L(\hat x)^{-1}] F(\hat w))_h + L(\hat x)_h \\ &= (\hat w^{-1} F(\hat w) [\hat w, L(\hat x)^{-1}])_h + L(\hat x)_h \\ &= (\hat w^{-1} F(\hat w))_h + [\hat w, L(\hat x)^{-1}]_h + L(\hat x)_h \\ &= -\alpha_i^\vee \otimes y_{f_j}^{q^{N/2} + 1} + L(\hat x)_h \\ &= -\alpha_j^\vee \otimes y_{f_j}^{q^{N/2} + 1} + \pi(\psi_i^{-1}(\hat x, 0)), \end{align*} $$

where the third (resp. the fifth) equality follows from Lemma 5.14 (resp. Lemma 5.13). The proof is finished.

Recall that $\mathcal {L}_{\chi _{h+}^h}$ is the multiplicative local system on $V = X_*(T) \otimes \overline {\mathbb {F}}_q$ corresponding to the character $\chi _{h+}^h: V^F \to \overline {\mathbb {Q}}_\ell ^\times $ .

Proposition 5.16. Let $\alpha \in \Phi \setminus \Phi _M$ and let $\kappa : \mathbb {G}_a \to V$ be the map given by $x \mapsto \alpha ^\vee \otimes x$ for $x \in \overline {\mathbb {F}}_q$ .

(1) $\kappa ^* \mathcal {L}_{\chi _{h+}^h}$ is nontrivial, and hence, $H_c^i(\mathbb {G}_a, \kappa ^* \mathcal {L}_{\chi _{h+}^h}) = 0$ for $i \in \mathbb {Z}$ ;

(2) if N is even and $F^{N/2}(\alpha ) = -\alpha $ , then

$$ \begin{align*} \dim H_c^i(\mathbb{G}_a, \tau^* \mathcal{L}_{\chi_{h+}^h}) = \begin{cases} q^{N/2} &\text{ if } i = 1; \\ 0, &\text{otherwise,} \end{cases} \end{align*} $$

where $\tau : \mathbb {G}_a \to V$ is given by $x \mapsto \alpha ^\vee \otimes x^{q^{N/2} + 1}$ . Moreover, in this case, $F^N$ acts on $H_c^1(\mathbb {G}_a, \tau ^* \mathcal {L}_{\chi _{h+}^h})$ via $-q^{N/2}$ .

Proof. Consider the composition of maps

$$\begin{align*}\theta: \mathbb{F}_{q^N} \overset \kappa \to V^{F^N} \overset {\mathrm{Nm}_N} \to V^F \overset {\chi_{h+}^h} \to \overline{\mathbb{Q}}_\ell^\times,\end{align*}$$

where $\mathrm {Nm}_N: V \to V$ is given by $v \mapsto v + \cdots + F^{N-1}(v)$ . As $\kappa $ is a homomorphism defined over $\mathbb {F}_{q^N}$ , we have $\kappa ^* \mathcal {L}_{\chi _{h+}^h} = \mathcal {L}_\theta $ by Lemma 5.8. Moreover, since $\alpha \in \Phi \setminus \Phi _M$ , $\theta $ is nontrivial by definition. Hence, $\mathcal {L}_\theta $ is nontrivial, and the statement (1) follows from [Reference Boyarchenko2, Lemma 9.4].

Assume that N is even and $F^{N/2}(\alpha ) = -\alpha $ . Then for $x \in \mathbb {F}_{q^{N/2}}$ , we have

$$ \begin{align*} \mathrm{Nm}_N(\alpha^\vee \otimes x) &= \sum_{i=0}^{N-1} F^i(\alpha^\vee) \otimes x^{q^i} \\ &= \sum_{i=0}^{N/2 - 1}(F^i(\alpha^\vee)\otimes x^{q^i} + F^{N/2 + i}(\alpha) \otimes x^{q^{N/2 + i}}) \\ &= \sum_{i=0}^{N/2 - 1}(F^i(\alpha^\vee)\otimes x^{q^i} + F^i(-\alpha^\vee) \otimes x^{q^i}) \\ &=0.\end{align*} $$

Hence, the (nontrivial) character $\theta $ of $\mathbb {F}_{q^N}$ restricts to a trivial character of $\mathbb {F}_{q^{N/2}}$ . Now the statement (2) follows from [Reference Boyarchenko and Weinstein5, Proposition 6.6.1].

Let Z be a locally closed subvariety of $\mathbb {G}_h^+$ with the natural embedding map $i_Z: Z \hookrightarrow \mathbb {G}_h^+$ . For a local system $\mathcal {F}$ on $\mathbb {G}_h^+$ , we write $H_c^j(Z, \mathcal {F}) = H_c^j(Z, i_Z^*\mathcal {F})$ for simplicity. We set $\pi = \pi _h^{\widetilde \Phi ^+:\widetilde \Phi ^h}: \mathbb {G}_h^+ = \mathbb {G}_r^+/\mathbb {G}_r^h \to \mathbb {A}_h \cong V$ .

Proposition 5.17. The following statements hold:

(1) $H_c^j(Y^{A_i}_h, \pi ^* \mathcal {L}_{\chi _{h+}^h}) \cong H_c^j(Y^{A_{i+1}}_h, \pi ^* \mathcal {L}_{\chi _{h+}^h})^{\oplus q^N}$ for $1 \leqslant i \leqslant m-1$ ;

(2) $H_c^j(Y^{A_m}_h, \pi ^* \mathcal {L}_{\chi _{h+}^h}) \cong H_c^{j-n-m+1}(Y^M_h, \pi _M^* \mathcal {L}_{\chi _{h+}^h})^{\oplus q^{(n-m+1)N/2}} ((-q^{N/2})^{n+m-1})$ .

Here, $Y^M_h = Y_h \cap \mathbb {M}_h^+$ , and $\pi _M$ is the restriction of $\pi $ to $\mathbb {M}_h^+$ .

Proof. By Proposition 5.15, we have an isomorphism

$$\begin{align*}\psi_i: Y^{A_i}_h \cong Y^{A_i}_{B_i} \times \mathbb{A}_{D_i^\flat}.\end{align*}$$

Let $p: Y^{A_i}_h \to Y^{A_i}_{B_i}$ be the natural projection. Set $\mathcal {L} = \mathcal {L}_{\chi _{h+}^h}$ .

Assume $1 \leqslant i \leqslant m-1$ . Let $Y^i = \{\hat x \in Y^{A_i}_h; x_{f_i}^{q^N} - x_{f_i} = 0\}$ . Then $\psi _i$ restricts to an isomorphism

$$\begin{align*}Y^i \cong Y_i \times \mathbb{A}_{D_i^\flat},\end{align*}$$

where $Y_i = \{\hat x \in Y^{A_i}_{B_i}; x_{f_i}^{q^N} - x_{f_i} = 0\}$ . In view of Proposition 5.15 (1), the restriction of $\pi $ to $Y^{A_i}_{B_i} \times \mathbb {A}_{D_i^\flat }$ is given by

$$\begin{align*}\pi(\psi_i^{-1}(\hat x, y)) = \pi_i(\psi_i^{-1}(\hat x, y)) = \eta(\hat x, y) + \pi_0(\hat x),\end{align*}$$

where $\eta (\hat x, y) = \alpha _{f_i}^\vee \otimes (x_{f_i}^{q^N}- x_{f_i}) y_{f_i}^{q^{n_i}}$ with $1 \leqslant n_i \leqslant N-1$ such that $F^{n_i}(f_i^\flat ) = h - f_i$ , and $\pi _0$ is the restriction of $\pi $ to $Y^{A_i}_{B_i} \times \{0\} \subseteq Y^{A_i}_{B_i} \times \mathbb {A}_{D_i^\flat }$ . As $\mathcal {L}$ is a multiplicative local system, we have $\pi ^* \mathcal {L} \cong \eta ^* \mathcal {L} \otimes p^* \pi _0^* \mathcal {L}$ . Hence, by the projection formula,

$$\begin{align*}p_! \pi^* \mathcal{L} \cong p_! \eta^* \mathcal{L} \otimes \pi_0^* \mathcal{L}.\end{align*}$$

For $\hat x \in Y^{A_i}_h$ , we define $\eta _{\hat x}: \mathbb {A}_{D_i^\flat } \to V$ to be the homomorphism given by $\eta _{\hat x}(y) = \eta (\hat x, y)$ . As $\alpha _{f_i} \in \Phi \setminus \Phi _M$ , it follows by Proposition 5.16 that $\eta _{\hat x}^* \mathcal {L}$ is a trivial multiplicative local system if and only if $x_{f_i}^{q^N} - x_{f_i} =0$ ; that is, $\hat x \in Y_i$ . Thus, $p_! \pi ^* \mathcal {L}$ is supported on $Y_i \times \mathbb {A}_{D_i^\flat } \cong Y^i$ , and hence, $p_! \pi ^* \mathcal {L} \cong p_! (\pi |_{Y^i})^* \mathcal {L}$ . Noticing that

$$\begin{align*}Y^i = \sqcup_{g \in (\mathbb{G}_h^{A_i})^F / (\mathbb{G}_h^{A_{i+1}})^F} g Y^{A_{i+1}}_h\end{align*}$$

and that $\#((\mathbb {G}_h^{A_i})^F / (\mathbb {G}_h^{A_{i+1}})^F) = \#(\mathbb {G}_h^{A_i} / \mathbb {G}_h^{A_{i+1}})^F = q^N$ , we have

$$\begin{align*}H_c^j(Y^{A_i}_h, \pi^* \mathcal{L}) \cong H_c^j(Y^i, \pi^* \mathcal{L}) \cong H_c^j(Y^{A_{i+1}}_h, \pi^* \mathcal{L})^{\oplus q^N},\end{align*}$$

and the first statement is proved.

By Proposition 5.15 (2), for $(\hat x, y) \in Y^{A_m}_{B_m} \times \mathbb {A}_{D_m^\flat } =Y^M_h \times \mathbb {A}_{D_m^\flat }$ , we have

$$\begin{align*}\pi(\hat x, y) = \tau(y) + \pi_M(\hat x),\end{align*}$$

where $\tau (y) = \sum _{j=m}^n \alpha _{f_j}^\vee \otimes y_{f_j}^{q^{N/2}+1}$ . Thus, $\pi ^* \mathcal {L} \cong \pi _M^* \mathcal {L} \boxtimes \tau ^* \mathcal {L}$ . By Künneth formula, we have

$$ \begin{align*} H_c^j(Y^{A_m}_h, \pi^* \mathcal{L}) &\cong \oplus_s H_c^s(\mathbb{A}_{D_m^\flat}, \tau^* \mathcal{L}) \otimes H_c^{j-s}(Y^M_h, \pi_M^* \mathcal{L}) \\ &\cong \otimes_{i=m}^n H_c^1(\mathbb{A}_{f_i}, \tau_i^* \mathcal{L}) \otimes \otimes_{i=1}^{m-1} H_c^2(\mathbb{A}_{f_i^\flat}, \overline{\mathbb{Q}}_\ell) \otimes H_c^{j-n-m+1}(Y^M_h, \pi_M^* \mathcal{L}) \\ &\cong H_c^{j-n-m+1}(Y^M_h, \pi_M^* \mathcal{L})^{q^{(n-m+1)N/2}}((-q^{N/2})^{m+n-1}), \end{align*} $$

where $\tau _i : \mathbb {G}_a \cong \mathbb {A}_{f_i} \to V$ is given by $x \mapsto \alpha _{f_i}^\vee \otimes x^{q^{N/2}+1}$ , and the last isomorphism follows from Proposition 5.16 (2). This finishes the proof of the second statement.

5.7 Proof of Theorem 5.5

Let $G'$ , $T'$ , $Y_f'$ be as in §5.5. Let $\chi '$ be the restriction of $\chi $ to $(\mathbb {T}^{\prime +}_f)^F$ . By Lemma 5.11, we have

$$\begin{align*}H_c^j(Y_f, \overline{\mathbb{Q}}_\ell)[\chi] \cong (\mathrm{ind}_{(\mathbb{T}^{\prime +}_f)^F}^{(\mathbb{T}_f^+)^F} (H_c^j(Y_f', \overline{\mathbb{Q}}_\ell)[\chi']))[\chi].\end{align*}$$

So it suffices to prove the theorem for semisimple reductive groups $G = G'$ .

We argue by induction on $f \in \widetilde \Delta ^+$ and $\# \Phi $ . Indeed, if $f = \min \widetilde \Delta ^+$ , then $(\mathbb {T}_f^+)^F = Y_f = \{1\}$ and the statement is trivial. However, if $\Phi $ is empty – that is, $G = T$ – then $Y_f = (\mathbb {T}_f^+)^F$ is a finite set and the statement is also true. Now we assume the theorem holds for all reductive groups L such that $\# \Phi _L < \# \Phi _G$ , and for all $Y_{f'}$ with $f' \leqslant f \in \widetilde \Delta ^+$ .

If $f \in \Delta _{\mathrm {aff}}^+$ , by Corollary 5.9 (1), we have a $(\mathbb {T}_f^+)^F$ -equivariant isomorphism

$$\begin{align*}H_c^i(Y_{f+}, \overline{\mathbb{Q}}_\ell) = H_c^{i-2}(Y_f, \overline{\mathbb{Q}}_\ell)(-q^N).\end{align*}$$

Then the statement follows by induction hypothesis.

Now we assume $f = h \in \mathbb {Z}_{\geqslant 1}$ . Let notation be as in §5.6. By Corollary 5.9 (2),

$$\begin{align*}H_c^i(Y_{h+}, \overline{\mathbb{Q}}_\ell)[\chi] \cong H_c^i(Y_h, \pi^* \mathcal{L}_{\chi_{h+}^h})[\chi].\end{align*}$$

If $\chi _{h+}^h$ is trivial, then $\mathcal {L}_{\chi _{h+}^h} = \overline {\mathbb {Q}}_\ell $ and $H_c^i(Y_{h+}, \overline {\mathbb {Q}}_\ell )[\chi ] \cong H_c^i(Y_h, \overline {\mathbb {Q}}_\ell )[\chi ]$ . Hence, the statement also follows by induction hypothesis. Assume $\chi _{h+}^h$ is nontrivial and let notation be as in §5.6. By Condition 2.1 and Lemma 2.2, we have $M = M_\chi \neq G$ . By Proposition 5.17 and Corollary 5.9 (2), we have

$$ \begin{align*} &\quad\ H_c^i(Y_h, \pi^* \mathcal{L}_{\chi_{h+}^h}) [\chi] = H_c^i(Y^{A_1}_h, \pi^* \mathcal{L}_{\chi_{h+}^h})[\chi] \\ &\cong (H_c^{i-m-n+1}(Y^M_h, \pi_M^* \mathcal{L}_{\chi_{h+}^h})[\chi])^{\oplus q^{(m+n-1)N/2}}((-q)^{(m+n-1)N/2}) \\ &\cong (H_c^{i-m-n+1}(Y_{h+}^M, \overline{\mathbb{Q}}_\ell) [\chi])^{\oplus q^{(m+n-1)N/2}}((-q)^{(m+n-1)N/2}), \end{align*} $$

so the statement follows by induction hypothesis. The proof is finished.

5.8 Computation of cohomological degree

Let $\chi $ be a smooth character $\mathcal {T}^+(\mathcal {O}_k)$ , which factors through $(\mathbb {T}_r^+)^F$ . We have the Howe factorization of an arbitrary lift of $\chi $ to a smooth character of $T(k)$ from [Reference Kaletha21, §3.6]. We use notation from loc. cit. In particular, we have the integers $(r \geq ) r_d \geq r_{d-1}> r_{d-2} > \dots > r_0 > 0$ at which the breaks happen and the increasing subsets $R_i:= R_{r_i} \subseteq \Phi $ (which are the roots systems of the twisted Levi subgroups $M_\chi $ appearing in §5.7). Moreover, $r_{-1} = 0$ , $R_d = \Phi $ by definition. Let $R_i^{\mathrm {red}} = R_i \cap \Phi ^{\mathrm {red}}$ , where $\Phi ^{\mathrm {red}}$ is as in §4.3.

Proposition 5.18. We have

$$\begin{align*}N s_{\chi,r} = 2r\cdot\#\Phi - \#\Phi^{\mathrm{red}} - \#R_0^{\mathrm{red}} - \sum_{i=0}^{d-1} r_i(\#R_{i+1} - \#R_i). \end{align*}$$

Proof. We can argue by induction on $\# \Phi $ (or on the number of jumps d). If $\Phi = \varnothing $ , the statement is trivial. Suppose it is true for all reductive groups L with $\#\Phi _L < \#\Phi $ . Then in view of §5.7 (where we can assume that $\chi $ is trivial over $\mathbb {T}_r^{h+1}$ with $h = r_{d-1}$ ), we have

$$\begin{align*}s_{\chi,r} = 2(r-r_{d-1}) \cdot \#\Delta +(m+n-1) + s_{\chi, r_{d-1}}^M,\end{align*}$$

where $s_{\chi ,r_{d-1}}^M$ is the unique integer i such that $H_c^i(Y_{r_{d-1}}^M, \overline {\mathbb {Q}}_\ell )[\chi ] \neq 0$ . Now,

$$ \begin{align*} m+n-1 &= \#D \\ &= \#\{ f \in \Delta_{\mathrm{aff}} \colon f(\textbf{x})> 0, f < r_{d-1} \} \cap (\widetilde R_d {\,\setminus\,} \widetilde R_{d-1})) \\ &= \frac{1}{N} \left( r_{d-1} (\#R_d - \#R_{d-1}) - (\#R_d^{\mathrm{red}} - \# R_{d-1}^{\mathrm{red}}) \right), \end{align*} $$

where $ \widetilde R_{d-1} \subseteq \widetilde \Phi $ is the preimage of $R_i$ under the natural projection $\widetilde \Phi \to \Phi \sqcup \{0\}$ . Note that $N \cdot \#\Delta = \#\Phi =\#R_d$ . The statement now follows by induction hypothesis.

This generalizes the formula from [Reference Chan and Ivanov8, Theorem 6.1.1]

Corollary 5.19. For the integer $s_\chi $ from Theorem 1.1, we have

$$\begin{align*}Ns_\chi = -\#\Phi^{\mathrm{red}} + \#R_0^{\mathrm{red}} + \sum_{i=0}^{d-1} r_i(\#R_{i+1} - \#R_i). \end{align*}$$

Proof. By Lemma 4.4, $2N\dim Y_r = 2N(r \cdot \#\Delta - \#\Delta ^{\mathrm {red}}) = 2(r \cdot \#\Phi - \#\Phi ^{\mathrm {red}})$ . As $Ns_\chi = 2N\dim Y_r - Ns_{\chi ,r}$ , the claim follows.

Note that when $\chi $ is sufficiently generic, this, combined with Corollary 1.3, gives a formula for the formal degree of the corresponding supercuspidal representation. Moreover, note that the essential parts of the formulas of Corollary 5.19 and of [Reference Schwein26, Theorem A] agree.

We now explicate the formula from Corollary 5.19 in two special cases.

Example 5.20. Assume that $\chi $ is toral; that is, $d = 1$ and the Howe filtration of $\chi $ has exactly one jump happening at some integer $r_0$ satisfying $0 < r_0 \leq r_1 := \mathrm {depth}(\chi )$ . Then $R_1 = \Phi $ and $R_0 = R_{-1} = \varnothing $ , and the formula from Corollary 5.19 simplifies to

$$\begin{align*}Ns_{\chi} = (r_0-1)\#\Phi. \end{align*}$$

Example 5.21. Assume that $G = \mathrm {GL}_n$ and $\mathcal {G}$ is hyperspecial, so that $\Phi = \Phi ^{\mathrm {red}}$ and $\#\Phi =n(n-1)$ . We can take $T \subseteq G$ to be the diagonal torus so that $\Phi = \{\alpha _{i,j}\}_{1\leq i\neq j \leq n}$ with $\alpha _{i,j}$ corresponding to $(i,j)$ th matrix entry. Moreover, we may take $c = (1,2,\dots ,n)$ ; then the (root systems of the) twisted Levi subgroups containing T are in bijection with divisors $0<k|n$ , with k corresponding to $R(k) = \{\alpha _{i,j}\}_{i-j \equiv 0\ \mod \frac {n}{k}}$ , so that $R(k) \subseteq R(k')$ if and only if $k|k'$ and $\#R(k) = k(k-1)$ (in particular, $R(1) = \varnothing $ corresponds to T itself). Given the integers $r_i$ ( $-1 \leq i \leq d$ ) as above, along with a corresponding sequence of divisors $1 = k_{-1}|k_0 | \dots |k_{d-1}|k_d=n$ of n, all of which are mutually different (except possibly $k_0 = 1$ ), formula from Corollary 5.19 then becomes

$$\begin{align*}Ns_\chi = -n(n-1) + k_0(k_0-1) + \sum_{i=0}^{d-1}r_i(k_{i+1}(k_{i+1}-1) - k_i(k_i-1)). \end{align*}$$

6 Traces

We combine Theorem 5.5 with [Reference Boyarchenko3, Lemma 2.12] to express the traces of all $g \in \mathcal {G}(\mathcal {O}_k)$ on $H_{s_\chi }(Y,\overline {\mathbb {Q}}_\ell )[\chi ]$ in terms of the geometry of $Y_h$ . In particular, we determine the dimension of $H_{s_\chi }(Y,\overline {\mathbb {Q}}_\ell )[\chi ]$ in terms of the non-vanishing degree $s_{\chi }$ .

Proposition 6.1. Let $\chi \colon \mathcal {T}^+(\mathcal {O}_k) \rightarrow \overline {\mathbb {Q}}_\ell ^\times $ be a smooth character which factors through $(\mathbb {T}_r^+)^F$ . Let $g \in \mathcal {G}^+(\mathcal {O}_k)$ with image $\bar g \in (\mathbb {G}_r^+)^F$ . Then

$$\begin{align*}\mathrm{tr}(\bar g, H_c^{s_{\chi,r}}(Y_r,\overline{\mathbb{Q}}_\ell)[\chi]) = \frac{1}{\#(\mathbb{T}_r^+)^F \cdot q^{s_{\chi,r} N/2}} \sum_{t \in (\mathbb{T}_r^+)^F} \chi(t)\cdot \#S_{g,t}, \end{align*}$$

where $S_{g,t} = \{x \in Y_r(\overline {\mathbb {F}}_q) \colon gF^N(x) = x t\}$ . For $g=1$ , this simplifies to

$$\begin{align*}\dim_{\overline{\mathbb{Q}}_\ell} H_c^{s_{\chi,r}}(Y_r,\overline{\mathbb{Q}}_\ell)[\chi] = \frac{\#(\mathbb{G}_r^+)^F}{\#(\mathbb{T}_r^+)^F \cdot q^{s_{\chi,r} N/2}}. \end{align*}$$

Proof. The first statement follows from Theorem 5.5 by [Reference Boyarchenko3, Lemma 2.12]. For the second statement, we have to compute the trace for $g=1$ . Therefore, let $x \in S_{1,t}$ for some $t \in (\mathbb {T}_h^+)^F$ and put $u = x^{-1}F(x)$ . Then $x \in S_{1,t}$ implies $t = x^{-1}F^N(x) = \prod _{i=0}^{N-1} F^i(u)$ . We claim that this implies $t=u=1$ . Let $A \subseteq \widetilde \Phi ^+$ be an F-stable and closed subset. Suppose that we have already shown that $t,u \in \mathbb {G}_r^A$ . Let $f \in A$ be such that $f(\textbf {x})$ is minimal among all roots in A. Then $A {\,\setminus \,} \mathcal {O}_f \subseteq A$ is F-stable and closed, and $A + A \subseteq A {\,\setminus \,} \mathcal {O}_f$ , so that $\mathbb {G}_h^{A{\,\setminus \,} \mathcal {O}_f} \subseteq \mathbb {G}_h^A$ is normal with abelian quotient. By induction on A, it suffices to show that $t,u \in \mathbb {G}_h^{A {\,\setminus \,} \mathcal {O}_f}$ . Let $\bar t,\bar u \in \mathbb {G}_h^A / \mathbb {G}_h^{A {\,\setminus \,} \mathcal {O}_f}$ denote the images of $t,u$ . If $f \in \mathbb {Z}_{>0}$ , then $\bar u = 1$ , and hence also $\bar t = 1$ , so that we are done. If $f \not \in \mathbb {Z}_{>0}$ , then $t = 1$ and $\mathbb {G}_h^A / \mathbb {G}_h^{A {\,\setminus \,} \mathcal {O}_f} \cong \prod _{i=0}^{N-1} \mathbb {G}_a$ , with F-action given by $F((x_i)_{i=0}^{N-1}) = (x_{i-1}^q)_{i=0}^{N-1}$ (the ith copy of $\mathbb {G}_a$ corresponds to $F^i(f)$ ). Now, as $u \in \overline {\mathbb {U}}_h \cap F\mathbb {U}_h^-$ by assumption, $\bar u$ corresponds under this isomorphism to an element of the form $(a,0, \dots , 0)$ with $a\in \mathbb {G}_a$ , and the equation $\prod _{i=0}^{N-1} F^i(\bar u) = 1$ in $\mathbb {G}_h^A / \mathbb {G}_h^{A {\,\setminus \,} \mathcal {O}_f}$ thus corresponds to $(a,a^q, \dots , a^{q^{N-1}}) = 0$ . Thus, $a = 0$ (i.e., $\bar u = 1$ ), and our original claim follows by induction on A. The claim immediately implies $S_{1,t} = \varnothing $ unless $t=1$ and $S_{1,1} = (\mathbb {G}_h^+)^F$ which proves the proposition.

Proof of Corollary 1.3.

Let $r \in \mathbb {Z}_{\geqslant 1}$ such that $\chi $ factors through $\mathbb {T}_r^+$ . It follows from §2.7 that $s_{\chi ,r} = 2\dim (Y_r) - s_\chi = 2\dim (\overline {\mathbb {U}}_r^+ \cap F\mathbb {U}_r^+) - s_\chi $ . Note that

$$\begin{align*}N \dim (\overline{\mathbb{U}}_r^+ \cap F\mathbb{U}_r^+) = \# (\widetilde \Phi_r^+ \cap \widetilde \Phi_{\mathrm{aff}}) = \dim \mathbb{G}_r^+ - \dim \mathbb{T}_r^+.\end{align*}$$

Thus,

$$ \begin{align*} q^{s_{\chi,r}N/2} &= q^{N\dim (\overline{\mathbb{U}}_r^+ \cap F\mathbb{U}_r^+) - \frac{s_\chi N}{2}} = q^{\dim (\mathbb{G}_r^+ / \mathbb{T}_r^+) - \frac{s_\chi N}{2}} \\ &= \frac{\#(\mathbb{G}_r^+)^F}{\#(\mathbb{T}_r^+)^F} \cdot q^{-s_\chi N/2}. \end{align*} $$

Inserting this into the second formula of Proposition 6.1 gives the result.

Corollary 6.2. Assume that p satisfies Condition 2.1. The varieties $Y_f$ for $f \in \widetilde \Phi _r^+$ are maximal. In particular, the varieties $X_r^{(\mathbb {T}_{0+})}$ for $r \in \mathbb {Z}_{\geqslant 0}$ are maximal.

Proof. By definition, we need to show that either $H_c^s(Y_f, \overline {\mathbb {Q}}_\ell ) = 0$ or $F^N$ acts on $H_c^s(Y_f, \overline {\mathbb {Q}}_\ell )$ by the scalar $(-1)^s q^{sN/2}$ with $sN$ even. By Proposition 5.3, we can replace f with $h = \min \{n \in \mathbb {Z}: n \geqslant f\}$ .

Assume that $H_c^s(Y_h, \overline {\mathbb {Q}}_\ell ) \neq 0$ . Then there exists a character $\chi $ of $\mathbb {T}_h^+$ such that $H_c^s(Y_h, \overline {\mathbb {Q}}_\ell )[\chi ] \neq 0$ . By Theorem 5.5, $s = s_{h, \chi }$ and $F^N$ acts on $H_c^s(Y_h, \overline {\mathbb {Q}}_\ell )[\chi ]$ by the scalar $(-1)^s q^{sN/2}$ . It remains to show $sN$ is even. In view of Proposition 5.18, this question is combinatorial, and we may assume that q is a suitable prime number. Then it follows from Corollary 1.3 that $s_\chi N$ is even. As $s = 2\dim Y_h - s_\chi $ , we deduce that $s N$ is even, as desired.

7 Irreducibility

Until the end of this article, we assume that $(T, U)$ is a Coxeter pair.

Recall the minimal Drinfeld stratum $X^{(\mathbb {T}_{0+})}$ of $X\subseteq \mathbb {G}$ from §4.3. We have its subscheme Y and the slightly bigger subscheme

$$\begin{align*}Z = X^{(\mathbb{T}_{0+})} \cap \mathbb{T} \mathbb{G}^+ = \{g \in \mathbb{T} \mathbb{G}^+ \colon g^{-1} F(g) \in \overline{\mathbb{U}} \cap F\mathbb{U} \}. \end{align*}$$

We have the corresponding approximations $Y_r \subseteq \mathbb {G}_r^+$ , and $Z_r \subseteq \mathbb {T}_r \mathbb {G}_r^+$ ; $Y_r$ is equipped with an $(\mathbb {G}_r^+)^F \times (\mathbb {T}_r^+)^F$ -action and $Z_r$ is equipped with an $(\mathbb {T}_r\mathbb {G}_r^+)^F \times \mathbb {T}_r^F$ -action.

In Theorem 5.5, we have seen that for any $\chi \colon (\mathbb {T}_r^+)^F \rightarrow \overline {\mathbb {Q}}_\ell ^\times $ , $H^\ast _c(Y_r)[\chi ]$ is concentrated in one degree. By Lemma 4.3, the same holds also for $Z_r$ for any character $\chi \colon \mathbb {T}_r^F \rightarrow \overline {\mathbb {Q}}_\ell ^\times $ . Now we prove that these weight spaces are irreducible as $\mathbb {G}_r^F$ (resp. $(\mathbb {G}_r^+)^F$ -)representations and pairwise distinct.

Theorem 7.1. For any $\chi ,\chi ' \colon \mathbb {T}_r^F \rightarrow \overline {\mathbb {Q}}_\ell ^\times $ , we have

$$\begin{align*}\left\langle H^\ast_c(Z_r)[\chi],H^\ast_c(Z_r)[\chi'] \right\rangle_{(\mathbb{T}_r \mathbb{G}_r^+)^F} = \begin{cases} 1 & \text{if } \chi = \chi', \\ 0 & \text{otherwise.} \end{cases} \end{align*}$$

The same holds for $Y_r$ , when $(\mathbb {T}_r)^F$ , $(\mathbb {T}_r \mathbb {G}_r^+)^F$ are replaced by $(\mathbb {T}_r^+)^F$ , $(\mathbb {G}_r^+)^F$ .

Proof. Let

$$\begin{align*}\Sigma = \{(y, x, x') \in \mathbb{T}_r \mathbb{G}_r^+ \times (\overline{\mathbb{U}}_r^+ \cap F\mathbb{U}_r^+) \times (\overline{\mathbb{U}}_r^+ \cap F\mathbb{U}_r^+); y^{-1} x F(y) = x'\},\end{align*}$$

equipped with $\mathbb {T}_r^F\times \mathbb {T}_r^F$ -action by $(t,t') \colon (y,x,x') \mapsto (tyt^{\prime -1},txt^{-1},t'x't^{\prime -1})$ . As in [Reference Deligne and Lusztig12, §6.6], we have $\Sigma \cong (\mathbb {T}_r \mathbb {G}_r^+)^F \backslash (Z_r \times Z_r)$ . It thus suffices to show that $H^\ast _c(\Sigma ) \cong H_c^*(\mathbb {T}_r^F)$ .

By Iwahori decomposition, we have $y = \tau y_+ y_-$ with $y_+ \in \mathbb {U}_r^+$ , $\tau \in \mathbb {T}_r$ and $y_- \in \overline {\mathbb {U}}_r^+$ . Then the equality $y^{-1} x F(y) = x'$ is equivalent to

(a) $$ \begin{align} y_+^{-1} \tau^{-1} x F(\tau) F(y_+) F(y_-) = y_- x'. \end{align} $$

By Proposition 3.1 (2), there is a unique pair $(u_1, u_2) \in (\mathbb {U}_r^+ \cap F^{-1}\overline {\mathbb {U}}_r^+) \times \mathbb {U}_r^+$ such that

(*) $$ \begin{align} y_+^{-1} \tau^{-1} x F(\tau) F(y_+) = u_2 \tau^{-1} F(\tau) F(u_1), \end{align} $$

and moreover, the correspondence $(\tau , x, y_+) \mapsto (\tau , u_1, u_2)$ gives an isomorphism

$$\begin{align*}\mathbb{T}_r \times (\overline{\mathbb{U}}_r^+ \cap \mathbb{U}_r^+) \times \mathbb{U}_r^+ \cong \mathbb{T}_r \times (\mathbb{U}_r^+ \cap F^{-1}\overline{\mathbb{U}}_r^+) \times \mathbb{U}_r^+.\end{align*}$$

Now the equality (a) becomes

(b) $$ \begin{align} u_2 \tau^{-1} F(\tau) F(u_1) F(y_-) = y_- x'. \end{align} $$

Write $y_- = y_1 y_2$ with $y_1 \in \overline {\mathbb {U}}_r^+ \cap F^{-1}(\mathbb {U}_r^+)$ and $y_2 \in \overline {\mathbb {U}}_r^+ \cap F^{-1}(\overline {\mathbb {U}}_r^+)$ . By Propositon 3.1 (1), the map $(x', y_2) \mapsto u_- := y_2 x' F(y_2)^{-1}$ gives an isomorphism $(F\mathbb {U}_r^+ \cap \overline {\mathbb {U}}_r^+) \times (\overline {\mathbb {U}}_r^+ \cap F^{-1}\overline {\mathbb {U}}_r^+) \cong \overline {\mathbb {U}}_r^+$ . Thus, the equality (b) becomes

(c) $$ \begin{align} u_2 \tau^{-1} F(\tau) F(u_1 y_1) = y_1 u_-. \end{align} $$

Write $u_1 y_1 = z_1 z_0 z_2$ with $z_1 \in F^{-1}(\mathbb {U}_r^+)$ , $z_0 \in \mathbb {T}_r$ and $z_2 \in F^{-1}\overline {\mathbb {U}}_r^+$ . Then the equality (c) becomes

(d) $$ \begin{align} u_2 {}^{\tau^{-1} F(\tau)} F(z_1) \tau^{-1} F(\tau) F(z_0) F(z_2) = y_1 u_-. \end{align} $$

It follows from (d) that $u_2 = {}^{\tau ^{-1} F(\tau )} F(z_1)^{-1}$ , $\tau ^{-1} F(\tau ) = F(z_0)^{-1}$ and $u_- = y_1^{-1} F(z_2)$ . Thus, we deduce that

(e) $$ \begin{align} \Sigma \cong \{(\tau, u_1, y_1) \in \mathbb{T}_r \times (\mathbb{U}_r^+ \cap F^{-1}\overline{\mathbb{U}}^+) \times (\overline{\mathbb{U}}_r^+ \cap F^{-1}\mathbb{U}_r^+); \tau F(\tau)^{-1} = \mathrm{pr}_0(F(u_1 y_1))\}, \end{align} $$

where $\mathrm {pr}_0: \mathbb {T}_r\mathbb {G}_r^+ \cong \mathbb {U}_r^+ \times \mathbb {T}_r \times \overline {\mathbb {U}}_r^+ \to \mathbb {T}_r$ is the natural projection.

Note that $(\zeta , \xi ) \in \mathbb {T}_r^F \times \mathbb {T}_r^F$ acts on $\Sigma $ by $(y, x, x')\mapsto (\zeta y \xi ^{-1}, \zeta x \zeta ^{-1}, \xi x' \xi ^{-1})$ . Then $(\zeta , \xi )$ sends $(\tau , x, y_+, y_-)$ to $(\tau \zeta \xi ^{-1}, \zeta x \zeta ^{-1}, \xi y_+ \xi ^{-1}, \xi y_- \xi ^{-1})$ . Using the relation (*), we see that $(\zeta , \xi )$ sends $(u_1, u_2)$ to $(\xi u_1 \xi ^{-1}, \xi u_2 \xi ^{-1})$ . Therefore, in view of (e), $(\zeta , \xi )$ acts on $\Sigma $ by sending $(\tau , u_1, y_1)$ to $(\tau \zeta \xi ^{-1}, \xi u_1 \xi ^{-1}, \xi y_1 \xi ^{-1})$ .

Let $\eta \in \mathbb {T}_r$ . Consider the action of $\eta $ on $\Sigma $ by sending $(\tau , u_1, y_1)$ to $(\tau , \eta u_1 \eta ^{-1}, \eta y_1 \eta ^{-1})$ . Then the action of $\mathbb {T}_r$ commutes with the action of $\mathbb {T}_r^F \times \mathbb {T}_r^F$ . Thus, we have an $\mathbb {T}_r^F \times \mathbb {T}_r^F$ -equivariant isomorphism

$$\begin{align*}H_c^*(\Sigma) \cong H_c^*(\Sigma^{\mathbb{T}_{r, \mathrm{red}}}) \cong H_c^*(\mathbb{T}_r^F),\end{align*}$$

where $\mathbb {T}_{r, \mathrm {red}}$ denotes the reductive part of $\mathbb {T}_r$ . Now the statement follows. The proof for $Y_r$ is the same.

8 Relation to the orbit method

Let $r \in \mathbb {Z}_{\geq 1} \cup \{\infty \}$ . We have the groups $\mathbb {G}_r^+$ and $\mathbb {T}_r^+$ and the variety $Y_r$ with $(\mathbb {G}_r^+)^F \times (\mathbb {T}_r^+)^F$ -action (where we put $Y_\infty =Y$ , and similarly for $\mathbb {T}_\infty ^+$ ). Theorems 5.5 and 7.1 show that $H_c^{s_{\chi ,r}}(Y_r,\overline {\mathbb {Q}}_\ell )[\chi ]$ is an irreducible $(\mathbb {G}_r^+)^F$ -representation. However, if either $r<p$ , or $r = \infty $ and $(\mathbb {G}_r^+)^F$ is uniform (see below), Kirillov’s orbit method attaches irreducible $(\mathbb {G}_r^+)^F$ -representations to adjoint $(\mathbb {G}_r^+)^F$ -orbits in the dual of the Lie algebra of $(\mathbb {G}_r^+)^F$ . We state a conjecture about the relation between these two constructions and verify it in a nontrivial case.

8.1 Review of the orbit method

The orbit method was originally developed by Kirillov [Reference Kirillov22] and extended later to various related setups. We briefly recall it in the two setups relevant for our article. We refer to [Reference Boyarchenko and Sabitova4] (in particular, §2 therein), [Reference Boyarchenko and Drinfeld1, §2] and [Reference Dixon, du Satoy, Mann and Segal11] and references therein for more detailed discussions.

Assume that $p>2$ .Footnote 3 For the first setup, recall that a uniform Lie algebra is a (topological) Lie algebra $\mathfrak {g}$ over $\mathbb {Z}_p$ , which is free of finite rank as a $\mathbb {Z}_p$ -module and satisfies $[\mathfrak {g},\mathfrak {g}] \subseteq p\mathfrak {g}$ . Following Lazard, there is a pro-p group $\Gamma = \exp \mathfrak {g}$ attached to $\mathfrak {g}$ , whose underlying topological space is $\mathfrak {g}$ and on which the group law is defined (via $\exp $ and $\log $ ) by the Campbell–Hausdorff series. For $\Gamma = \exp \mathfrak {g}$ , one has mutually inverse homeomorphisms $\exp \colon \Gamma \rightarrow \mathfrak {g}$ and $\log \colon \mathfrak {g} \rightarrow \Gamma $ . Set up appropriately, the functor $\mathfrak {g} \mapsto \exp \mathfrak {g}$ even defines an isomorphism of categories. We denote the inverse functor by $\Gamma \mapsto \log \Gamma $ . A profinite group $\Gamma $ is called uniform (short for uniformly powerful) if there is a uniform Lie-algebra $\mathfrak {g}$ with $\Gamma \cong \exp \mathfrak {g}$ . There is a similar isomorphism of categories between finite p-groups $\Gamma $ of nilpotence class $<p$ and finite nilpotent Lie rings $\mathfrak {g}$ of p-power order and nilpotence class $<p$ . We use the same notation as in the uniform pro-p case.

For the moment, let $\Gamma $ be either

  1. (i) a uniform pro-p group, or

  2. (ii) a finite p-group of nilpotence class $<p$ .

Let $\mathfrak {g} = \log \Gamma $ denote the corresponding uniform Lie $\mathbb {Z}_p$ -algebra resp. finite Lie ring. Let $\widehat \Gamma $ denote the set of isomorphism classes of smooth irreducible $\overline {\mathbb {Q}}_\ell $ -representations of $\Gamma $ . Note that there is an adjoint action of $\Gamma $ on $\mathfrak {g}$ . More precisely, for any $g \in \Gamma $ , we have the automorphism ${\mathrm {Ad}}\, g \colon \mathfrak {g} \rightarrow \mathfrak {g}$ given by $x \mapsto \log (g\exp (x)g^{-1})$ . Let

$$\begin{align*}\mathfrak{g}^\ast = \operatorname{\mathrm{Hom}}_{\mathrm{cont}}(\mathfrak{g},\overline{\mathbb{Q}}_\ell^\times) \end{align*}$$

be the dual of $\mathfrak {g}$ . The adjoint action of $\Gamma $ on $\mathfrak {g}$ induces an action of $\Gamma $ on $\mathfrak {g}^\ast $ . Kirillov’s orbit method, in the present setup established in [Reference Boyarchenko and Sabitova4], describes a natural bijection between $\widehat \Gamma $ and the set of $\Gamma $ -orbits in $\mathfrak {g}^\ast $ .

Theorem 8.1 (Theorem 2.6 in [Reference Boyarchenko and Sabitova4]).

Assume $p\geq 3$ and $\Gamma $ is either a uniform pro-p-group or a p-group of nilpotence class $<p$ and let $\mathfrak {g} = \mathrm {Lie}\, \Gamma $ . Then there exists a bijection $\Omega \leftrightarrow \rho _\Omega $ between $\Gamma $ -orbits $\Omega \subseteq \mathfrak {g}^\ast $ and $\widehat \Gamma $ , characterized by

$$\begin{align*}\operatorname{\mathrm{tr}}(g,\rho_\Omega) = \frac{1}{\#\Omega^{1/2}} \cdot \sum_{f\in \Omega} f(\log(g)). \end{align*}$$

Groups of the form $\Gamma = (\mathbb {G}_r^+)^F$ may or may not satisfy the assumptions of Theorem 8.1, as the following examples show.

Example 8.2. Suppose $r = \infty $ . Then $\Gamma = (\mathbb {G}_r^+)^F$ is the maximal pro-p subgroup of the parahoric group $\mathcal {G}(\mathcal {O}_k)$ . If $\operatorname {\mathrm {char}} k = p$ , $\Gamma $ always contains torsion, and hence is never uniform. Suppose now that $\operatorname {\mathrm {char}} k = 0$ . Then $\Gamma $ might or might not be uniform. For example, $1 + p M_n(\mathbb {Z}_p)$ is uniform. However, if $k/\mathbb {Q}_p$ has ramification index $e>1$ , then $1+\varpi M_n(\mathcal {O}_k)$ is not uniform. In general, it is true that a topological group has the structure of a p-adic Lie group if and only if it contains an open uniform subgroup [Reference Dixon, du Satoy, Mann and Segal11, Theorems 8.1 and 4.2].

Example 8.3. Suppose $r<\infty $ . If ${\mathbf x}$ is hyperspecial, $\Gamma = (\mathbb {G}_r^+)^F$ is of nilpotency class $\leq r-1$ (as $f({\mathbf x})$ is integral for all $f \in \widetilde \Phi $ , $\mathbb {G}_r^+ = \mathbb {G}_r^1$ , and the subgroups $\{\mathbb {G}_r^i\}_{i=1}^{r}$ form a central series of length $r-1$ ). Thus, if $r \leq p$ , the orbit method applies to the finite p-group $\Gamma $ . In contrast to Example 8.2, there is no assumption on the characteristic of k.

8.2 Cohomological induction vs. the orbit method

For brevity, we write $\Gamma = (\mathbb {G}_r^+)^F$ and . Note that satisfies condition (i) or (ii) in §8.1 and let denote its Lie algebra. As is abelian, is not only a homeomorphism, but also an isomorphism of groups with inverse . Also, as is abelian, we may identify with . By Theorem 1.1, we get a map

where the second map is

$$\begin{align*}\chi \mapsto (-1)^{s_\chi} H_{s_\chi}(Y,\overline{\mathbb{Q}}_\ell)[\chi] \end{align*}$$

However, Theorem 8.1 gives a map

$$\begin{align*}\rho \colon \mathfrak{g}^\ast \rightarrow \mathfrak{g}^\ast/{\mathrm{Ad}} \, \Gamma \stackrel{\sim}{ \rightarrow } \widehat \Gamma, \end{align*}$$

where the first arrow is the natural projection. It is a natural question, how these two maps are related, and we make the following conjecture in this direction. Note that there is a canonical projection

$$\begin{align*}\delta \colon \mathfrak{g} \twoheadrightarrow \mathfrak{t} \end{align*}$$

(as on the level of (geometric points of) the Lie algebras, $\mathfrak {t}$ is the weight $0$ subspace of the adjoint representation of on $\mathfrak {g}$ ; then one takes Frobenius invariants). Let $\delta ^\ast \colon \mathfrak {t}^\ast \rightarrow \mathfrak {g}^\ast $ be the dual map.

Conjecture 8.4. We have $\rho \circ \delta ^\ast = R_{\mathrm {log}}$ .

With other words, if is a character, then Conjecture 8.4 predicts that $H_{s_\chi }(Y_r,\overline {\mathbb {Q}}_\ell )[\chi ] \cong \rho _{\Omega }$ , where $\Omega \in \mathfrak {g}^\ast /{\mathrm {Ad}}\, \Gamma $ is the orbit of . Note that to be able to state the conjecture, we need (only) Theorem 7.1, but to verify it in a special case in §8.3, we use Theorem 5.5.

Remark 8.5. (1) Combined with [Reference Boyarchenko and Drinfeld1, Theorems 2.9 and 2.11], Conjecture 8.4 allows a realization of $H_{s_\chi }(Y_r,\overline {\mathbb {Q}}_\ell )[\chi ]$ as an induced representation (at least in the case when $\Gamma $ is finite).

(2) In the light of Examples 8.2 and 8.3, Conjecture 8.4 says that $\chi \mapsto H_{s_\chi }(Y_r,\overline {\mathbb {Q}}_\ell )[\chi ]$ is a generalization of the orbit method (for those adjoint orbits containing an element of $\mathfrak {t}$ ) to all groups of the form $\Gamma = (\mathbb {G}_r^+)^F$ . The collection of all such groups is neither contained in, nor containing the family of groups for which the orbit method applies.

8.3 An example

Assume that $\mathrm {char}(k) = p> 2$ , let $G = \mathrm {GL}_2$ and $r=3$ . Let $\mathcal {G}$ be the standard model of G over $\mathcal {O}_k$ . We verify Conjecture 8.4 in this case, that is for the group

$$\begin{align*}\Gamma = 1 + \varpi M_2(\mathbb{F}_q[\![\varpi]\!])/1+ \varpi^3 M_2(\mathbb{F}_q[\![\varpi]\!]), \end{align*}$$

where $M_2$ denotes the $2\times 2$ -matrices. ( $\Gamma $ is of nilpotency class $2<p$ , and hence, the orbit method applies.)

Write $\overline {R} := \overline {\mathbb {F}}_q[\varpi ]/(\varpi ^2)$ with Frobenius $\sigma (a + \varpi b) = a^q + \varpi b^q$ and let $R := \overline {R}^\sigma $ and $R_2 := \overline {R}^{\sigma ^2}$ . Write

$$\begin{align*}x(g_1,g_3) := 1 + \varpi \left(\begin{smallmatrix} g_1 & \sigma(g_3) \\ g_3 & \sigma(g_1) \end{smallmatrix}\right) \in 1 + \varpi M_2(\overline{R}) \cong \frac{1 + \varpi M_2(\overline{\mathbb{F}}_q[\![\varpi]\!])}{1+ \varpi^3 M_2(\overline{\mathbb{F}}_q[\![\varpi]\!])} \end{align*}$$

with $g_i = g_{i0} + \varpi g_{i1} \in \overline {R}$ for $i=1,3$ . Let be the twisted Frobenius on $1 + \varpi M_2(\overline {R})$ , such that the diagonal torus in $\Gamma $ becomes the unramified elliptic torus. We get a presentation of $\Gamma $ as

$$\begin{align*}\Gamma \cong \left(1 + \varpi M_2(\overline{R})\right)^F = \left\{ x(g_1,g_3) \colon g_1,g_3 \in R_2 \text{ for } i=1,3 \right\}, \end{align*}$$

which will be in use until the end of §8.3. Then and the corresponding deep level Deligne–Lusztig space $Y_3$ is given by

$$\begin{align*}Y_3 = \left\{ x(v_1,v_3) \in 1 + \varpi M_2(\overline{R}) \colon \det x(v_1,v_3) \in R ^\times \right\}. \end{align*}$$

The condition $\det x(v_1,v_3) \in R ^\times $ is equivalent to the conditions $v_{10} \in \mathbb {F}_{q^2}$ and $v_{11}^{q^2} - v_{11} = v_{30}^{q^2+q} - v_{30}^{q+1}$ . Next, we describe how $\Gamma $ and act on a point $x(v_1,v_3) \in Y_3$ . Let with $\tau = \tau _0 + \varpi \tau _1 \in R_2$ . Then we have

$$\begin{align*}x(v_1,v_3).t = x(v_{10} + \tau_0 + \varpi(v_{11} + \tau_1 + v_{10}\tau_0),v_{30} + \varpi (v_{31} + v_{30}\tau_0)). \end{align*}$$

Let $g = g(g_1,g_3) \in \Gamma $ . Then

$$ \begin{align*} g.x(v_1,v_3) = x(v_{10} + g_{10} + &\varpi(v_{11} + g_{11} + g_{10}v_{10} + g_{30}^q v_{30}), \\ &v_{30} + g_{30} + \varpi(v_{31} + g_{31} + g_{30}v_{10} + g_{10}^q v_{30})). \end{align*} $$

Lemma 8.6. There exists a constant $C \in \mathbb {Q}^\times $ such that for all $g = x(g_1,g_3) \in \Gamma $ , one has

$$\begin{align*}\operatorname{\mathrm{tr}}(g, H_{s_\chi}(Y_3,\overline{\mathbb{Q}}_\ell)[\chi]) = \begin{cases} Cq\cdot \chi(x(g_1, 0)) & \text{if } g_{30} = 0, \\ C \cdot\sum\limits_{\lambda \in \mathbb{F}_{q^2} \colon \lambda^{q}+\lambda = g_{30}^{q+1}} \chi(x(g_{10} + \varpi (g_{11} - \lambda), 0)) & \text{otherwise.} \end{cases} \end{align*}$$

Proof. $Y_3$ is defined over $\mathbb {F}_{q^2}$ . Combining Theorem 5.5 with [Reference Boyarchenko3, Lemma 2.12], we see that there is some $C_1 \in \mathbb {Q}^\times $ , such that for any $g = g(g_1,g_3) \in G$ ,

$$\begin{align*}\operatorname{\mathrm{tr}}(g,|R_\chi|) = C_1 \cdot \sum_{t \in T} \#S_{g,t} \cdot \chi(t), \end{align*}$$

where

$$\begin{align*}S_{g,t} = \{x \in X \colon g.F^2(x) = x.t \}. \end{align*}$$

Write $t = x(\tau ,0)$ . Using the above description of the actions on X, one easily sees that $S_{g,t} = \varnothing $ unless $g_{10} = \tau _0$ . Assume that $g_{10} = \tau _0$ holds. Using the determinant condition above, one easily deduces that $\#S_{g,t} = q^6 \cdot \#S_{g,t}'$ , where

$$\begin{align*}S_{g,t}' = \{v_{30} \in \mathbb{F}_{q^2} \colon v_{30} - v_{30}^{q^2} = g_{30}\text{ and } g_{11} - \tau_1 - g_{30}^{q+1}= v_{30}^q g_{30} - v_{30}g_{30}^q \}. \end{align*}$$

If $g_{30} = 0$ , the claim of the lemma becomes clear now. Assume $g_{30}\neq 0$ . Suppose first that $\tau _1$ is such that $S_{g,t}' \neq \varnothing $ . Then, if $v_{30} \in S_{g,t}'$ is arbitrary, writing $y := v_{30}^q g_{30} - v_{30}g_{30}^q$ , we see (using that $v_{30}^{q^2} = v_{30}-g_{30}$ ) that $y^q = -y - g_{30}^{q+1}$ . But on the other hand, $g_{11} - \tau _1 = y + g_{30}^{q+1}$ , and hence, we deduce (using that $g_{30} \in \mathbb {F}_q$ ) that

$$\begin{align*}(g_{11} - \tau_1)^q + (g_{11} - \tau_1) = (y + g_{30}^{q+1})^q + (y + g_{30}^{q+1}) = y^q + y + 2g_{30}^{q+1} = g_{30}^{q+1}. \end{align*}$$

With other words, $S_{g,t}' = \varnothing $ , unless

(8.1) $$ \begin{align} (g_{11} - \tau_1)^q + (g_{11} - \tau_1) = g_{30}^{q+1}. \end{align} $$

Assume now that this equality holds. Note that $v_{30}^q g_{30} - v_{30}g_{30}^q = g_{11} - \tau _1 - g_{30}^{q+1}$ , regarded as an equation in $v_{30}$ , has precisely q different solutions in $\overline {\mathbb {F}}_q$ (as $g_{30} \neq 0$ ). Moreover, if $v_{30}$ is one of its solutions, then (applying the transformation $X \mapsto X^q + X$ to both sides of this equation) one verifies using (8.1) that $v_{30}$ also satisfies $v_{30}^{q^2} - v_{30} = -g_{30}$ ; that is $v_{30} \in S_{g,t}'$ . Altogether, $\#S_{g,t}' = q$ if (8.1) holds and $\#S_{g,t}' = 0$ otherwise. The lemma follows immediately from this by taking $\lambda := g_{11} - \tau _1$ for those $\tau _1$ which satisfy (8.1).

Now we consider the orbit method side. Write $y(g_1,g_3) := x(g_1,g_3) - 1 \in \varpi M_2(\overline {R}) = \mathfrak {g} = \mathrm {Lie}\, \Gamma $ . The map $\log \colon \Gamma \rightarrow \mathfrak {g}$ is given by $\log (1+\varpi z) = \varpi z - \frac {\varpi ^2 z}{2}$ . Let

$$\begin{align*}\delta \colon \mathfrak{g} \rightarrow \mathfrak{t}, \quad y(g_1,g_3) \mapsto y(g_1, 0) \quad \text{ and let } \quad \varepsilon := \chi \circ \exp_T \circ \delta \in \mathfrak{g}^\ast. \end{align*}$$

Consider first the $\Gamma $ -orbit $\Omega _\delta $ of $\delta $ ( $\Gamma $ acts on the first factor in $\operatorname {\mathrm {Hom}}(\mathfrak {g},\mathfrak {t})$ by conjugation). First, note that the action of $\Gamma $ factors through $\Gamma = (\mathbb {G}_3^+)^F \twoheadrightarrow (\mathbb {G}_2^+)^F = 1+ \varpi M_2(\mathbb {F}_{q^2})$ . Moreover, for $h = x(h_{10},h_{30})\in (\mathbb {G}_2^+)^F$ , we have

$$ \begin{align*} ({\mathrm{Ad}} h)(\delta)(y(g_1,g_3)) &= \delta(hy(g_1,g_3)h^{-1}) \\ &= y(g_{10} + \varpi(g_{11} + h_{10}^q g_{30} - h_{10} g_{30}^q),0) =: \delta_{h_{10}}(g). \end{align*} $$

Thus, $\Omega _\delta = \{\delta _{h_{10}} \colon h_{10} \in \mathbb {F}_{q^2}\}$ has cardinality $q^2$ . As is an isomorphism, the $\Gamma $ -orbit of has the same cardinality as $\Omega _\delta $ . Let now $h_{10} \neq h_{10}' \in \mathbb {F}_{q^2}$ . An easy computation shows that if and only if $\chi |_{1 + \varpi ^2\mathbb {F}_q^{-}}$ is trivial, where we set $\mathbb {F}_q^- := \{x \in \mathbb {F}_{q^2} \colon x+x^q = 0\}$ .

Suppose first that $\chi |_{1 + \varpi ^2\mathbb {F}_q^{-}}$ is nontrivial. Then composition with induces a bijection $\Omega _\delta \stackrel {\sim }{ \rightarrow } \Omega _{\varepsilon }$ . Unraveling the trace formula from Theorem 8.1 we then get that for $g = x(g_1,g_3)$ ,

(8.2) $$ \begin{align} \operatorname{\mathrm{tr}}(g,\rho_{\Omega_{\varepsilon}}) = C_2 \cdot \sum_{\alpha \in \mathbb{F}_{q^2}}\chi\left(x\left(g_{10} + \varpi \left(g_{11} - \frac{g_{30}^{q+1}}{2} + \alpha^q g_{30} - \alpha g_{30}^q\right)\right)\right), \end{align} $$

for some constant $C_2 \in \mathbb {Q}^\times $ . If $g_{30} = 0$ , this clearly agrees with the trace from Lemma 8.6 up to a (non-zero) scalar. Assume $g_{30} \neq 0$ . Then the homomorphism $\alpha \mapsto \alpha ^q g_{30} - \alpha g_{30}^q \colon \mathbb {F}_{q^2} \rightarrow \mathbb {F}_{q^2}$ is easily seen to have image $\mathbb {F}_q^-$ . Thus, (8.2) transforms to

$$\begin{align*}\operatorname{\mathrm{tr}}(g,\rho_{\Omega_{\varepsilon}}) = C_2 \cdot q \sum_{\mu \in \mathbb{F}_q^-}\chi \left(x\left(g_{10}+\varpi \left(g_{11} - \frac{g_{30}^{q+1}}{2} + \mu \right)\right)\right). \end{align*}$$

Now it is immediate to check that the map $\mu \mapsto \lambda := \frac {g_{30}^{q+1}}{2} - \mu $ defines a bijection between $\mathbb {F}_q^-$ and the set $\{\lambda \in \overline {\mathbb {F}}_q \colon \lambda ^q + \lambda = g_{30}^{q+1}\}$ . Thus, the trace of $\rho _{\Omega _\varepsilon }$ agrees with the trace from Lemma 8.6 up to a non-zero scalar, which does not depend on g. As we know that $H_{s_\chi }(Y_3,\overline {\mathbb {Q}}_\ell )[\chi ]$ and $\rho _{\Omega _\varepsilon }$ are both irreducible $\Gamma $ -representations, it follows that they must be isomorphic.

In the case that $\chi |_{1 + \varpi ^2\mathbb {F}_q^{-}}$ is trivial, a similar (and easier) computation leads to the same conclusion. Altogether, we have shown the following:

Proposition 8.7. For , we have $H_{s_\chi }(Y_3,\overline {\mathbb {Q}}_\ell )[\chi ] \cong \rho _{\Omega _{\varepsilon }}$ as $\Gamma $ -representations. Thus, Conjecture 8.4 holds for $\Gamma $ .

A Algorithm for the Steinberg cross-section

The algorithm used in the proof of Proposition 3.1 consists of two procedures (implemented in SAGE, v8.6), which we now describe.

find_candidate_for_one_step (procedure 1):

Input: an element $w \in W$ , a set $\Psi \subsetneq \Phi ^+$ of positive roots

Output: a (non-empty) set of positive roots or False.

  1. 1. Compute the set $\Phi _w = \{\alpha \in \Phi ^+ \colon w\sigma (\alpha ) < 0\}$ .

  2. 2. Set $\Phi ^+ {\,\setminus \,} \Psi = \{\beta _1,\dots ,\beta _s\}$ with $s \geq 1$ .

  3. 3. For i running through $1, 2, \dots , s$ do:

    1. 3.1. Set $\Psi _1^{(i)} = \Psi \cup \{\beta _i\}$ and $\Psi _2^{(i)} = \Psi _1^{(i)} {\,\setminus \,} \Phi _w$ .

    2. 3.2. Check whether the following conditions hold: (a) $\Psi _1^{(i)}$ and $\Psi _2^{(i)}$ are closed under addition; (b) for all $\alpha ,\beta \in \Psi _1^{(i)}$ such that $\alpha +\beta \in \Phi ^+$ , one has $\alpha + \beta \in \Psi _2^{(i)}$ ; (c) $w\sigma (\Psi _2^{(i)}) \subseteq \Psi _1^{(i)}$ .

    3. 3.3 If (a)–(c) hold, return $\Psi _1^{(i)}$ and stop. Otherwise, continue with the next i.

  4. 4. Return False.

iterate_steps (procedure 2):

Input: an element $w \in W$ , and $\Psi $ , which is either a subset of $\Phi ^+$ or False.

Output: a (non-empty) set of positive roots or False or True.

  1. 1. Compute the set $\Phi _w := \{\alpha \in \Phi ^+ \colon w\sigma (\alpha ) < 0\}$ .

  2. 2. If $\Phi _w = \Phi ^+$ , return True and stop.

  3. 3. If find_candidate_for_one_step $(w, \Psi ) =$ False, return False and stop.

  4. 4. If $\Psi = \Phi ^+$ , return True and stop.

  5. 5. Set $\Psi ' := $ find_candidate_for_one_step $(w,\Psi )$ . Return iterate_steps $(w,\Psi ')$ .

To check if Lemma [Reference Ivanov20, Lemma 5.7] holds for an element $w \in W$ , one runs the (recursive) procedure iterate_steps with arguments w and $\Phi _w = \{\alpha \in \Phi ^+ \colon w\sigma (\alpha ) < 0\}$ . The recursion stops after finitely many steps. If the final output is True, the lemma holds. This holds true if w is twisted Coxeter.

Remark A.1. Note that the final output True of iterate_steps $(w,\Phi _w)$ is a sufficient but not a necessary condition for Lemma [Reference Ivanov20, Lemma 5.7] to hold for $w \in W$ . In fact, there are (non-Coxeter) elements $w \in W$ for which [Reference Ivanov20, Lemma 5.7] holds true, but iterate_steps $(w,\Phi _w)$ outputs False.

Acknowledgements.

The first author gratefully acknowledges the support of the German Research Foundation (DFG) via the Heisenberg program (grant nr. 462505253). He would like to thank Moritz Firsching for answering his questions related to SAGE. The second author would like to thank Miaofen Chen, Xuhua He, Jilong Tong and Weizhe Zheng for answering his questions and for helpful discussions. The second author is partially supported by National Key R&D Program of China, No. 2020YFA0712600, CAS Project for Young Scientists in Basic Research, Grant No. YSBR-003, and National Natural Science Foundation of China, Nos. 11922119, 12288201, and 12231001.

Competing interests

On behalf of all authors, the corresponding author states that there is no conflict of interest.

Data availability statement

Data sharing is not applicable to this article as no datasets were generated or analyzed during the current study.

Footnotes

1 This holds if the derived group of G is simply connected and $p\geq 5$ ; it also always holds if p does not divide the order of the Weyl group of G.

2 Note that for $H=T$ , there is no conflict of notation with §2.3 as the closure of T in $\mathcal {G}$ is the connected Néron model of T by [Reference Yu32, 4.7.4 Lemma and 8.2 Corollary].

3 This assumption can be weakened at the cost of more technical results.

References

Boyarchenko, M and Drinfeld, V (2010) A motivated introduction to character sheaves and the orbit method for unipotent groups in positive characteristic. arXiv preprint math/0609769.Google Scholar
Boyarchenko, M (2010) Characters of unipotent groups over finite fields. Selecta Math. (N.S.) 16(4), 857933.CrossRefGoogle Scholar
Boyarchenko, M (2012) Deligne-Lusztig constructions for unipotent and $p$ -adic groups. arXiv preprint arXiv:1207.5876.Google Scholar
Boyarchenko, M and Sabitova, M (2008) The orbit method for profinite groups and a $p$ -adic analogue of Brown’s theorem. Israel J. Math. 165, 6791.CrossRefGoogle Scholar
Boyarchenko, M and Weinstein, J (2016) Maximal varieties and the local Langlands correspondence for $\mathrm{GL}(n)$ . J. Amer. Math. Soc. 29, 177236.CrossRefGoogle Scholar
Chan, C (2020) The cohomology of semi-infinite Deligne–Lusztig varieties. J. Reine Angew. Math. (Crelle) 2020(768), 93147.CrossRefGoogle Scholar
Chan, C and Ivanov, AB (2019) Cohomological representations of parahoric subgroups. Represent. Theory 25, 126.CrossRefGoogle Scholar
Chan, C and Ivanov, AB (2021) The Drinfeld stratification for ${\mathrm{GL}}_{\mathrm{n}}$ . Selecta Math. New Ser. 27(50).CrossRefGoogle Scholar
Chan, C and Ivanov, AB (2023) On loop Deligne–Lusztig varieties of Coxeter type for ${\mathrm{GL}}_{\mathrm{n}}$ . Camb. J. Math. 11(2) (2023), 441505.CrossRefGoogle Scholar
Chan, C and Oi, M (2023) Geometric $L$ -packets of Howe-unramified toral supercuspidal representations. J. Eur. Math. Soc. 162, doi: 10.4171/JEMS/1396.Google Scholar
Dixon, JD, du Satoy, MPF, Mann, A and Segal, D (1999) Analytic Pro- $p$ -groups, 2nd edn. Advanced Mathematics, vol. 61. Cambridge: Cambridge University Press.Google Scholar
Deligne, P and Lusztig, G (1976) Representations of reductive groups over finite fields. Ann. Math. 103(1), 103161.CrossRefGoogle Scholar
Dudas, O and Ivanov, A (2024) Orthogonality relations for deep level Deligne-Lusztig schemes of Coxeter type. Forum Math. Sigma. 12(e660), 127.CrossRefGoogle Scholar
Feng, T (2024) Modular functoriality in the Local Langlands Correspondence. arXiv preprint 2312.12542.Google Scholar
Fargues, L and Scholze, P (2021) Geometrization of the local langlands correspondence. arXiv preprint 2102.13459.Google Scholar
Gordon, A (2024) The Closed Stratum of a Parahoric Deligne-Lusztig Variety. PhD dissertation, University of Michigan.Google Scholar
He, X and Lusztig, G (2012) A generalization of Steinberg’s cross-section. J. Amer. Math. Soc. 25(3) (2012), 739757.CrossRefGoogle Scholar
Ivanov, AB and Mann, L. The 5-functor formalism for schemes. In preparation.Google Scholar
Ivanov, AB (2023) Arc-descent for the perfect loop functor and $p$ -adic Deligne–Lusztig spaces. J. Reine Angew. Math. (Crelle) 2023(794), 154.Google Scholar
Ivanov, AB (2023) On a decomposition of $p$ -adic Coxeter orbits. Épijournal de Géom. Alg7.Google Scholar
Kaletha, T (2019) Regular supercuspidal representations. J. Amer. Math. Soc. 32, 10711170.CrossRefGoogle Scholar
Kirillov, AA, Unitary representations of nilpotent Lie groups. Uspehi Mat. Nauk 106(4) (1962), 57110.Google Scholar
Malten, W (2021) From braids to transverse slices in reductive groups. arXiv preprint 2111.01313.Google Scholar
Moy, A and Prasad, G (1994) Unrefined minimal K-types for $p$ -adic groups. Invent. Math. 116, 393408.CrossRefGoogle Scholar
Nie, S (2024) Steinberg’s cross-section of Newton strata. Math. Ann. https://doi.org/10.1007/s00208-024-02872-2.CrossRefGoogle Scholar
Schwein, D (2024) Formal degree of regular supercuspidals. J. Eur. Math. Soc. 1–53. doi: 10.4171/JEMS/1412.Google Scholar
The Stacks Project Authors (2014) Stacks Project, http://stacks.math.columbia.edu.Google Scholar
Steinberg, R (1965) Regular elements of semisimple algebraic groups. Publ. Math. IHES 25, 4980.CrossRefGoogle Scholar
Steinberg, R (1975) Torsion in reductive groups. Adv. Math. 15, 6392.CrossRefGoogle Scholar
The Sage Developers (2022) SageMath, the Sage Mathematics Software System. doi: 10.5281/zenodo.6259615.CrossRefGoogle Scholar
Yu, JK (2001) Construction of tame supercuspidal representations. J. Amer. Math. Soc. 14, 579622.CrossRefGoogle Scholar
Yu, JK (2015) Smooth models associated to concave functions in Bruhat-Tits theory. In Autour des schémas en groupes, Vol. III, vol. 47. Panor. Syntheses.Google Scholar
Zhu, X (2017) Affine Grassmannians and the geometric Satake in mixed characteristic. Ann. Math. 185(2), 403492.CrossRefGoogle Scholar
Zhu, X (2020) Coherent sheaves on the stack of Langlands parameters. arXiv preprint 2008.02998.Google Scholar