Hostname: page-component-745bb68f8f-d8cs5 Total loading time: 0 Render date: 2025-02-06T19:55:24.177Z Has data issue: false hasContentIssue false

A group-theoretic generalization of the p-adic local monodromy theorem

Published online by Cambridge University Press:  29 June 2021

Shuyang Ye*
Affiliation:
Department of Mathematics, East China Normal University, Shanghai, China
Rights & Permissions [Opens in a new window]

Abstract

Let G be a connected reductive group over a p-adic number field F. We propose and study the notions of G- $\varphi $ -modules and G- $(\varphi ,\nabla )$ -modules over the Robba ring, which are exact faithful F-linear tensor functors from the category of G-representations on finite-dimensional F-vector spaces to the categories of $\varphi $ -modules and $(\varphi ,\nabla )$ -modules over the Robba ring, respectively, commuting with the respective fiber functors. We study Kedlaya’s slope filtration theorem in this context, and show that G- $(\varphi ,\nabla )$ -modules over the Robba ring are “G-quasi-unipotent,” which is a generalization of the p-adic local monodromy theorem proved independently by Y. André, K. S. Kedlaya, and Z. Mebkhout.

Type
Article
Copyright
© Canadian Mathematical Society 2021

1 Introduction

Let p be a prime number and q a power of p. Let K be a complete nonarchimedean discretely valued field of characteristic $0$ equipped with an automorphism $\varphi $ , the Frobenius, inducing the q-power map on the residue field $\kappa \supseteq \mathbb {F}_q$ . We also require K to be unramified over the fixed subfield F under $\varphi $ . See Hypothesis 2.1 for a concrete example.

The Robba ring $\mathcal {R}=\mathcal {R}(K,t)$ is the ring of bidirectional power series $\sum \limits _{i\in \mathbb {Z}}c_it^i$ in one variable t with coefficients in K which converge in an annulus $[\alpha ,1)$ for some series-dependent $0<\alpha <1$ . The Robba ring $\mathcal {R}$ is endowed with an absolute Frobenius lift $\varphi $ which extends the Frobenius on K and lifts the q-power map on $\kappa (\!( t )\!)$ , and with the derivation $\partial =d/dt$ .

A $(\varphi ,\nabla )$ -module over $\mathcal {R}$ is a triple $(M,\Phi ,\nabla )$ , where M is a finite free $\mathcal {R}$ -module, $\Phi $ is a Frobenius, i.e., a $\varphi $ -linear endomorphism of M whose image spans M over $\mathcal {R}$ , and $\nabla \colon M\to M\bigotimes _{\mathcal {R}} \mathcal {R} dt$ is a connection. Moreover, $\Phi $ and $\nabla $ should satisfy the gauge compatibility condition, which says that, after choosing an $\mathcal {R}$ -basis for M, the actions $\Phi $ and $\nabla $ are given by matrices A and N, respectively, and these matrices should satisfy $N=\operatorname {\mathrm {\boldsymbol \mu }}\cdot A(\varphi (N))A^{-1}-\partial (A)A^{-1}$ , where $\operatorname {\mathrm {\boldsymbol \mu }}:= \partial (\varphi (t))$ .

The $(\varphi ,\nabla )$ -modules, also known as the overconvergent $\mathrm {F}$ -isocrystals in the literature, are closely related to p-adic local systems on $\operatorname {\mathrm {Spec}} \kappa (\!( t )\!)$ (for a summary, we refer to [Reference Kedlaya13]), for which the correct monodromy theorem is the p-adic local monodromy theorem (pLMT), conjectured by Crew [Reference Crew5], and proved independently by André [Reference André1], Kedlaya [Reference Kedlaya9], and Mebkhout [Reference Mebkhout17]. It states that every $(\varphi ,\nabla )$ -module over $\mathcal {R}$ is quasi-unipotent. Concretely, a $(\varphi ,\nabla )$ -module M over $\mathcal {R}$ , after an étale extension to $\mathcal {R}_L$ (the Robba ring canonically associated to some finite separable extension L of $\kappa (\!( t )\!)$ ), admits a filtration by $(\varphi ,\nabla )$ -submodules such that the connections induced on the gradiation are trivial. A matricial description of the theorem is given as follows. Let d be the rank of M over $\mathcal {R}$ , and let $A\in \operatorname {\mathrm {GL}}_d(\mathcal {R})$ (resp. $N\in \operatorname {\mathrm {Mat}}_{d,d}(\mathcal {R})$ ) be the matrix of $\Phi $ (resp. $\nabla $ ) in some basis. Then, there exists $B\in \operatorname {\mathrm {GL}}_d(\mathcal {R}_L)$ such that $BNB^{-1}-\partial (B)B^{-1}$ is an upper-triangular block matrix with zero blocks in the diagonal.

We mention two applications of the pLMT in p-adic Hodge theory.

  • In [Reference Berger3], Berger associated to every p-adic de Rham representation V a $(\varphi ,\nabla )$ -module $\mathrm {N}_{\operatorname {\mathrm {dR}}}(V)$ over a Robba ring. He showed that V is potentially semistable if and only if $\mathrm {N}_{\operatorname {\mathrm {dR}}}(V)$ is quasi-unipotent. Using the pLMT, he could prove the p-adic monodromy theorem (previously a conjecture of Fontaine): every p-adic de Rham representation is potentially semistable.

  • In [Reference Marmora16], Marmora used the pLMT to construct a functor from the category of $(\varphi ,\nabla )$ -modules over $\mathcal {R}$ to that of $K^{\operatorname {\mathrm {nr}}}$ -valued Weil–Deligne representations of the Weil group $\mathcal {W}_{\kappa (\!( t )\!)}$ , where $K^{\operatorname {\mathrm {nr}}}$ is the maximal unramified extension of K in a fixed algebraic closure of K.

Rather than the general linear group, a Galois representation may take value in some connected reductive group G, such as a special linear group or a symplectic group. In order to have appropriate formulations of the above results in this context, it is helpful to establish a G-version of the pLMT, which is the main motivation of our present paper.

In this paper, we introduce the notion of G- $\varphi $ -modules over $\mathcal {R}$ (resp. G- $(\varphi ,\nabla )$ -modules over $\mathcal {R}$ ), which are exact faithful F-linear tensor functors from the category $\operatorname {\mathrm {\textbf {Rep}}}_F(G)$ of G-representations on finite-dimensional F-vector spaces to the category $\operatorname {\mathrm {\textbf {Mod}^{\varphi }_{\mathcal {R}}}}$ of $\varphi $ -modules over $\mathcal {R}$ (resp. to the category $\operatorname {\mathrm {\textbf {Mod}^{\varphi ,\nabla }_{\mathcal {R}}}}$ of $(\varphi ,\nabla )$ -modules over $\mathcal {R}$ ), commuting with the respective fiber functors. These constructions are inspired by that of G-isocrystals introduced in [Reference Dat, Orlik and Rapoport6, Section IX.1].

Before coming to the main theorem, we first explain the group-theoretic gauge compatibility condition (Definition 4.6). Let G be an affine algebraic F-group, and let $\mathfrak {g}$ be its Lie algebra. For any $y\in G(\mathcal {R})$ and $Y\in \mathfrak {g}\bigotimes _F \mathcal {R}$ , we define $\Gamma _y(Y) := \operatorname {\mathrm {Ad}}(y)(Y)-\operatorname {\mathrm {dlog}}(y)$ , where $\operatorname {\mathrm {Ad}} \colon G \to \operatorname {\mathrm {GL}}_{\mathfrak {g}}$ is the adjoint representation, and $\operatorname {\mathrm {dlog}} \colon G(\mathcal {R}) \to \mathfrak {g}\bigotimes _F\mathcal {R}$ is defined in Construction 4.4. We say $g\in G(\mathcal {R})$ and $X\in \mathfrak {g}\bigotimes _F\mathcal {R}$ satisfy the gauge compatibility condition if $X=\Gamma _g\big (\operatorname {\mathrm {\boldsymbol \mu }}\varphi (X) \big )$ . When $G=\operatorname {\mathrm {GL}}_d$ , we have $\operatorname {\mathrm {Ad}}(y)(Y)=yYy^{-1}$ and $\operatorname {\mathrm {dlog}}(y)=\partial (y)y^{-1}$ . In this context, the group-theoretic gauge compatibility condition coincides with the aforementioned matrical one.

Our main theorem is the following G-version of the pLMT.

Theorem 1.1 (Theorem 4.21)

Let G be a connected reductive F-group, and let $\mathfrak {g}$ be its Lie algebra. If $g\in G(\mathcal {R})$ and $X\in \mathfrak {g}\bigotimes _F\mathcal {R}$ satisfy $X=\Gamma _g\big (\operatorname {\mathrm {\boldsymbol \mu }}\varphi (X) \big )$ , then there exists a finite separable extension L over $\kappa (\!( t )\!)$ and an element $b\in G(\mathcal {R}_L)$ such that $\Gamma _b(X)\in \operatorname {\mathrm {Lie}} \big (U_{G_{\mathcal {R}}}(-\lambda _g)\big )\bigotimes _{\mathcal {R}} {\mathcal {R}_L}$ .

Here, $\lambda _g \colon \mathbb {G}_{m,\mathcal {R}} \to G_{\mathcal {R}}$ is a cocharacter associated to g whose reciprocal is denoted by $-\lambda _g$ , and $U_{G_{\mathcal {R}}}(-\lambda _g)$ denotes the unipotent radical of the parabolic subgroup of $G_{\mathcal {R}}$ associated to $-\lambda _g$ . When $G=\operatorname {\mathrm {GL}}_d$ , g (resp. X) should be thought as the matrix of the Frobenius (resp. the matrix of the connection), and $\Gamma _b(\rule {2mm}{0.15mm})$ as the matrix of a connection under the change-of-basis via $b^{-1}$ . Moreover, $\operatorname {\mathrm {Lie}} \big (U_{G_{\mathcal {R}}}(-\lambda _g)\big )\bigotimes _{\mathcal {R}} {\mathcal {R}_L}$ consists of upper-triangular matrices over $\mathcal {R}_L$ with zero blocks (of certain sizes) in the diagonal. As such, Theorem 1.1 recovers the matricial pLMT described above.

In Proposition 4.9, we show that G- $(\varphi ,\nabla )$ -modules over $\mathcal {R}$ are indeed pairs $(g,X)$ subject to the gauge compatibility condition in the theorem. In this sense, the theorem can be interpreted as saying that G- $(\varphi ,\nabla )$ -modules over $\mathcal {R}$ are “G-quasi-unipotent.” In Examples 4.10 and 4.11, we give examples of the existence of such pairs for G a special linear group and a symplectic group, respectively.

More examples of G- $(\varphi ,\nabla )$ -modules are expected from Berger’s functor $\mathrm {N}_{\operatorname {\mathrm {dR}}}$ mentioned previously. Explicitly, we hope to show in a future work that if a p-adic de Rham representation V takes value in a connected reductive group G, then $\mathrm {N}_{\operatorname {\mathrm {dR}}}(V)$ is a G- $(\varphi ,\nabla )$ -module. As another future work, we intend to use Theorem 1.1 to formulate a G-version of Marmora’s functor, namely, to construct a functor from the category of G- $(\varphi ,\nabla )$ -modules over $\mathcal {R}$ to that of Weil–Deligne representations of the Weil group $\mathcal {W}_{\kappa (\!( t )\!)}$ taking value in $G(K^{\operatorname {\mathrm {nr}}})$ .

Our approach to the theorem closely follows that of the pLMT in [Reference Kedlaya9] for absolute Frobenius lifts, wherein the author used his slope filtration theorem (along with applying the pushforward functor and twisting to each quotient of the filtration) to reduce the problem to the unit-root case, and then apply the unit-root pLMT attributed to Tsuzuki [Reference Tsuzuki23] to finish. More precisely, we use Kedlaya’s slope filtration theorem to construct a $\mathbb {Q}$ -filtered fiber functor $\operatorname {\mathrm {{HN}}}_g$ from $\operatorname {\mathrm {\textbf {Rep}}}_F(G)$ to $ \mathbb {Q}$ - $\operatorname {\mathrm {\textbf {Fil}}}_{\mathcal {R}}$ , the category of $\mathbb {Q}$ -filtered modules over $\mathcal {R}$ (see Theorem 3.4). We then reduce $\operatorname {\mathrm {{HN}}}_g$ to a $\mathbb {Z}$ -filtered fiber functor $\operatorname {\mathrm {{HN}}}_g^{\mathbb {Z}}$ from $\operatorname {\mathrm {\textbf {Rep}}}_F(G)$ to $\mathbb {Z}$ - $\operatorname {\mathrm {\textbf {Fil}}}_{\mathcal {R}}$ , the category of $\mathbb {Z}$ -filtered modules over $\mathcal {R}$ (see Lemma 3.10). Then, a result of Ziegler (Theorem 2.12) immediately implies that $\operatorname {\mathrm {{HN}}}_g^{\mathbb {Z}}$ is splittable, i.e., factors through a $\mathbb {Z}$ -graded fiber functor (see Proposition 3.11). In particular, for any splitting of $\operatorname {\mathrm {{HN}}}_g^{\mathbb {Z}}$ , we construct a morphism $\lambda _g\colon \mathbb {G}_{m,\mathcal {R}} \to G_{\mathcal {R}}$ of $\mathcal {R}$ -groups in Section 3.4, which is called the $\mathbb {Z}$ -slope morphism of g. With this, we can reduce the G- $(\varphi ,\nabla )$ -module $(g,X)$ over $\mathcal {R}$ , involving the (generalized) pushforward functor and twisting, to a unit-root one (see Corollary 4.20). Theorem 1.1 then follows from the unit-root pLMT and a Tannakian argument.

The paper is organized as follows. In Section 2, we set up basic notation and conventions, and then recall some necessary background on the theory of slopes and Tannakian formalism. In Section 3, we study G- $\varphi $ -modules over the Robba ring, and construct slope morphisms. In Section 4, we consider G- $(\varphi ,\nabla )$ -modules over the Robba ring, and prove our main result, Theorem 1.1, in the last subsection.

2 Preliminaries

2.1 Notation and conventions

When k is a field, we denote by $\operatorname {\mathrm {\textbf {Vec}}}_k$ the category of finite-dimensional k-vector spaces. When R is a k-algebra,Footnote 1 we denote by $\operatorname {\mathrm {\textbf {Mod}}}_R$ the category of R-modules, and by $\operatorname {\mathrm {\textbf {Alg}}}_R$ the category of R-algebras. When $V,W \in \operatorname {\mathrm {\textbf {Vec}}}_k$ , we write $V_R$ for $V\bigotimes _k R$ , and write $\alpha _R:= \alpha \otimes R$ , the R-linear extension of $\alpha $ , for all k-linear maps $\alpha \colon V\to W$ . When G is an affine algebraic k-group, we denote by $k[G]$ the Hopf algebra of G, by $G_R:= G \times _{\operatorname {\mathrm {Spec}} k} \operatorname {\mathrm {Spec}} R$ the base extension, by $H^1(k,G):= H^1 \big (\operatorname {\mathrm {Gal}}(k^{\operatorname {\mathrm {sep}}}/k),G(k^{\operatorname {\mathrm {sep}}}) \big )$ the first Galois cohomology set, and by $\operatorname {\mathrm {\textbf {Rep}}}_k(G)$ the category of representations of G on finite-dimensional k-vector spaces.

By a reductive k-group, we mean a (not necessarily connected) affine algebraic k-group G such that every smooth connected unipotent normal subgroup of $G_{\bar k}$ is trivial, where $\bar {k}$ is an algebraic closure of k.

For the rest of this paper, we work under the following hypothesis.

Hypothesis 2.1 Let p be a prime number and $q=p^f$ an integral power of p. Let F be a finite extension of $\mathbb {Q}_p$ with the ring of integers $\mathcal {O}_F$ , a fixed uniformizer $\varpi _F$ and the residue field $\mathbb {F}_q$ of q elements. Let $\kappa $ be a perfect field containing $\mathbb {F}_q$ . Let $\mathcal {O}_K=\mathcal {O}_F\bigotimes _{W(\mathbb {F}_q)} W(\kappa )$ , where $W(\mathbb {F}_q)$ (resp. $W(\kappa )$ ) denotes the ring of Witt vectors with coefficients in $\mathbb {F}_q$ (resp. in $\kappa $ ). Then, $K:= \operatorname {\mathrm {Frac}}(\mathcal {O}_K)\cong F\bigotimes _{W(\mathbb {F}_q)} W(\kappa )$ is a complete discretely valued field with ring of integers $\mathcal {O}_K$ , a uniformizer $\varpi := \varpi _F\otimes 1$ and residue field $\kappa $ . Let $\operatorname {\mathrm {Frob}}$ be the ring endomorphism of $W(\kappa )$ induced by the p-power map on $\kappa $ , and let

$$ \begin{align*}\varphi:= \operatorname{\mathrm{Id}}_F\otimes \operatorname{\mathrm{Frob}}^f \colon K\operatorname{\mathrm{\longrightarrow}} K\end{align*} $$

be the Frobenius automorphism on K relative to F. Then, $\varphi $ reduces to the q-power map on $\kappa $ , and the fixed field of $\varphi $ on K is $F\bigotimes _{W(\mathbb {F}_q)} W(\mathbb {F}_q) \cong F$ .

2.2 The Robba ring and its variants

For $\alpha \in (0,1)$ , we put

$$ \begin{align*}\mathcal{R}_\alpha:= \Big\{\operatorname*{\mathrm{\sum}}\limits_{i\in\mathbb{Z}} c_it^i~\Big|~\ c_i\in K, \lim\limits_{i\to\pm\infty} |c_i|\rho^i=0,~\forall\rho\in[\alpha,1) \Big\}.\end{align*} $$

For any $\rho \in [\alpha ,1)$ , we define the $\rho $ -Gauss norm on $\mathcal {R}_\alpha $ by setting $\big |\sum \limits _i c_it^i \big |_\rho :=\sup _i\{|c_i|\rho ^i\}$ . The Robba ring is defined to be the union $\mathcal {R} := \mathcal {R}(K,t):= \bigcup \limits _{\alpha \in (0,1)}\mathcal {R}_\alpha $ . For any $\sum \limits _i c_it^i\in \mathcal {R}$ , we define $\big |\sum \limits _i c_it^i\big |_1:=\sup _i\{|c_i|\}\in \mathbb {R}_{\geq 0}\cup \{\infty \}$ , the $1$ -Gauss norm.

The bounded Robba ring $\mathcal {E}^{\dagger }=\mathcal {E}^{\dagger }(K,t)$ is the subring of $\mathcal {R}$ consisting of bounded elements (i.e., elements with finite $1$ -Gauss norm), which is actually a Henselian discretely valued field w.r.t. the $1$ -Gauss norm with residue field $\kappa (\!( t)\!)$ .

Let $R\in \{\mathcal {R}, \mathcal {E}^{\dagger }\}$ . An absolute q-power Frobenius lift on R is a ring endomorphism $\varphi \colon R\to R$ given by $\sum \limits _{i\in \mathbb {Z}} c_it^i \longmapsto \sum \limits _{i\in \mathbb {Z}} \varphi (c_i)u^i$ for $u=\varphi (t)\in R$ such that $|u-t^q|_1<1$ .

For any $\alpha \in (0,1)$ , we define $\tilde {\mathcal {R}}_\alpha $ to be the ring of formal sums $\operatorname *{\mathrm {\sum }}\limits _{i\in \mathbb {Q}} c_it^i$ with $c_i\in K$ , subject to the following properties.

  • For any $c>0$ , the set $\{i\in \mathbb {Q}\mid |c_i|\geq c\}$ is well-ordered.

  • For any $\rho \in [\alpha ,1)$ , we have $\lim \limits _{i\to \pm \infty } |c_i|\rho ^i=0$ .

For any $\rho \in [\alpha ,1)$ , we define the $\rho $ -Gauss norm on $\tilde {\mathcal {R}}_\alpha $ by setting

$$ \begin{align*}\Big|\operatorname*{\mathrm{\sum}}\limits_i c_it^i \Big|_\rho=\sup\limits_{i\in\mathbb{Q}}\{|c_i|\rho^i\}.\end{align*} $$

We define $\tilde {\mathcal {R}} := \tilde {\mathcal {R}}(K,t):= \bigcup \limits _{\alpha \in (0,1)} \tilde {\mathcal {R}}_\alpha $ , the extended Robba ring. The absolute Frobenius lift on $\tilde {\mathcal {R}}$ is the ring automorphism on $\operatorname {\mathrm {\tilde {\mathcal {R}}}}$ given by $\sum \limits _{i\in \mathbb {Q}} c_it^i \mapsto \sum \limits _{i\in \mathbb {Q}} \varphi (c_i)t^{iq}$ . We denote by $\operatorname {\mathrm {\tilde {\mathcal {E}}}}^{\dagger }$ the subring of $\operatorname {\mathrm {\tilde {\mathcal {R}}}}$ consisting of bounded elements. By [Reference Kedlaya11, Proposition 2.2.6], we have a $\varphi $ -equivariant embedding $\psi :\mathcal {R}\to \tilde {\mathcal {R}}$ such that $|\psi (x)|_\rho =|x|_\rho $ for $\rho $ sufficiently close to $1$ .

2.3 The slope filtration theorem

We recall Kedlaya’s theory of slopes. Let $R\in \{\mathcal {E}^{\dagger },\mathcal {R},\operatorname {\mathrm {\tilde {\mathcal {E}}}}^{\dagger },\operatorname {\mathrm {\tilde {\mathcal {R}}}}\}$ equipped with a Frobenius lift $\varphi $ . For the notions of $\varphi $ -modules and $(\varphi ,\nabla )$ -modules over R, we refer to [Reference Kedlaya9, Section 2.5]. We denote by $\operatorname {\mathrm {\textbf {Mod}^{\varphi }_R}}$ (resp. $\operatorname {\mathrm {\textbf {Mod}^{\varphi ,\nabla }_R}}$ ) the category of $\varphi $ -modules (resp. $(\varphi ,\nabla )$ -modules) over R.

Let $(M,\varphi )\in \operatorname {\mathrm {\textbf {Mod}^{\varphi }_R}}$ , and let n be a positive integer. Then, $(M,\varphi ^n)$ is a $\varphi ^n$ -module over $(R,\varphi ^n)$ . The n-pushforward functor is given by

$$ \begin{align*}[n]_*\colon \operatorname{\mathrm{\textbf{Mod}^{\varphi}_R}} \operatorname{\mathrm{\longrightarrow}} \textbf{Mod}^{\varphi^n}_R, \;\;\;\;\; (M,\varphi) \longmapsto (M,\varphi^n).\end{align*} $$

For any $s\in \mathbb {Z}$ , we define the twist $M(s)$ of $(M,\varphi )$ by s to be the $\varphi $ -module $(M,\varpi ^s\Phi )$ .

Now, let M be a $\varphi $ -module over R of rank d.

  1. (i) We say that M is a unit-root $\varphi $ -module if there exists a basis $\textbf {v}_1,\ldots ,\textbf {v}_d$ of M over R in which $\varphi $ acts via an invertible matrix in $\operatorname {\mathrm {GL}}_d(\mathcal {O}_{\mathcal {E}^{\dagger }})$ if $R\in \{\mathcal {E}^{\dagger },\mathcal {R}\}$ , or $\operatorname {\mathrm {GL}}_d(\mathcal {O}_{\operatorname {\mathrm {\tilde {\mathcal {E}}}}^{\dagger }})$ if $R\in \{\operatorname {\mathrm {\tilde {\mathcal {E}}}}^{\dagger },\operatorname {\mathrm {\tilde {\mathcal {R}}}}\}$ .

  2. (ii) Let $\mu =s/r \in \mathbb {Q}$ with $r>0$ and $(s,r)=1$ . We say that M is pure of slope $\mu $ if $([r]_*M)(-s)$ is unit-root.

Let $M\in \operatorname {\mathrm {\textbf {Mod}^{\varphi }_R}}$ . We have a canonical filtration $0=M_0 \subseteq M_1 \subseteq \cdots \subseteq M_l=M$ of $\varphi $ -submodules over R such that each quotient $M_i/M_{i-1}$ is pure of some slope $\mu _i$ with $\mu _1< \cdots <\mu _l$ , by [Reference Kedlaya11, Theorem 1.7.1] if $R=\mathcal {R}$ or [Reference Kedlaya11, Proposition 1.4.15 and Theorem 2.1.8]) if $R=\operatorname {\mathrm {\tilde {\mathcal {R}}}}$ . This is called the slope filtration of M. We call $\mu _1,\ldots ,\mu _l$ the jumps of the slope filtration. The (uniquely determined, not necessarily strictly) increasing sequence $(\mu _1,\ldots ,\mu _1,\ldots ,\mu _l,\ldots ,\mu _l)$ , with each $\mu _i$ appearing $\operatorname {\mathrm {rk}}_R(M_i/M_{i-1})$ times, is said to be the slope sequence for M. We call $\operatorname {\mathrm {rk}}_R (M_i/M_{i-1})$ the multiplicity of $\mu _i$ for all $1\leq i\leq l$ .

Moreover, if M is a $(\varphi ,\nabla )$ -module over $\mathcal {R}$ , then the slope filtration can be refined to a filtration of $(\varphi ,\nabla )$ -submodules. This is [Reference Kedlaya9, Theorem 6.12], and is referred to as the slope filtration theorem for $(\varphi ,\nabla )$ -modules over $\mathcal {R}$ .

To continue, we need to recall some notions introduced in [Reference Kedlaya12, Section 14]. A difference ring (resp. difference field) is a ring (resp. field) R equipped with an endomorphism $\phi $ . A difference module over R is an R-module M equipped with a $\phi $ -linear endomorphism $\Phi $ . A finite free difference module M over R is said to be dualizable (resp. trivial) if M admits a basis over R such that $\Phi $ acts via an invertible matrix (resp. the identity matrix). For example, a $\varphi $ -module over R is a dualizable difference module over R where R is any of the rings constructed in Section 2.2. A dualizable difference module M over R is said to be standard if it admits a basis $e_1,\ldots ,e_d$ such that $e_i=\Phi (e_{i-1})$ for $2\leq i \leq d$ and $\Phi (e_d)=\lambda e_1$ for some $\lambda \in R^{\times }$ . A difference field $(k,\phi _k)$ is called strongly difference-closed if $\phi _k$ is an automorphism and any dualizable difference module over k is trivial.

Let k be a complete nonarchimedean valued field and $(k,\phi _k)$ is a difference field in which $\phi _k$ is bijective. An admissible extension of $(k,\phi _k)$ is a difference field $(\ell ,\phi _\ell )$ , where $\ell $ is a field extension of k complete for the valuation extending the one on k with the same value group, and $\phi _\ell $ is an automorphism of $\ell $ extending $\phi _k$ . (See [Reference Kedlaya11, Definition 3.2.1].)

Lemma 2.2 [Reference Liu15, Lemma 1.5.3] The field K admits an admissible extension E such that the residue field $\kappa _E$ of E is strongly difference-closed.

The following lemma is a recollection of some results which will be used in the sequel.

Lemma 2.3 Let E be an admissible extension of K such that $\kappa _E$ is strongly difference-closed.

  1. (i) Let $M \in \operatorname {\mathrm {\textbf {Mod}^{\varphi }_{\mathcal {R}}}}$ . Then, tensoring the slope filtration of M with $\tilde {\mathcal {R}}(E,t)$ gives the slope filtration of $M\bigotimes _{\mathcal {R}}\tilde {\mathcal {R}}(E,t)$ .

  2. (ii) Let $0 \operatorname {\mathrm {\longrightarrow }} M_1 \operatorname {\mathrm {\longrightarrow }} M \operatorname {\mathrm {\longrightarrow }} M_2 \operatorname {\mathrm {\longrightarrow }} 0$ be a short exact sequence of $\varphi $ -modules over $\tilde {\mathcal {R}}(E,t)$ such that the slopes of $M_1$ are all less than the smallest slope of $M_2$ . Then, the sequence splits.

  3. (iii) Every $\varphi $ -module over $\tilde {\mathcal {R}}(E,t)$ admits a Dieudonné–Manin decomposition, i.e., it is a direct sum of standard $\varphi $ -submodules.

  4. (iv) Let M and N be $\varphi $ -modules over $\tilde {\mathcal {R}}(E,t)$ . If the slopes of M are all less than the smallest slope of N, then no nonzero morphism from M to N exists.

Assertion (i) is [Reference Liu15, Proposition 1.5.6]. Assertion (ii) is [Reference Liu15, Proposition 1.5.11], and assertion (iii) is Proposition 1.5.12 in loc. cit. Assertion (iv) is [Reference Kedlaya11, Proposition 1.4.18].

2.4 The Tannakian duality

In this subsection, k denotes a field unless otherwise specified. We follow the definitions and notations in [Reference Deligne and Milne7]. We denote by $\omega ^G$ the forgetful functor $\operatorname {\mathrm {\textbf {Rep}}}_k(G)\to \operatorname {\mathrm {\textbf {Vec}}}_k$ , which is called the fiber functor.

The following Tannakian duality will be repeatedly used in this paper, whose proof can be found, e.g., in [Reference Milne18, Theorem 9.2].

Theorem 2.4 Let G be an affine algebraic k-group, and let $R\in \operatorname {\mathrm {\textbf {Alg}}}_k$ . Suppose that for any $(V,\rho _V) \in \operatorname {\mathrm {\textbf {Rep}}}_k(G)$ , we are given an R-linear map $\lambda _V\colon V_R\to V_R$ . If the family $\{\lambda _V\mid (V,\rho _V)\in \operatorname {\mathrm {\textbf {Rep}}}_k(G)\}$ satisfies

  1. (i) $\lambda _{V\bigotimes W}=\lambda _V\otimes \lambda _W$ for all $V,W\in \operatorname {\mathrm {\textbf {Rep}}}_k(G)$ ;

  2. (ii) is the identity map where is the trivial representation on k;

  3. (iii) for all G-equivariant maps $\alpha \colon V\to W$ , we have $\lambda _W\circ \alpha _R=\alpha _R\circ \lambda _V$ .

Then, there exists a unique $g\in G(R)$ such that $\lambda _V=\rho _V(g)$ for all $(V,\rho _V) \in \operatorname {\mathrm {\textbf {Rep}}}_k(G)$ .

Corollary 2.5 Let G be an affine algebraic k-group. We have an isomorphism $G \cong \operatorname {\mathrm {\underline {Aut}}}^{\otimes }(\omega ^G)$ of affine algebraic k-groups.

Corollary 2.6 Let G be a smooth affine algebraic k-group. Let $\ell /k$ be a field extension, and let $\eta \colon \operatorname {\mathrm {\textbf {Rep}}}_\ell (G)\to \operatorname {\mathrm {\textbf {Vec}}}_\ell $ be a fiber functor over $\ell $ . Then, $\operatorname {\mathrm {\underline {Hom}}}^{\otimes } (\omega ^G,\eta )$ is a G-torsor over $\ell $ . In particular, if $H^1(\ell ,G)=\{1\}$ and $G(\ell )\neq \emptyset $ , then $\omega ^G$ is isomorphic to $\eta $ over $\ell $ .

Proof Notice that we have an action

$$ \begin{align*}\operatorname{\mathrm{\underline{Hom}}}^{\otimes} (\omega^G,\eta) \times \operatorname{\mathrm{\underline{Aut}}}^{\otimes} (\omega^G) \operatorname{\mathrm{\longrightarrow}} \operatorname{\mathrm{\underline{Hom}}}^{\otimes} (\omega^G,\eta)\end{align*} $$

by precomposition. By [Reference Deligne and Milne7, Theorem 3.2(i)], $\operatorname {\mathrm {\underline {Hom}}}^{\otimes } (\omega ^G,\eta )$ is an $\operatorname {\mathrm {\underline {Aut}}}^{\otimes } (\omega ^G)$ -torsor. In particular, it is a G-torsor over $\ell $ by Corollary 2.5.

Because G is a $\ell $ -group variety, G-torsors over $\ell $ are $\ell $ -varieties by [Reference Milne18, Proposition 2.69], whose isomorphism classes are classified by $H^1(\ell ,G)$ . It follows from the triviality of $H^1(\ell ,G)$ that $\operatorname {\mathrm {\underline {Hom}}}^{\otimes } (\omega ^G,\eta )(\ell )\cong G(\ell )$ ; hence, $\operatorname {\mathrm {\underline {Hom}}}^{\otimes } (\omega ^G,\eta )(\ell )\neq \emptyset $ . [Reference Deligne and Milne7, Proposition 1.13] then implies the second assertion.▪

To end this subsection, we give a Lie algebra version of Theorem 2.4. We start with recalling the notion of the Lie algebra of a k-group functor. (See [Reference Demazure and Gabriel8, Chapitre II, Section 4] for more details. Notice that k denotes a ring in loc. cit.)

For any $R\in \operatorname {\mathrm {\textbf {Alg}}}_k$ , we define the R-algebra of dual numbers $R[\varepsilon ]:= R[X]/(X^2)$ . Put $\varepsilon := X+(X^2)$ ; we then have the canonical projection $\pi _R\colon R[\varepsilon ]\to R,~\varepsilon \mapsto 0$ . Let G be a k-group functor. We define

$$ \begin{align*}\operatorname{\mathrm{Lie}}(G)(R) := \operatorname{\mathrm{Ker}} G(\pi_R).\end{align*} $$

Let $f\colon G\to H$ be a morphism of k-group functors. The commutative diagram

(1)

implies that $f(R[\epsilon ])\circ \iota _G(X)\in \operatorname {\mathrm {Lie}}(H)(R)$ for all $X\in \operatorname {\mathrm {Lie}}(G)(R)$ . We define $\operatorname {\mathrm {Lie}}(f):= f(R[\epsilon ])\circ \iota _G\colon \operatorname {\mathrm {Lie}}(G)(R) \to \operatorname {\mathrm {Lie}}(H)(R)$ . Hence, $\operatorname {\mathrm {Lie}}(\rule {2mm}{0.15mm})(R)$ is a functor from the category of k-group functors to that of abelian groups.

For an affine algebraic k-group G, we write I for the kernel of the counit $\epsilon _G\colon k[G]\to k$ . Because $k[G]$ is Noetherian, $I/I^2$ is a finite-dimensional vector space over $k\cong k[G]/I$ . We then have $\operatorname {\mathrm {Hom}}_k(I/I^2,R) \cong \operatorname {\mathrm {Hom}}_k(I/I^2,k) \bigotimes _k R$ . By [Reference Demazure and Gabriel8, Corollaire II.3.3], we have canonical group isomorphisms $\operatorname {\mathrm {Lie}}(G)(R)\cong \operatorname {\mathrm {Hom}}_k(I/I^2,R)$ and $\mathfrak {g}=\operatorname {\mathrm {Lie}}(G)(k) \cong \operatorname {\mathrm {Hom}}_k(I/I^2,k)$ , whence $\operatorname {\mathrm {Lie}}(G)(R) \cong \mathfrak {g}_R$ . The Lie structure on $\mathfrak {g}$ then canonically gives a Lie structure on $\mathfrak {g}_R$ and hence on $\operatorname {\mathrm {Lie}}(G)(R)$ . We call $\operatorname {\mathrm {Lie}}(G)(R)$ the Lie algebra of G over R, and will identify it with $\mathfrak {g}_R$ . Moreover, $\operatorname {\mathrm {Lie}}(\rule {2mm}{0.15mm})(R)$ is a functor from the category of affine algebraic k-groups to that of Lie algebras over R.

Remark 2.7 More generally, let k be a commutative ring with $1$ and let G be a smooth k-group scheme. For any k-algebra R, we can similarly define $\operatorname {\mathrm {Lie}}(G)(R)$ as above. Because the $\mathscr {O}_G$ -module $\Omega _{G/k}^1$ is finite locally free, we have $\operatorname {\mathrm {Lie}}(G)(R)\cong \operatorname {\mathrm {Lie}}(G)(k)\bigotimes _k R$ by [Reference Demazure and Gabriel8, Proposition II.4.8].

Remark 2.8 For any d-dimensional G-representation $(V,\rho _V)$ , we write $\operatorname {\mathrm {\mathfrak {gl}}}_V:= \operatorname {\mathrm {Lie}}(\operatorname {\mathrm {GL}}_V)(k)$ . We then have $\operatorname {\mathrm {\mathfrak {gl}}}_{V,R}=\{I_d+\varepsilon B \mid B\in \operatorname {\mathrm {Mat}}_{d,d}(R)\}$ , after choosing a k-basis for V. Then, $I_d+\varepsilon B \mapsto B$ gives a group isomorphism from $\operatorname {\mathrm {\mathfrak {gl}}}_{V,R}$ to $\operatorname {\mathrm {End}}_R(V_R)$ . Henceforth, we will identify $\operatorname {\mathrm {Lie}}(\rho _V)(X)$ as an endomorphism of $V_R$ , for all $X\in \mathfrak {g}_R$ .

Replacing H with $\operatorname {\mathrm {GL}}_V$ and f with $\rho _V$ in diagram (1), we obtain a morphism $\operatorname {\mathrm {Lie}}(\rho _V)=\rho _V(R[\epsilon ])\circ \iota _G \colon \mathfrak {g}_R \to \operatorname {\mathrm {\mathfrak {gl}}}_{V,R}$ of Lie algebras over R. Let $(W,\rho _W)\in \operatorname {\mathrm {\textbf {Rep}}}_k(G)$ , and let $\alpha \in \operatorname {\mathrm {Hom}}_G(V,W)$ . We then have $\alpha _R\circ \operatorname {\mathrm {Lie}}(\rho _V)(X)=\operatorname {\mathrm {Lie}}(\rho _W)(X)\circ \alpha _R$ for all $X\in \mathfrak {g}_R$ .

Applying the functor $\operatorname {\mathrm {Lie}}(\rule {2mm}{0.15mm})(R)$ on both sides of the isomorphism in Corollary 2.5 gives us an isomorphism $\mathfrak {g}_R\cong \operatorname {\mathrm {Lie}}(\operatorname {\mathrm {\underline {Aut}}}^{\otimes } (\omega ^G))(R)$ of Lie algebras over R. The following corollary indicates that elements in $\operatorname {\mathrm {Lie}}(\operatorname {\mathrm {\underline {Aut}}}^{\otimes } (\omega ^G))(R)$ are exactly the derivatives (in the sense of taking derivations of conditions (i–iii) in Theorem 2.4) of elements in $\operatorname {\mathrm {\underline {Aut}}}^{\otimes } (\omega ^G)(R)$ .

Corollary 2.9 Let G be an affine algebraic k-group, and let R be a k-algebra. Suppose that for any $(V,\rho _V)\in \operatorname {\mathrm {\textbf {Rep}}}_k (G)$ , we are given an R-linear endomorphism $\theta _V$ of $V_R$ subject to the conditions:

  1. (i) $\theta _{V\bigotimes W}=\theta _V\otimes \operatorname {\mathrm {Id}}_{W_R} +\operatorname {\mathrm {Id}}_{V_R}\otimes \theta _W$ for all $V,W\in \operatorname {\mathrm {\textbf {Rep}}}_k(G)$ ;

  2. (ii) , where is the trivial G-representation;

  3. (iii) $\theta _W\circ \alpha _R=\alpha _R\circ \theta _V$ for all $\alpha \in \operatorname {\mathrm {Hom}}_G(V,W)$ .

Then, there exists a unique element $X\in \mathfrak {g}_R$ such that $\theta _V=\operatorname {\mathrm {Lie}}(\rho _V)(X)$ for all $(V,\rho _V)\in \operatorname {\mathrm {\textbf {Rep}}}_k(G)$ .

Proof For any $(V,\rho _V)\in \operatorname {\mathrm {\textbf {Rep}}}_k(G)$ and $\theta _V\colon V_R\to V_R$ , we define an $R[\varepsilon ]$ -linear map

$$ \begin{align*}\varepsilon \theta_V \colon V_{R[\varepsilon]} \operatorname{\mathrm{\longrightarrow}} V_{R[\varepsilon]}, ~~~~~ v\otimes(x+y\varepsilon) \longmapsto \theta_V(v\otimes x)\otimes \varepsilon.\end{align*} $$

We define an $R[\varepsilon ]$ -linear endomorphism

$$ \begin{align*}\tilde \theta_V := \operatorname{\mathrm{Id}}_{V_{R[\varepsilon]}}+\varepsilon \theta_V \colon V_{R[\varepsilon]} \operatorname{\mathrm{\longrightarrow}} V_{R[\varepsilon]}. \end{align*} $$

Then, $\tilde \theta _V\in \operatorname {\mathrm {Lie}}(\operatorname {\mathrm {GL}}_V)(R)\subseteq \operatorname {\mathrm {GL}}_V(R[\varepsilon ])$ , because $\pi _R(\tilde \theta _V)=\operatorname {\mathrm {Id}}_{V_R}$ .

We claim that the family

(2) $$ \begin{align} \big\{\tilde \theta_V\colon V_{R[\varepsilon]}\to V_{R[\varepsilon]} \mid (V,\rho_V)\in \operatorname{\mathrm{\textbf{Rep}}}_k (G) \big\} \end{align} $$

of $R[\varepsilon ]$ -linear endomorphisms satisfies conditions (i–iii) in Theorem 2.4. Granting this claim for a moment, we conclude that $\tilde \theta \in \operatorname {\mathrm {\underline {Aut}}}^{\otimes }(\omega ^G)(R[\varepsilon ])$ . In particular, there exists a unique element $X\in G(R[\varepsilon ])$ such that $\tilde \theta _V=\rho _V(X)$ for all $(V,\rho _V)\in \operatorname {\mathrm {\textbf {Rep}}}_k(G)$ . Because $\pi _R(\tilde \theta )=\operatorname {\mathrm {Id}}\in \operatorname {\mathrm {\underline {Aut}}}^{\otimes }(\omega ^G)(R)$ , we have $\tilde \theta \in \operatorname {\mathrm {Lie}}(\operatorname {\mathrm {\underline {Aut}}}^{\otimes }(\omega ^G))(R)$ . The isomorphism $\mathfrak {g}_R\cong \operatorname {\mathrm {Lie}}(\operatorname {\mathrm {\underline {Aut}}}^{\otimes } (\omega ^G))(R)$ then implies that $X\in \mathfrak {g}_R$ . Furthermore, it follows from the construction that $\theta _V=\operatorname {\mathrm {Lie}}(\rho _V)(X)$ for all $(V,\rho _V)\in \operatorname {\mathrm {\textbf {Rep}}}_k(G)$ , and the corollary follows.

It remains to prove the claim. Condition (ii) is clear from the construction. Given $(W,\rho _W)\in \operatorname {\mathrm {\textbf {Rep}}}_k(G)$ , we compute

$$ \begin{align*} \tilde \theta_{V\bigotimes W} &=\operatorname{\mathrm{Id}}_{(V\bigotimes W)_R}+\varepsilon \theta_{V\bigotimes W}\\ &=\operatorname{\mathrm{Id}}_{(V\otimes W)_R}+\varepsilon(\theta_V\otimes \operatorname{\mathrm{Id}}_{W_R} +\operatorname{\mathrm{Id}}_{V_R}\otimes \theta_W)\\ &=(\operatorname{\mathrm{Id}}_{V_R}+\varepsilon \theta_V)\otimes (\operatorname{\mathrm{Id}}_{W_R}+\varepsilon \theta_W)\\ &=\tilde \theta_V\otimes \tilde \theta_W. \end{align*} $$

Hence, (2) satisfies condition (i). It remains to show that Theorem 2.4 satisfies condition (iii). Let $\alpha \in \operatorname {\mathrm {Hom}}_G(V,W)$ . For any $v\otimes (x+y\varepsilon )\in V_{R[\varepsilon ]}$ , we compute

$$ \begin{align*} \alpha_{R[\varepsilon]}\circ \varepsilon \theta_V(v\otimes(x+y\varepsilon)) &=\alpha_{R[\varepsilon]} (\theta_V(v \otimes x)\otimes \varepsilon) =(\alpha_R\circ \theta_V)(v\otimes x )\otimes \varepsilon \\ &=(\theta_W\circ \alpha_R)(v\otimes x ) \otimes \varepsilon =\theta_W (\alpha(v)\otimes x)\otimes \varepsilon \\ &=\varepsilon\theta_W ( \alpha(v)\otimes (x+y\epsilon)) =\varepsilon\theta_W \circ \alpha_{R[\varepsilon]} (v\otimes (x+y\varepsilon)). \end{align*} $$

It follows that

$$ \begin{align*} \alpha_{R[\varepsilon]}\circ \tilde \theta_V&=\alpha_{R[\varepsilon]}\circ (\operatorname{\mathrm{Id}}_{V_{R[\varepsilon]}}+\varepsilon \theta_V) =\alpha_{R[\varepsilon]}+\alpha_{R[\varepsilon]}\circ \varepsilon \theta_V\\ &=\alpha_{R[\varepsilon]}+\varepsilon\theta_W \circ \alpha_{R[\varepsilon]} =(\operatorname{\mathrm{Id}}_{W_{R[\varepsilon]}}+\varepsilon \theta_W)\circ \alpha_{R[\varepsilon]}\\ &=\tilde \theta_W \circ \alpha_{R[\varepsilon]}, \end{align*} $$

as desired.▪

2.5 Filtered and graded fiber functors

We recall the notions of filtered and graded fiber functors on Tannakian categories following [Reference Ziegler25]. Let $\Gamma $ be a totally ordered abelian group (written additively), and let $R\in \operatorname {\mathrm {\textbf {Alg}}}_k$ . A $\Gamma $ -graded R-module is an R-module M together with a direct sum decomposition $M=\bigoplus \limits _{\gamma \in \Gamma } M_\gamma $ . A morphism between two $\Gamma $ -graded R modules M and N is an R-linear map $f\colon M\to N$ such that $f(M_\gamma )\subseteq N_\gamma $ for all $\gamma \in \Gamma $ . We denote by $\Gamma $ - $\operatorname {\mathrm {\textbf {Grad}}}_R$ the category of $\Gamma $ -graded modules over R. For $M,N\in \Gamma $ - $\operatorname {\mathrm {\textbf {Grad}}}_R$ , we define the tensor product $(M\bigotimes _R N)_\gamma = \bigoplus \limits _{\gamma '+\gamma ''=\gamma } \big ( M_{\gamma '}\bigotimes _R N_{\gamma ''}\big )$ .

Let M be an R-module. A $\Gamma $ -filtration on M is an increasing map

$$ \begin{align*}\mathcal{F}\colon \Gamma \operatorname{\mathrm{\longrightarrow}} \{R\text{-submodules of}~ M\},~~~\gamma \longmapsto \mathcal{F}^{\gamma} M ,\end{align*} $$

such that $\mathcal {F}^{\gamma } M=0$ for $\gamma \ll 0$ and $\mathcal {F}^{\gamma } M=M$ for $\gamma \gg 0$ , which is increasing in the sense that $\mathcal {F}^{\gamma } M \subseteq \mathcal {F}^{\gamma '} M$ whenever $\gamma \leq \gamma '$ . A $\Gamma $ -filtered R-module is an R-module M with a $\Gamma $ -filtration. To abbreviate notations, we sometimes denote $\mathcal {F}^{\gamma } M$ by $M^{\gamma }$ if no confusion shall arise. A morphism between two $\Gamma $ -filtered R-modules M and N is an R-linear map $f\colon M\to N$ such that $f(M^{\gamma })\subseteq N^{\gamma }$ for all $\gamma \in \Gamma $ . We denote by $\Gamma $ - $\operatorname {\mathrm {\textbf {Fil}}}_R$ the category of $\Gamma $ -filtered modules over R.

Let M be a $\Gamma $ -filtered module over R. For any $\gamma \in \Gamma $ , we put $\mathcal {F}^{\gamma -} M:= \operatorname *{\mathrm {\sum }}\limits _{\gamma '<\gamma } \mathcal {F}^{\gamma '} M$ . We define

$$ \begin{align*}\operatorname{\mathrm{gr}}_{\mathcal{F}}^{\gamma} M:= \mathcal{F}^{\gamma} M/\mathcal{F}^{\gamma-}M.\end{align*} $$

Then, $\operatorname {\mathrm {gr}}_{\mathcal {F}} M:= \bigoplus \limits _{\gamma \in \Gamma } \operatorname {\mathrm {gr}}_{\mathcal {F}}^{\gamma } M$ is a $\Gamma $ -graded R module, and is called the $\Gamma $ -graded R-module associated to $\mathcal {F}$ . We thus have a functor

$$ \begin{align*}\operatorname{\mathrm{gr}}\colon \Gamma\text{-}\operatorname{\mathrm{\textbf{Fil}}}_R \operatorname{\mathrm{\longrightarrow}} \Gamma\text{-}\operatorname{\mathrm{\textbf{Grad}}}_R.\end{align*} $$

Elements $\gamma \in \Gamma $ such that $\operatorname {\mathrm {gr}}_{\mathcal {F}}^{\gamma } M\neq 0$ are said to be the $\Gamma $ -jumps (or simply jumps) of $\mathcal {F}$ .

The tensor product structure in $\Gamma $ - $\operatorname {\mathrm {\textbf {Fil}}}_R$ is defined by

$$ \begin{align*}\mathcal{F}^{\gamma} (M\bigotimes_R N)=\operatorname*{\mathrm{\sum}}\limits_{\gamma'+\gamma''=\gamma} \mathcal{F}^{\gamma'} M\bigotimes_R \mathcal{F}^{\gamma''}N,\end{align*} $$

for all $\Gamma $ -filtered modules M and N over R.

A morphism $f\colon M\to N$ in $\Gamma $ - $\operatorname {\mathrm {\textbf {Fil}}}_R$ is said to be admissible (or strict) if

$$ \begin{align*}f(M^{\gamma})=f(M)\cap N^{\gamma},~~~~~\forall \gamma\in \Gamma.\end{align*} $$

Following [Reference Ziegler25, Section 4.1], we say that a short sequence in $\Gamma $ - $\operatorname {\mathrm {\textbf {Fil}}}_R$ is exact if both of $f'$ and $f''$ are admissible, and the underlying short sequence in $\operatorname {\mathrm {\textbf {Mod}}}_R$ is exact.

Let $\mathcal T$ be a Tannakian category over k, and let R be a k-algebra.

  1. (i) A $\Gamma $ -graded fiber functor on $\mathcal T$ over R is an exact faithful k-linear tensor functor $\tau \colon \mathcal T \to \Gamma $ - $\operatorname {\mathrm {\textbf {Grad}}}_R$ .

  2. (ii) A $\Gamma $ -filtered fiber functor on $\mathcal T$ over R is an exact faithful k-linear tensor functor $\eta \colon \mathcal T \to \Gamma $ - $\operatorname {\mathrm {\textbf {Fil}}}_R$ .

  3. (iii) Given an object $M=\bigoplus \limits _{\gamma \in \Gamma } M_\gamma $ in $\Gamma $ - $\operatorname {\mathrm {\textbf {Grad}}}_R$ , we put $\mathcal {F}^{\gamma }(M):=\bigoplus \limits _{\gamma '\leq \gamma } M_{\gamma '}$ . This gives rise to a functor $\operatorname {\mathrm {fil}}\colon \Gamma $ - $\operatorname {\mathrm {\textbf {Grad}}}_R \to \Gamma $ - $\operatorname {\mathrm {\textbf {Fil}}}_R$ .

  4. (iv) A $\Gamma $ -filtered fiber functor $\eta $ is called splittable if there exists a $\Gamma $ -graded fiber functor $\tau $ such that $\eta =\operatorname {\mathrm {fil}}\circ \tau $ , and $\tau $ is called a splitting of $\eta $ .

Remark 2.10 More concretely, a $\Gamma $ -filtered fiber functor is a k-linear functor $\eta \colon \mathcal T \to \Gamma $ - $\operatorname {\mathrm {\textbf {Fil}}}_R$ satisfying the following properties (cf. [Reference Dat, Orlik and Rapoport6, Definition 4.2.6 and Remark 4.2.7]).

  1. (i) It is admissibly (or strictly) functorial, i.e., for any morphism $\alpha \colon X\to Y$ in $\mathcal T$ , we have $\eta (\alpha )\big (\mathcal {F}^{\gamma } \eta (X)\big )=\eta (\alpha )(\eta (X))\cap \mathcal {F}^{\gamma } \eta (Y)$ for all $\gamma \in \Gamma $ .

  2. (ii) It is compatible with tensor products, i.e., we have

    $$ \begin{align*} \mathcal{F}^{\gamma} \big(\eta(X\bigotimes Y)\big)=\operatorname*{\mathrm{\sum}}\limits_{\gamma'+\gamma''=\gamma} \mathcal{F}^{\gamma'} \big(\eta(X)\big)\bigotimes \mathcal{F}^{\gamma''} \big(\eta(Y)\big),\end{align*} $$
    for all $X,Y\in \operatorname {\mathrm {Ob}}(\mathcal T)$ and $\gamma \in \Gamma $ .
  3. (iii)

    where is the identity object in $\mathcal T$ . Note that is the identity object in $\Gamma $ - $\operatorname {\mathrm {\textbf {Fil}}}_R$ .

Construction 2.11 Let $(M,\mathcal {F})\in \mathbb {Z}$ - $\operatorname {\mathrm {\textbf {Fil}}}_R$ be a $\mathbb {Z}$ -filtered module with $\mathbb {Z}$ -jumps $\jmath _1< \cdots <\jmath _n$ . For any $\gamma \in \mathbb {Q}_{>0}$ , we define a $\mathbb {Q}$ -filtered module $(M,[\gamma ]_*\mathcal {F})$ by

$$ \begin{align*}([\gamma]_*\mathcal{F})^{x} M:= \left\{ \begin{array}{ll} 0 &\text{for}~x< \jmath_1\gamma, \\ M^{\jmath_i} ~~~ &\text{for}~ \jmath_i\gamma \leq x <\jmath_{i+1}\gamma,~ 1\leq i\leq n-1,\\ M &\text{for}~x\geq \jmath_n\gamma. \end{array} \right.\end{align*} $$

We then have a fully faithful embedding $[\gamma ]_*\colon \mathbb {Z}$ - $\operatorname {\mathrm {\textbf {Fil}}}_R \to \mathbb {Q}$ - $\operatorname {\mathrm {\textbf {Fil}}}_R$ . Similarly, we have a fully faithful embedding $[\gamma ]_*\colon \mathbb {Z}$ - $\operatorname {\mathrm {\textbf {Grad}}}_R \to \mathbb {Q}$ - $\operatorname {\mathrm {\textbf {Grad}}}_R$ by defining $[\gamma ]_* := \operatorname {\mathrm {gr}}\circ [\gamma ]_*\circ \operatorname {\mathrm {fil}}$ .

To end this subsection, we exhibit the following theorem for latter use. (Be aware that in [Reference Ziegler25], the author only considers $\Gamma $ -gradings and $\Gamma $ -filtrations for $\Gamma = \mathbb{Z}$ .)

Theorem 2.12 [Reference Ziegler25, Theorem 4.15] Let $\mathcal T$ be a Tannakian category over a field k, and let R be a k-algebra. Let $\eta \colon \mathcal T \to \mathbb {Z}$ - $\operatorname {\mathrm {\textbf {Fil}}}_R$ be a $\mathbb {Z}$ -filtered fiber functor. If $\operatorname {\mathrm {\underline {Aut}}}^{\otimes }_R(\operatorname {\mathrm {forg}}\circ \eta )$ is prosmooth (i.e., a limit of smooth algebraic group schemes) over R, where $\operatorname {\mathrm {forg}}\colon \mathbb {Z}$ - $\operatorname {\mathrm {\textbf {Fil}}}_R \to \operatorname {\mathrm {\textbf {Mod}}}_R$ is the forgetful functor, then $\eta $ is splittable.

3 G- $\varphi $ -modules over the Robba ring

We fix an affine algebraic F-group G in this section.

3.1 Definition

Let $R\in \{\mathcal {E}^{\dagger },\mathcal {R},\operatorname {\mathrm {\tilde {\mathcal {E}}}}^{\dagger },\operatorname {\mathrm {\tilde {\mathcal {R}}}}\}$ equipped with an absolute Frobenius lift $\varphi $ . The following definition is motivated by that of G-isocrystals introduced in [Reference Dat, Orlik and Rapoport6, Section IX.1].

Definition 3.1 A G- $\varphi $ -module over R is an exact faithful F-linear tensor functor

$$ \begin{align*}\operatorname{\mathrm{I}}\colon \operatorname{\mathrm{\textbf{Rep}}}_F(G) \operatorname{\mathrm{\longrightarrow}} \operatorname{\mathrm{\textbf{Mod}^{\varphi}_R}},\end{align*} $$

which satisfies $\operatorname {\mathrm {forg}} \circ \operatorname {\mathrm {I}} = \omega ^G \otimes R$ , where $\operatorname {\mathrm {forg}} \colon \operatorname {\mathrm {\textbf {Mod}^{\varphi }_R}} \to \operatorname {\mathrm {\textbf {Mod}}}_R$ is the forgetful functor. The category of G- $\varphi $ -modules over R is denoted by $\operatorname {\mathrm {G-\textbf {Mod}^{\varphi }_R}}$ , whose morphisms are morphisms of tensor functors.

Let $(V,\rho )\in \operatorname {\mathrm {\textbf {Rep}}}_F(G)$ , and let $g\in G(R)$ . We define $\operatorname {\mathrm {I}}(g)(V):= (V_R,g\varphi )$ , where

$$ \begin{align*} g\varphi\colon V_R \operatorname{\mathrm{\longrightarrow}} V_R, \;\;\;\;\; v\otimes f \longmapsto \rho(g)(v\otimes 1)\varphi(f). \end{align*} $$

Let $V,W\in \operatorname {\mathrm {\textbf {Rep}}}_F(G)$ . We have a canonical isomorphism $(V\otimes W)_{\mathcal {R}} \cong V_{\mathcal {R}} \otimes _{\mathcal {R}} W_{\mathcal {R}}$ , and we will henceforth identify them. Given any $\alpha \in \operatorname {\mathrm {Hom}}_G(V,W)$ , we define $\operatorname {\mathrm {I}}(g)(\alpha ):= \alpha _R$ . We thus have the following G- $\varphi $ -module over R (associated to g).

$$ \begin{align*}\operatorname{\mathrm{I}}(g)\colon \operatorname{\mathrm{\textbf{Rep}}}_F(G) \operatorname{\mathrm{\longrightarrow}} \operatorname{\mathrm{\textbf{Mod}^{\varphi}_R}},\;\;\;\;\; V\longmapsto (V_R,g\varphi).\end{align*} $$

We call $\operatorname {\mathrm {I}}(g)(V)=(V_R,g\varphi )$ the G- $\varphi $ -module over R associated to g.

For any $g\in G(R)$ , we sometimes write $\Phi _g= \Phi _{g,V}$ for the $\varphi $ -linear action $g\varphi $ on $V_R$ . Both notations have their own advantages in practice.

Remark 3.2 For any $g\in G(R)$ , we define $\Phi (g):= G(\varphi )(g)$ . For any $(V,\rho )\in \operatorname {\mathrm {\textbf {Rep}}}_F(G)$ , we have a commutative diagram

Hence, $\rho (\varphi (g))=\varphi (\rho (g))$ . For any $h\in G(R)$ and $n,m\geq 0$ , we have the following formula in $G(R)\rtimes \operatorname {\mathrm {\langle }}\varphi \operatorname {\mathrm {\rangle }}$ :

$$ \begin{align*}(h\varphi^n)\circ (g\varphi^m)=\big(h\varphi^n(g)\big)\varphi^{n+m}. \end{align*} $$

3.2 The $\mathbb {Q}$ -filtered fiber functor $\operatorname {\mathrm {{HN}}}_g$

We fix an element $g\in G(\mathcal {R})$ .

Construction 3.3 For any $V\in \operatorname {\mathrm {\textbf {Rep}}}_F(G)$ , we have a $\varphi $ -module $(V_{\mathcal {R}},g\varphi )$ over $\mathcal {R}$ . Kedlaya’s slope filtration theorem [Reference Kedlaya9, Theorem 6.10] then provides a filtration

$$ \begin{align*}0\subseteq V_{\mathcal{R}}^{\mu_1}\subseteq \cdots\subseteq V_{\mathcal{R}}^{\mu_l}=V_{\mathcal{R}},\end{align*} $$

satisfying

  • $V_{\mathcal {R}}^{\mu _1}$ is pure of some slope $\mu _1\in \mathbb {Q}$ and each $V_{\mathcal {R}}^{\mu _i}/V_{\mathcal {R}}^{\mu _{i-1}}$ is pure of some slope $\mu _i\in \mathbb {Q}$ for $2\leq i\leq l$ ;

  • $\mu _1<\cdots <\mu _l$ .

We thus have an increasing map

$$ \begin{align*} \mathcal{HN}_g \colon \mathbb{Q} &\operatorname{\mathrm{\longrightarrow}} \{\mathcal{R}\text{-modules of}~V_{\mathcal{R}}\}\\ x &\longmapsto \mathcal{HN}_g^x(V_{\mathcal{R}}), \end{align*} $$

where

$$ \begin{align*}\mathcal{HN}_g^x(V_{\mathcal{R}})=\left\{ \begin{array}{ll} 0 &\text{for}~x<\mu_1,\\ V_{\mathcal{R}}^{\mu_i} ~~~ &\text{for}~ \mu_i\leq x <\mu_{i+1}, 1\leq i\leq l-1,\\ V_{\mathcal{R}} &\text{for}~x\geq \mu_l. \end{array} \right.\end{align*} $$

Then, $(V_{\mathcal {R}},\mathcal {HN}_g)$ is a $\mathbb {Q}$ -filtered module over $\mathcal {R}$ with $\mathbb {Q}$ -jumps $\mu _1<\cdots <\mu _l$ . We will denote $\mathcal {HN}_g^x(V_{\mathcal {R}})$ by $V_{\mathcal {R}}^x$ when $\mathcal {HN}_g$ is clear in the context.

Theorem 3.4 The assignments

$$ \begin{align*}V \longmapsto (V_{\mathcal{R}},\mathcal{HN}_g) ~~~\text{and}~~~\alpha \longmapsto \alpha_{\mathcal{R}},\end{align*} $$

for all $\alpha \in \operatorname {\mathrm {Hom}}_G(V,W)$ , define a $\mathbb {Q}$ -filtered fiber functor

$$ \begin{align*}\operatorname{\mathrm{{HN}}}_g \colon \operatorname{\mathrm{\textbf{Rep}}}_F(G) \operatorname{\mathrm{\longrightarrow}} \mathbb{Q}\text{-}\operatorname{\mathrm{\textbf{Fil}}}_{\mathcal{R}}.\end{align*} $$

Proof This is Propositions 3.5 and 3.6 below.▪

For any admissible extension E of K, we first remark that the $\varphi $ -equivariant embedding $\psi \colon \mathcal {R}\to \operatorname {\mathrm {\tilde {\mathcal {R}}}}(E,t)$ is faithfully flat (see [Reference Kedlaya11, Remark 3.5.3]). We also remark that, if $M_1$ and $M_2$ are pure $\varphi $ -modules over $\mathcal {R}$ of slopes $\mu _1$ and $\mu _2$ , respectively, then $M_1 \bigotimes _{\mathcal {R}} M_2$ is pure of slope $\mu _1+\mu _2$ (cf. [Reference Kedlaya11, Corollary 1.6.4]). These facts will be repeatedly used in the sequel.

Proposition 3.5 The assignments in Theorem 3.4 yield a faithful F-linear tensor functor $\operatorname {\mathrm {{HN}}}_g \colon \operatorname {\mathrm {\textbf {Rep}}}_F(G) \to \mathbb {Q}$ - $\operatorname {\mathrm {\textbf {Fil}}}_{\mathcal {R}}$ .

Proof Let be the trivial G-representation. Then, is of rank $1$ with slope $0$ , proving that $\operatorname {\mathrm {{HN}}}_g$ preserves identity objects.

We claim that $\operatorname {\mathrm {{HN}}}_g$ is functorial. Let $\alpha \in \operatorname {\mathrm {Hom}}_G(V,W)$ be a morphism of finite-dimensional G-modules. We need to show that $\alpha _{\mathcal {R}}(V_{\mathcal {R}}^x)\subseteq W_{\mathcal {R}}^x$ for all ${x\in \mathbb {Q}}$ . Choose by Lemma 2.2 an admissible extension E of K such that $\kappa _E$ is strongly difference-closed. For any fixed $x\in \mathbb {Q}$ , we set $V_{\operatorname {\mathrm {\tilde {\mathcal {R}}}}(E,t)}^x := V_{\mathcal {R}}^x\bigotimes _{\mathcal {R}} {\operatorname {\mathrm {\tilde {\mathcal {R}}}}(E,t)}$ , and $W_{\operatorname {\mathrm {\tilde {\mathcal {R}}}}(E,t)}^x := W_{\mathcal {R}}^x\bigotimes _{\mathcal {R}} {\operatorname {\mathrm {\tilde {\mathcal {R}}}}(E,t)}$ . By Lemma 2.3(iii), we have a decomposition $W_{\operatorname {\mathrm {\tilde {\mathcal {R}}}}(E,t)}=W_{\operatorname {\mathrm {\tilde {\mathcal {R}}}}(E,t)}^x \operatorname *{\mathrm {\bigoplus }} W_{\operatorname {\mathrm {\tilde {\mathcal {R}}}}(E,t)}'$ of $\varphi $ -modules over $\operatorname {\mathrm {\tilde {\mathcal {R}}}}(E,t)$ , where $W_{\operatorname {\mathrm {\tilde {\mathcal {R}}}}(E,t)}^x$ (resp. $W_{\operatorname {\mathrm {\tilde {\mathcal {R}}}}(E,t)}'$ ) has slopes less or equal to x (resp. greater than x). By Lemma 2.3(iv), the induced morphism $V_{\operatorname {\mathrm {\tilde {\mathcal {R}}}}(E,t)}^x \to W_{\operatorname {\mathrm {\tilde {\mathcal {R}}}}(E,t)}'$ of $\varphi $ -modules is zero. We thus have $\alpha _{\operatorname {\mathrm {\tilde {\mathcal {R}}}}(E,t)} \big (V_{\operatorname {\mathrm {\tilde {\mathcal {R}}}}(E,t)}^x \big )\subseteq W_{\operatorname {\mathrm {\tilde {\mathcal {R}}}}(E,t)}^x$ . Given any $\textbf {v}\in V_{\mathcal {R}}^x$ , we may write $\alpha _{\operatorname {\mathrm {\tilde {\mathcal {R}}}}(E,t)}(\textbf {v}\otimes 1)=\alpha _{\mathcal {R}}(\textbf {v})\otimes 1=\operatorname *{\mathrm {\sum }}\limits _{i\in I} \textbf {w}_i\otimes s_i$ for some finite set I, with $\textbf {w}_i\in W_{\mathcal {R}}^x$ and $s_i\in {\operatorname {\mathrm {\tilde {\mathcal {R}}}}(E,t)}$ for all $i\in I$ . Let M be the $\mathcal {R}$ -submodule of $W_{\mathcal {R}}$ generated by $\alpha _{\mathcal {R}}(\textbf {v})$ and the $\textbf {w}_i$ , and let N be the $\mathcal {R}$ -submodule of $W_{\mathcal {R}}^x$ generated by the $\textbf {w}_i$ . We then have $(M/N)\bigotimes _{\mathcal {R}} \operatorname {\mathrm {\tilde {\mathcal {R}}}}(E,t) \cong (M\bigotimes _{\mathcal {R}} \operatorname {\mathrm {\tilde {\mathcal {R}}}}(E,t))/ (N\bigotimes _{\mathcal {R}} \operatorname {\mathrm {\tilde {\mathcal {R}}}}{(E,t))=0}$ . It follows that $M/N=0$ as $\mathcal {R}\to \operatorname {\mathrm {\tilde {\mathcal {R}}}}(E,t)$ is faithfully flat. We thus have $\alpha _{\mathcal {R}}(\textbf {v})\in N\subseteq W_{\mathcal {R}}^x$ , as desired.

It remains to show that $\operatorname {\mathrm {{HN}}}_g$ preserves tensor products (in the sense of Remark 2.10(ii)). Let V and W be two finite-dimensional G-modules, and suppose that the slope filtration of $(V_{\mathcal {R}},g\varphi )$ (resp. $(W_{\mathcal {R}},g\varphi )$ ) has jumps $\mu _1<\cdots <\mu _{l_V}$ (resp. ${\nu _1<\cdots <\nu _{l_W}}$ ). By [Reference Kedlaya12, Lemma 16.4.3], $\big ((V\bigotimes _F W)_{\mathcal {R}},g\varphi \big )$ has jumps $\{\mu _i+\nu _j \mid 1\leq i\leq {l_V}, 1\leq j\leq {l_W}\}$ . Fix any $1\leq l\leq {l_V}$ and $1\leq s\leq {l_W}$ ; we need to show

(3) $$ \begin{align} (V \bigotimes_F W)_{\mathcal{R}}^{\mu_{l}+\nu_{s}}= \operatorname*{\mathrm{\sum}}\limits_{\scriptstyle \begin{array}{c} x,y\in\mathbb{Q} \\ x+y=\mu_{l}+\nu_{s}\end{array}} V_{\mathcal{R}}^x \bigotimes_{\mathcal{R}} W_{\mathcal{R}}^y, \end{align} $$

and we will do so in the remainder of the proof.

We claim that

$$ \begin{align*}\operatorname*{\mathrm{\sum}}\limits_{\scriptstyle \begin{array}{c} x,y\in\mathbb{Q} \\ x+y=\mu_{l}+\nu_{s}\end{array}} V_{\mathcal{R}}^x \bigotimes_{\mathcal{R}} W_{\mathcal{R}}^y= \operatorname*{\mathrm{\sum}}\limits_{\scriptstyle \begin{array}{c} \mu_i+\nu_j\leq \mu_l+\nu_s \\ 1\leq i\leq {l_V}, 1\leq j\leq {l_W}\end{array}}V_{\mathcal{R}}^{\mu_i} \bigotimes_{\mathcal{R}} W_{\mathcal{R}}^{\nu_j}.\end{align*} $$

It is clear that the RHS is contained in the LHS; we now show the reverse inclusion. Let $x,y\in \mathbb {Q}$ such that $x+y=\mu _{l}+\nu _{s}$ . If $x<\mu _1$ or $y<\nu _1$ , then $V_{\mathcal {R}}^x\bigotimes _{\mathcal {R}} W_{\mathcal {R}}^y=0$ which is contained in the RHS. Otherwise, there exists the largest integer $1\leq i\leq {l_V}$ (resp. $1\leq j\leq {l_W}$ ) with the property that $\mu _i\leq x$ (resp. $\nu _j\leq y$ ). We then have $V_{\mathcal {R}}^x\bigotimes _{\mathcal {R}} W_{\mathcal {R}}^y=V_{\mathcal {R}}^{\mu _i}\bigotimes _{\mathcal {R}} W_{\mathcal {R}}^{\nu _j}$ and $\mu _i+\nu _j\leq \mu _{l}+\nu _{s}$ . The claim is thus proved.

From Lemma 2.3(iii), we see that

$$ \begin{align*} \big(V \bigotimes_F W \big)_{\operatorname{\mathrm{\tilde{\mathcal{R}}}}(E,t)}^{\mu_{l}+\nu_{s}}= \Big( \operatorname*{\mathrm{\sum}}\limits_{\scriptstyle \begin{array}{c} \mu_i+\nu_j\leq \mu_{l}+\nu_{s} \\ 1\leq i\leq {l_V}, 1\leq j\leq {l_W}\end{array}} V_{\mathcal{R}}^{\mu_i}\bigotimes_{\mathcal{R}} W_{\mathcal{R}}^{\nu_j} \Big)\bigotimes_{\mathcal{R}} {\operatorname{\mathrm{\tilde{\mathcal{R}}}}(E,t)}.\end{align*} $$

Therefore, we have

$$ \begin{align*} \big(V \bigotimes W \big)_{\mathcal{R}}^{\mu_{l}+\nu_{s}}=\operatorname*{\mathrm{\sum}}\limits_{\scriptstyle \begin{array}{c}\mu_i+\nu_j\leq \mu_{l}+\nu_{s} \\ 1\leq i\leq {l_V}, 1\leq j\leq {l_W}\end{array}} V_{\mathcal{R}}^{\mu_i}\bigotimes_{\mathcal{R}} W_{\mathcal{R}}^{\nu_j}\end{align*} $$

by Lemma 2.3(i) and the fact that $\mathcal {R} \to {\operatorname {\mathrm {\tilde {\mathcal {R}}}}(E,t)}$ is faithfully flat. The desired equality (1) then follows from the preceding claim.▪

Let $(M,\varphi )$ be a $\varphi $ -module over $\operatorname {\mathrm {\tilde {\mathcal {R}}}}$ of rank d. Then, $\Phi $ is invertible, because the Frobenius lift on $\operatorname {\mathrm {\tilde {\mathcal {R}}}}$ is bijective, and $(M,\varphi ^{-1})$ is a $\varphi ^{-1}$ -module over $\operatorname {\mathrm {\tilde {\mathcal {R}}}}$ . More explicitly, let $A\in \operatorname {\mathrm {GL}}_d(\operatorname {\mathrm {\tilde {\mathcal {R}}}})$ be the matrix of action of $\varphi $ in some basis for M over $\operatorname {\mathrm {\tilde {\mathcal {R}}}}$ . Then, in the same basis, the matrix of action of $\varphi ^{-1}$ is $\varphi ^{-1}(A^{-1})$ . For example, if $M=V_{\operatorname {\mathrm {\tilde {\mathcal {R}}}}}$ for some $V\in \operatorname {\mathrm {\textbf {Rep}}}_F(G)$ , and $\Phi =\psi (g)\varphi $ where $\psi $ denotes (by abuse of notation) the group morphism $G(\mathcal {R}) \to G(\operatorname {\mathrm {\tilde {\mathcal {R}}}})$ induced by the embedding $\psi \colon \mathcal {R} \to \operatorname {\mathrm {\tilde {\mathcal {R}}}}$ recalled above Proposition 3.5, then

$$ \begin{align*} \big(\psi(g)\varphi \big) \cdot \big(\varphi^{-1}(\psi(g^{-1}))\varphi^{-1}\big)=1\end{align*} $$

in $G(\operatorname {\mathrm {\tilde {\mathcal {R}}}})\rtimes \operatorname {\mathrm {\langle }}\varphi \operatorname {\mathrm {\rangle }}$ (cf. Remark 3.2), which implies that $\varphi ^{-1}=\varphi ^{-1}(\psi (g^{-1}))\varphi ^{-1}$ .

Let M be a standard $\varphi $ -module over $\operatorname {\mathrm {\tilde {\mathcal {R}}}}$ of slope $\mu =s/r$ with $r>0$ and $(s,r)=1$ . Namely, we have a standard basis $e_1,\ldots ,e_r$ in which $\varphi $ acts via

$$ \begin{align*}A= \left(\begin{smallmatrix} 0 & & &\varpi^s\\ 1 &\ddots & & \\ &\;\ddots &\;\;\ddots &\\ && 1 &0 \end{smallmatrix}\right). \end{align*} $$

Then,

$$ \begin{align*}\varphi^{-1}(A^{-1})= \left(\begin{smallmatrix} 0 &1 & &\\ &\ddots &\ddots & \\ & & \;\;\ddots & {{\scriptstyle 1}} \\ \varpi^{-s} & & &0 \end{smallmatrix}\right), \end{align*} $$

which implies that $(M,\varphi ^{-1})$ is a standard $\varphi ^{-1}$ -module pure of slope $-\mu $ .

Proposition 3.6 The functor $\operatorname {\mathrm {{HN}}}_g\colon \operatorname {\mathrm {\textbf {Rep}}}_F(G) \to \mathbb {Q}\text {-}\operatorname {\mathrm {\textbf {Fil}}}_{\mathcal {R}}$ is exact.

Proof Let $\alpha \in \operatorname {\mathrm {Hom}}_G(V,W)$ be a morphism of finite-dimensional G-modules. We need to show that $\alpha _{\mathcal {R}}(V_{\mathcal {R}}^x)=\alpha _{\mathcal {R}}(V_{\mathcal {R}})\cap W_{\mathcal {R}}^x$ for all $x\in \mathbb {Q}$ . For any fixed $x\in \mathbb {Q}$ , the functoriality in Proposition 3.5 already implies that $\alpha _{\mathcal {R}}(V_{\mathcal {R}}^x)\subseteq \alpha _{\mathcal {R}}(V_{\mathcal {R}})\cap W_{\mathcal {R}}^x$ . Thus, it suffices to show that for any nonzero element $\textbf {v}\in V_{\mathcal {R}}$ such that $\alpha _{\mathcal {R}}(\textbf {v})\in W_{\mathcal {R}}^x$ , there exists $\textbf {v}'\in V_{\mathcal {R}}^x$ with $\alpha _{\mathcal {R}}(\textbf {v})=\alpha _{\mathcal {R}}(\textbf {v}')$ .

By Lemma 2.3(iii), we have decompositions

(4) $$ \begin{align} V_{\operatorname{\mathrm{\tilde{\mathcal{R}}}}(E,t)}=V_{\operatorname{\mathrm{\tilde{\mathcal{R}}}}(E,t)}^x \operatorname*{\mathrm{\bigoplus}} V_{\operatorname{\mathrm{\tilde{\mathcal{R}}}}(E,t)}'~~~\text{and}~~~W_{\operatorname{\mathrm{\tilde{\mathcal{R}}}}(E,t)}=W_{\operatorname{\mathrm{\tilde{\mathcal{R}}}}(E,t)}^x \operatorname*{\mathrm{\bigoplus}} W_{\operatorname{\mathrm{\tilde{\mathcal{R}}}}(E,t)}' \end{align} $$

of $\varphi $ -modules over $\operatorname {\mathrm {\tilde {\mathcal {R}}}}(E,t)$ , in which $V_{\operatorname {\mathrm {\tilde {\mathcal {R}}}}(E,t)}^x$ and $W_{\operatorname {\mathrm {\tilde {\mathcal {R}}}}(E,t)}^x$ have slopes less or equal to x, while $V_{\operatorname {\mathrm {\tilde {\mathcal {R}}}}(E,t)}'$ and $W_{\operatorname {\mathrm {\tilde {\mathcal {R}}}}(E,t)}'$ have slopes greater than x. Notice that the composition

is a morphism of $\varphi $ -modules. We claim that $\xi =0$ . We write $\varphi =\psi (g)\varphi $ , then $\varphi ^{-1}=\varphi ^{-1}(\psi (g^{-1}))\varphi ^{-1}$ . Because $\alpha $ is G-equivariant and $\varphi ^{-1}(\psi (g^{-1}))\in G(\operatorname {\mathrm {\tilde {\mathcal {R}}}}(E,t))$ , we have that $\alpha _{\operatorname {\mathrm {\tilde {\mathcal {R}}}}}\colon (V_{\operatorname {\mathrm {\tilde {\mathcal {R}}}}(E,t)},\varphi ^{-1}) \to (W_{\operatorname {\mathrm {\tilde {\mathcal {R}}}}(E,t)},\varphi ^{-1})$ is a morphism of $\varphi ^{-1}$ -modules. On the other hand, we also have decompositions of $\varphi ^{-1}$ -modules as in (2), together with the induced morphism $\xi \colon V_{\operatorname {\mathrm {\tilde {\mathcal {R}}}}(E,t)}' \to W_{\operatorname {\mathrm {\tilde {\mathcal {R}}}}(E,t)}^x$ of $\varphi ^{-1}$ -modules. But in this case, $V_{\operatorname {\mathrm {\tilde {\mathcal {R}}}}(E,t)}'$ has slopes less than x, while $W_{\operatorname {\mathrm {\tilde {\mathcal {R}}}}(E,t)}^x$ has slopes greater or equal to x. It then follows from Lemma 2.3(iv) that $\xi =0$ , as claimed.

Therefore, we find $\textbf {v}_1,\ldots ,\textbf {v}_n \in V_{\mathcal {R}}^x$ and $s_1,\ldots ,s_n\in \operatorname {\mathrm {\tilde {\mathcal {R}}}}(E,t)$ such that

$$ \begin{align*} \alpha_{\operatorname{\mathrm{\tilde{\mathcal{R}}}}(E,t)}(\textbf{v}\otimes 1)=\alpha_{\mathcal{R}}(\textbf{v})\otimes 1 =\operatorname*{\mathrm{\sum}}\limits_{i=1}^n \alpha_{\mathcal{R}}(\textbf{v}_i)\otimes s_i.\end{align*} $$

Let M be the submodule of $W_{\mathcal {R}}$ generated by $\alpha _{\mathcal {R}}(\textbf {v})$ and the $\alpha _{\mathcal {R}}(\textbf {v}_i)$ , and let N be the submodule generated by the $\alpha _{\mathcal {R}}(\textbf {v}_i)$ . We then have

$$ \begin{align*}(M/N)\bigotimes_{\mathcal{R}} \operatorname{\mathrm{\tilde{\mathcal{R}}}}(E,t) \cong (M\bigotimes_{\mathcal{R}} \operatorname{\mathrm{\tilde{\mathcal{R}}}}(E,t))/ (N\bigotimes_{\mathcal{R}} \operatorname{\mathrm{\tilde{\mathcal{R}}}}(E,t))=0.\end{align*} $$

It follows that $M/N=0$ as $\mathcal {R}\to \operatorname {\mathrm {\tilde {\mathcal {R}}}}(E,t)$ is faithfully flat, and hence, $\alpha _{\mathcal {R}}(\textbf {v})=\sum \limits _{i=1}^n r_i\alpha _{\mathcal {R}}(\textbf {v}_i)\in W_{\mathcal {R}}^x$ for some $r_i\in \mathcal {R}$ . Put $\textbf {v}':= \sum \limits _{i=1}^n r_i\textbf {v}_i \in V_{\mathcal {R}}^x$ , we then have $\alpha _{\mathcal {R}}(\textbf {v}')=\alpha _{\mathcal {R}}(\textbf {v})$ , as desired.▪

3.3 Splittings of $\operatorname {\mathrm {{HN}}}_g$

As before, we fix an element $g\in G(\mathcal {R})$ . In Section 3.2, we have constructed a $\mathbb {Q}$ -filtered fiber functor $\operatorname {\mathrm {{HN}}}_g\colon \operatorname {\mathrm {\textbf {Rep}}}_F(G) \to \mathbb {Q}$ - $\operatorname {\mathrm {\textbf {Fil}}}_{\mathcal {R}}$ . In this subsection, we show that $\operatorname {\mathrm {{HN}}}_g$ is splittable whenever G is smooth. Our strategy goes as follows. We first use Lemma 3.10 reducing $\operatorname {\mathrm {{HN}}}_g$ to a $\mathbb {Z}$ -filtered fiber functor $\operatorname {\mathrm {{HN}}}_g^{\mathbb {Z}}$ to which Theorem 2.12 is applicable. This $\operatorname {\mathrm {{HN}}}_g^{\mathbb {Z}}$ then admits a $\mathbb {Z}$ -splitting. Finally, in Theorem 3.12, we lift such a $\mathbb {Z}$ -splitting to a $\mathbb {Q}$ -splitting of $\operatorname {\mathrm {{HN}}}_g$ .

Definition 3.7 We define the support of $\operatorname {\mathrm {{HN}}}_g$ by

$$ \begin{align*}\operatorname{\mathrm{Supp}}(\operatorname{\mathrm{{HN}}}_g) := \{x\in \mathbb{Q} \mid \operatorname{\mathrm{gr}}^x_{\operatorname{\mathrm{{HN}}}_g}(V)\neq 0~\text{for some}~ V\in \operatorname{\mathrm{\textbf{Rep}}}_F(G)\}.\end{align*} $$

Notice that $\operatorname {\mathrm {Supp}}(\operatorname {\mathrm {{HN}}}_g)$ is the set of jumps of the slope filtrations of $(V_{\mathcal {R}},g\varphi )$ for all $V\in \operatorname {\mathrm {\textbf {Rep}}}_F(G)$ .

The general idea of the following construction was addressed in [Reference Anschütz2], after Definition 2.5 in loc. cit.; we will make it more explicit in our case.

Construction 3.8 Let $W\in \operatorname {\mathrm {\textbf {Rep}}}_F(G)$ be a faithful representation. Because G is algebraic, W is a tensor generator for $\operatorname {\mathrm {\textbf {Rep}}}_F(G)$ , i.e., any representation V of G is a subquotient of a direct sum of representations $\bigotimes ^m (W \bigoplus W^{\vee })$ for various $m\in \mathbb {N}$ . (See [Reference Milne18, Theorem 4.14].) Therefore, $\operatorname {\mathrm {Supp}}(\operatorname {\mathrm {{HN}}}_g)$ is the additive subgroup of $\mathbb {Q}$ finitely generated by the $\mathbb {Q}$ -jumps $\nu _1,\ldots ,\nu _n$ of $(W_{\mathcal {R}},g\varphi )$ . We write $\nu _i=s_i/d_i$ with $d_i>0$ and $(s_i,d_i)=1$ for $1\leq i\leq n$ . Let $d_g\in \mathbb {N}$ be the least common multiple of the $d_i$ . We then have $d_g\nu _i\in \mathbb {Z}$ for $1\leq i\leq n$ . In particular, we have

$$ \begin{align*}d_g=\min \{m\in\mathbb{N} \mid mx\in \mathbb{Z}, \forall x\in \operatorname{\mathrm{Supp}}(\operatorname{\mathrm{{HN}}}_g)\}.\end{align*} $$

Therefore, $d_g$ is uniquely determined by g. We call $d_g$ the least common denominator of g.

Remark 3.9 We conclude from Construction 3.8 that $\operatorname {\mathrm {Supp}}(\operatorname {\mathrm {{HN}}}_g)$ is isomorphic to $\mathbb {Z}$ or $0$ . In fact, if $(W_{\mathcal {R}},g\varphi )$ has only one jump $0$ , then $\operatorname {\mathrm {Supp}}(\operatorname {\mathrm {{HN}}}_g)=0$ . Otherwise, the choice of $d_g$ implies that $\gcd (d_g\nu _1,\ldots ,d_g\nu _n)=1$ . We then have $d_g\operatorname {\mathrm {Supp}}(\operatorname {\mathrm {{HN}}}_g)=\mathbb {Z}$ , because the $d_g\nu _i$ generate $\mathbb {Z}$ as a $\mathbb {Z}$ -module. Therefore, $x \mapsto d_gx$ gives an isomorphism $\operatorname {\mathrm {Supp}}(\operatorname {\mathrm {{HN}}}_g)\cong \mathbb {Z}$ .

Lemma 3.10 $\operatorname {\mathrm {{HN}}}_g$ factors through a $\mathbb {Z}$ -filtered fiber functor $\operatorname {\mathrm {{HN}}}^{\mathbb {Z}}_g\colon \operatorname {\mathrm {\textbf {Rep}}}_F(G) \to \mathbb {Z}$ - $\operatorname {\mathrm {\textbf {Fil}}}_{\mathcal {R}}$ which makes the diagram

commute.

We remark that the functor ${[d_g^{-1}]}_*$ (see Construction 2.11) is nothing but relabeling the jumps by multiplying all jumps with $d_g^{-1}$ . In particular, this lemma implies that $\operatorname {\mathrm {gr}}^x_{\operatorname {\mathrm {{HN}}}_g}(V)=\operatorname {\mathrm {gr}}^{d_g^{-1}x}_{\operatorname {\mathrm {{HN}}}^{\mathbb {Z}}_g}(V)$ for all $x\in \mathbb {Q}$ and $V\in \operatorname {\mathrm {\textbf {Rep}}}_F(G)$ .

Proof of Lemma 3.10 Let $V\in \operatorname {\mathrm {\textbf {Rep}}}_F(G)$ , and let $\mu _1,\ldots ,\mu _l$ be the $\mathbb {Q}$ -jumps of $(V_{\mathcal {R}},g\varphi )$ . We then have $d_g\mu _i\in \mathbb {Z}$ for all i. We have an increasing map

$$ \begin{align*} \mathcal{F}_g \colon \mathbb{Z} &\operatorname{\mathrm{\longrightarrow}} \{\mathcal{R}\text{-submodules of}~V_{\mathcal{R}}\},\\ x &\longmapsto \mathcal{F}_g^x(V_{\mathcal{R}}), \end{align*} $$

where

$$ \begin{align*}\mathcal{F}_g^x(V_{\mathcal{R}}):= \left\{ \begin{array}{ll} 0 &\text{for}~x<d_g\mu_1,\\ \mathcal{HN}_g^{\mu_i}(V_{\mathcal{R}}) ~~~ &\text{for}~ d_g\mu_i\leq x <d_g\mu_{i+1}, 1\leq i\leq l-1,\\ V_{\mathcal{R}} &\text{for}~x\geq d_g\mu_l. \end{array} \right.\end{align*} $$

Then, $(V_{\mathcal {R}},\mathcal {F}_g)$ is a $\mathbb {Z}$ -filtered module over $\mathcal {R}$ with $\mathbb {Z}$ -jumps $d_g\mu _1<\cdots <d_g\mu _l$ . We thus have a $\mathbb {Z}$ -filtered fiber functor

$$ \begin{align*} \operatorname{\mathrm{{HN}}}^{\mathbb{Z}}_g \colon \operatorname{\mathrm{\textbf{Rep}}}_F(G) &\operatorname{\mathrm{\longrightarrow}} \mathbb{Z}\text{-}\operatorname{\mathrm{\textbf{Fil}}}_{\mathcal{R}},\\ V &\longmapsto (V_{\mathcal{R}},\mathcal{F}_g), \end{align*} $$

satisfying $\operatorname {\mathrm {{HN}}}_g=[d_g^{-1}]_*\circ \operatorname {\mathrm {{HN}}}_g^{\mathbb {Z}}$ .▪

By the definition of $\operatorname {\mathrm {\underline {Aut}}}^{\otimes }$ and Corollary 2.5, we have $\operatorname {\mathrm {\underline {Aut}}}^{\otimes }(\omega ^G)(R)=\operatorname {\mathrm {Aut}}^{\otimes } (\omega ^G_R)\cong G(R)$ for all $R\in \operatorname {\mathrm {\textbf {Alg}}}_k$ . For any R-algebra S, we then have

$$ \begin{align*}\operatorname{\mathrm{\underline{Aut}}}^{\otimes}(\omega_R^G)(S)=\operatorname{\mathrm{Aut}}^{\otimes} (\omega^G_R\otimes S)=\operatorname{\mathrm{Aut}}^{\otimes} (\omega^G_S)\cong G_R(S).\end{align*} $$

Proposition 3.11 Let G be a smooth F-group. Then, $\operatorname {\mathrm {{HN}}}^{\mathbb {Z}}_g$ is splittable.

Proof Because $\operatorname {\mathrm {forg}}\circ \operatorname {\mathrm {{HN}}}^{\mathbb {Z}}_g=\omega ^G\otimes \mathcal {R}$ , we have

$$ \begin{align*} \operatorname{\mathrm{\underline{Aut}}}^{\otimes}_{\mathcal{R}}(\operatorname{\mathrm{forg}}\circ \operatorname{\mathrm{{HN}}}^{\mathbb{Z}}_g)= \operatorname{\mathrm{\underline{Aut}}}^{\otimes}_{\mathcal{R}}(\omega^G_{\mathcal{R}})\cong G_{\mathcal{R}}.\end{align*} $$

Notice that $G_{\mathcal {R}}$ is smooth over $\mathcal {R}$ ; the proposition then follows from Theorem 2.12.▪

Theorem 3.12 Let G be a smooth F-group. Then, the $\mathbb {Q}$ -filtered fiber functor $\operatorname {\mathrm {{HN}}}_g$ is splittable.

Proof Choose a splitting $\tau _g\colon \operatorname {\mathrm {\textbf {Rep}}}_F(G)\to \mathbb {Z}$ - $\operatorname {\mathrm {\textbf {Grad}}}_{\mathcal {R}}$ of $\operatorname {\mathrm {{HN}}}^{\mathbb {Z}}_g$ by Proposition 3.11, we then have a $\mathbb {Q}$ -graded fiber functor ${[d_g^{-1}]}_*\circ \tau _g\colon \operatorname {\mathrm {\textbf {Rep}}}_F(G)\to \mathbb {Q}$ - $\operatorname {\mathrm {\textbf {Grad}}}_{\mathcal {R}}$ . On the other hand, we have the diagram

(5)

with the upper-left, the upper-right, and the bottom triangles commutative. Here, the commutativity of the upper-left (resp. the upper-right) triangle follows from Proposition 3.11 (resp. Lemma 3.10); for the bottom one, we note that ${[d_g^{-1}]}_*\circ \operatorname {\mathrm {fil}}=\operatorname {\mathrm {fil}}\circ {[d_g^{-1}]}_*$ . Hence, the outer diagram also commutes, which implies that $\operatorname {\mathrm {{HN}}}_g$ factors through the $\mathbb {Q}$ -graded fiber functor ${[d_g^{-1}]}_*\circ \tau _g$ , as desired.▪

3.4 The slope morphism

Let R be a commutative ring with $1$ , and let $\Gamma $ be an abelian group (not necessarily finitely generated). We first continue the discussions in Section 2.5 to see how $\Gamma $ -gradings over R are related to $D_R(\Gamma )$ -modules, for some affine group scheme $D_R(\Gamma )$ which will be defined as follows.

The group algebra $R[\Gamma ]:= \bigoplus \limits _{\gamma \in \Gamma } Re_\gamma $ carries a Hopf algebra structure, where the comultiplication is given by $\Delta (e_\gamma )=e_\gamma \otimes e_\gamma $ , the counit is given by $\epsilon (e_\gamma )=1$ , and the antipode is given by $S(e_\gamma )=e_{-\gamma }$ , for all $\gamma \in \Gamma $ . We denote by $D_R(\Gamma )$ the affine R-group scheme represented by $R[\Gamma ]$ . For any $\gamma \in \Gamma $ , the Hopf algebra morphism $R[\mathbb {Z}] \to R[\Gamma ], e_1 \mapsto e_\gamma $ gives rise to a character $\chi _\gamma \colon D_R(\Gamma ) \to \mathbb {G}_{m,R}$ of $D_R(\Gamma )$ . For the remainder of this paper, we denote by $\mathbb {D}_R$ the R-group scheme $D_R(\mathbb {Q})$ .

Let $M=\bigoplus _{\gamma \in \Gamma } M_\gamma $ be a $\Gamma $ -graded module over R. Then, M becomes a $D_R(\Gamma )$ -module where $D_R(\Gamma )$ acts on each $M_\gamma $ via $\chi _\gamma $ . The following lemma shows that this assignment gives an equivalence of categories.

Lemma 3.13 [Reference Demazure and Gabriel8, Proposition II.2.5] $\Gamma $ - $\operatorname {\mathrm {\textbf {Grad}}}_R$ is equivalent to the category of $D_{R}(\Gamma )$ -modules.

Corollary 3.14 For any $\gamma \in \mathbb {Q}_{>0}$ , the functor $[\gamma ]_*\colon \mathbb {Z}\text {-}\operatorname {\mathrm {\textbf {Grad}}}_R \to \mathbb {Q}\text {-}\operatorname {\mathrm {\textbf {Grad}}}_R$ corresponds to the character $\chi _\gamma \colon \mathbb {D}_R\to \mathbb {G}_{m,R}$ .

Proof Let $M\in \mathbb {Z}\text {-}\operatorname {\mathrm {\textbf {Grad}}}_R$ . By Lemma 3.13, we may write $M=\bigoplus \limits _{n\in \mathbb {Z}} M_n$ as a direct sum of eigenmodules. By construction, we have $[\gamma ]_*(M)=\bigoplus \limits _{n\in \mathbb {Z}} ([\gamma ]_*(M))_{\gamma n}$ with $([\gamma ]_*(M))_{\gamma n}=M_n$ for all n, which is also a decomposition into eigenmodules. Therefore, giving $[\gamma ]_*$ is equivalent to giving the commutative diagram

of R-modules for all $n\in \mathbb {Z}$ such that $M_n\neq 0$ . Here, the left (resp. the right) vertical arrow is given by $m\mapsto m\otimes e_n$ (resp. $m\mapsto m\otimes e_{\gamma n}$ ). The diagram then corresponds to $R[\mathbb {Z}]\to R[\mathbb {Q}],~e_1\mapsto e_\gamma $ , as desired.▪

We now apply the preceding discussions to the functors constructed in Section 3.3, following [Reference Kottwitz14, Section 4].

Construction 3.15 Let $g\in G(\mathcal {R})$ ; we fix a splitting $\tau _g$ of $\operatorname {\mathrm {{HN}}}_g^{\mathbb {Z}}$ given by Proposition 3.11. For any $(V,\rho )\in \operatorname {\mathrm {\textbf {Rep}}}_F(G)$ , $\tau _g$ gives a decomposition of $V_{\mathcal {R}}$ , which induces a morphism $\lambda _{\rho ,g}\colon \mathbb {G}_{m,\mathcal {R}} \to \operatorname {\mathrm {GL}}_{V,\mathcal {R}}$ by Lemma 3.13. Let S be an $\mathcal {R}$ -algebra, and let $a\in \mathbb {G}_{m,\mathcal {R}}(S)$ . We then have a family

$$ \begin{align*} \big\{ \lambda_{\rho,g}(a)\colon V_S \to V_S \mid (V,\rho)\in\operatorname{\mathrm{\textbf{Rep}}}_F(G) \big\} \end{align*} $$

of S-linear maps. Because $\tau _g$ is a tensor functor, this family satisfies conditions (i–iii) in Theorem 2.4. We thus find a unique element $b\in G_{\mathcal {R}}(S)$ such that $\lambda _{\rho ,g}(a)=\rho (b)$ for all $(V,\rho )\in \operatorname {\mathrm {\textbf {Rep}}}_F(G)$ . The assignment $a\mapsto b$ is functorial in S, because both $\lambda _{\rho ,g}$ and $\rho $ are functorial. We then have a morphism of $\mathcal {R}$ -groups

$$ \begin{align*}\lambda_g\colon \mathbb{G}_{m,\mathcal{R}} \operatorname{\mathrm{\longrightarrow}} G_{\mathcal{R}},\end{align*} $$

which is said to be the $\mathbb {Z}$ -slope morphism of g.

By Corollary 3.14, ${[d_g^{-1}]}_*$ gives a unique morphism $\chi _{d_g^{-1}}\colon \mathbb {D}_{\mathcal {R}} \to \mathbb {G}_{m,\mathcal {R}}$ . We define

$$ \begin{align*}\nu_g:= \lambda_g\circ \chi_{d_g^{-1}} \colon \mathbb{D}_{\mathcal{R}} \operatorname{\mathrm{\longrightarrow}} G_{\mathcal{R}},\end{align*} $$

which is said to be the $\mathbb {Q}$ -slope morphism of g.

The following example demonstrates explicitly how $\lambda _g$ and $\nu _g$ are related to the splittings constructed in Section 3.3 (see Diagram 3).

Example 3.16 Let $(V,\rho )\in \operatorname {\mathrm {\textbf {Rep}}}_F(G)$ and suppose that the slope filtration of $(V_{\mathcal {R}},g\varphi )$ is

$$ \begin{align*}0\subseteq V_{\mathcal{R}}^{\mu_1} \subseteq \cdots \subseteq V_{\mathcal{R}}^{\mu_l}=V_{\mathcal{R}}\end{align*} $$

with jumps $\mu _1<\cdots <\mu _l$ . By Theorem 3.12, the functor $[d_g^{-1}]_* \circ \tau _g \colon \operatorname {\mathrm {\textbf {Rep}}}_F(G) \to \mathbb {Q}\text {-}\operatorname {\mathrm {\textbf {Grad}}}_{\mathcal {R}}$ gives a splitting

(6) $$ \begin{align} \textstyle V_{\mathcal{R}}=V_{\mathcal{R},\mu_1}\operatorname*{\mathrm{\bigoplus}}\cdots \operatorname*{\mathrm{\bigoplus}} V_{\mathcal{R},\mu_l} \end{align} $$

of $\operatorname {\mathrm {{HN}}}_g(V)$ , i.e., we have $\bigoplus \limits _{i=1}^j V_{\mathcal {R},\mu _i}=V_{\mathcal {R}}^{\mu _j}$ for all $1\leq j\leq l$ .

First, we fix $1\leq i\leq l$ . Let $S\in \operatorname {\mathrm {\textbf {Alg}}}_{\mathcal {R}}$ and $a\in \mathbb {D}_{\mathcal {R}}(S)$ , then $\rho \circ \nu _g(a)$ acts on $V_{\mathcal {R},\mu _i}\bigotimes _{\mathcal {R}} S$ via multiplication by $\chi _{\mu _i}(a)$ . By Lemma 3.10, $\rho \circ \lambda _g(b)$ acts on $V_{\mathcal {R},\mu _i}$ via multiplication by $b^{d_g\mu _i}$ , for all $b\in \mathbb {G}_{m,\mathcal {R}}(S)$ . Notice that for any $\frac {m}{n}\in \mathbb {Q}$ , we have $e_{\frac {m}{n}}=(e_{\frac {1}{n}})^m \in \mathcal {R}[\mathbb {Q}]$ , and hence, $\chi _{\frac {m}{n}}=(\chi _{\frac {1}{n}})^m$ . In particular, we have $\chi _{\mu _i}=\chi _{\frac {d_g\mu _i}{d_g}}=(\chi _{d_g^{-1}})^{d_g\mu _i}$ . Then, on $V_{\mathcal {R},\mu _i}\bigotimes _{\mathcal {R}} S$ , we have

$$ \begin{align*}\rho\circ \nu_g(a)=\chi_{\mu_i}(a)=(\chi_{d_g^{-1}}(a))^{d_g\mu_i}=\rho\circ \lambda_g\big( \chi_{d_g^{-1}}(a)\big)=\rho\circ \lambda_g\circ \chi_{d_g^{-1}}(a).\end{align*} $$

We next apply this result to all $1\leq i\leq l$ . Because $V_{\mathcal {R}}=\bigoplus \limits _{i=1}^l V_{\mathcal {R},\mu _i}$ , we conclude that $\rho \circ \nu _g=\rho \circ \lambda _g\circ \chi _{d_g^{-1}}$ . It follows that $\nu _g=\lambda _g\circ \chi _{d_g^{-1}}$ once we choose a faithful representation, as is expected from the definition of $\nu _g$ .

If $G=\operatorname {\mathrm {GL}}_V$ for some $V\in \operatorname {\mathrm {\textbf {Vec}}}_F$ , we consider the standard representation $(V,\rho )$ of G where $\rho $ is the identity. The discussion in the above example then indicates that the image of $\lambda _g$ is contained in a split maximal torus in $G_{\mathcal {R}}$ ; we conjecture that this property holds true for an arbitrary split reductive F-group G, and we shall give one more evidence as follows.

Example 3.17 Fix a d-dimensional F-vector space V. For any $R\in \operatorname {\mathrm {\textbf {Alg}}}_F$ , we define

$$ \begin{align*}\operatorname{\mathrm{SL}}_V(R) := \{ g\in \operatorname{\mathrm{GL}}_V(R) \mid \det(g)=1\} .\end{align*} $$

The affine algebraic F-group $\operatorname {\mathrm {SL}}_V$ is called the special linear group (associated to V).

Fix an arbitrary $g\in \operatorname {\mathrm {SL}}_V(\mathcal {R})$ . With the notation as in Construction 4.14, we suppose the jumps of the slope filtration of $(V_{\mathcal {R}},\Phi _g)$ are $\mu _1,\ldots ,\mu _l$ and $\xi _g(V)=\bigoplus \limits _{i=1}^l V_{\mathcal {R},\mu _i}$ . For each i, we write $r_i=\operatorname {\mathrm {rk}}_{\mathcal {R}} (V_{\mathcal {R},\mu _i})$ , then the $r_i$ -th exterior product $\Lambda ^{r_i}(V_{\mathcal {R},\mu _i})$ is of rank  $1$ . We choose a generator $m_i$ , then $\Lambda ^{r_i}(\Phi _{g,\mu _i})(m_i)=f_i m_i$ for some $f_i\in \mathcal {R}^{\times }=(\mathcal {E}^{\dagger })^{\times }$ . Let $\nu $ be the valuation of the $1$ -Gauss norm on $\mathcal {E}^{\dagger }$ . We then have $\mu _i=\frac {\nu (f_i)}{r_i}$ by [Reference Kedlaya11, Definition 1.4.4].

Let $e_1,\ldots ,e_d$ be a basis for V over F, and let $A\in \operatorname {\mathrm {SL}}_d(\mathcal {R})$ be the matrix of action of $\Phi _g$ in $e_1\otimes 1,\ldots , e_d\otimes 1$ . Let $B\in \operatorname {\mathrm {GL}}_d(\mathcal {R})$ represent a change-of-basis over $\mathcal {R}$ . Then, in the new basis, the matrix of action of $\Phi _g$ is $B^{-1}A \varphi (B)$ . Notice that $\det (B)\in (\mathcal {E}^{\dagger })^{\times }$ and $\varphi $ preserves $\nu $ , we then have

$$ \begin{align*}\nu\big( \det(B^{-1}A \varphi(B) )\big)= \nu \big(\det(B^{-1}) \det(A) \varphi(\det(B))\big)=\nu(\det(A)), \end{align*} $$

which implies that the valuation of the determinant of the matrix of action of $\Phi _g$ is invariant under change-of-basis. We denote by $\nu (\det (\Phi _g))$ this invariant. In particular, we have $\nu (\det (\Phi _g))=0$ , because $\det (A)=1$ by assumption. Put $\Phi ^{\prime }_g := \bigoplus \limits _{i=1}^l \Phi _{g,\mu _i}$ , where each $\Phi _{g,\mu _i}$ is the projection of $\Phi _g$ to the $\mu _i$ -th graded piece of $\xi _g(V)$ (cf. Construction 4.14 below). We thus have

$$ \begin{align*}0=\nu(\det (\Phi_g))=\nu(\det (\Phi^{\prime}_g))=\nu(f_1)+\cdots +\nu(f_l)=r_1\mu_1+\cdots +r_l\mu_l.\end{align*} $$

Let $S\in \operatorname {\mathrm {\textbf {Alg}}}_{\mathcal {R}}$ and $t\in \mathbb {G}_{m,\mathcal {R}}(S)$ . Because $\lambda _g(t)$ acts on each $V_{\mathcal {R},\mu _i}\bigotimes _{\mathcal {R}} S$ via multiplication by $t^{d_g\mu _i}$ where $d_g$ is the least common denominator of g, we then have

$$ \begin{align*}\det(\lambda_g(t))= t^{d_g(r_1\mu_1+\cdots +r_l \mu_l)}=1.\end{align*} $$

Therefore, the image of $\lambda _g$ is contained in a split maximal torus in $\operatorname {\mathrm {SL}}_{V,\mathcal {R}}$ .

4 G- $(\varphi ,\nabla )$ -modules over the Robba ring

In this section, we fix an affine algebraic group F-group G.

4.1 Definition and an identification

Let $R\in \{\mathcal {E}^{\dagger },\mathcal {R}\}$ equipped with an absolute Frobenius lift $\varphi $ and the usual derivation $\partial =\partial _t=d/dt$ on R.

Definition 4.1 A G- $(\varphi ,\nabla )$ -module over R is an exact faithful F-linear tensor functor

$$ \begin{align*}\operatorname{\mathrm{I}}\colon \operatorname{\mathrm{\textbf{Rep}}}_F(G) \operatorname{\mathrm{\longrightarrow}} \textbf{Mod}^{\varphi,\nabla}_R,\end{align*} $$

which satisfies $\operatorname {\mathrm {forg}} \circ \operatorname {\mathrm {I}} = \omega ^G \otimes R$ , where $\operatorname {\mathrm {forg}} \colon \operatorname {\mathrm {\textbf {Mod}^{\varphi }_R}} \to \operatorname {\mathrm {\textbf {Mod}}}_R$ is the forgetful functor. The category of G- $(\varphi ,\nabla )$ -modules over R is denoted by $\operatorname {\mathrm {G-\textbf {Mod}^{\varphi ,\nabla }_R}}$ , whose morphisms are morphisms of tensor functors. A G- $(\varphi ,\nabla )$ -module $\operatorname {\mathrm {I}}$ over R is called unit-root if $\operatorname {\mathrm {I}}(V,\rho )$ is a unit-root $(\varphi ,\nabla )$ -module over R for all $(V,\rho ) \in \operatorname {\mathrm {\textbf {Rep}}}_F(G)$ .

Remark 4.2 We remark that $\operatorname {\mathrm {G-\textbf {Mod}^{\varphi ,\nabla }_R}}$ is a groupoid, because both $\operatorname {\mathrm {\textbf {Rep}}}_F(G)$ and $\textbf {Mod}^{\varphi ,\nabla }_R$ are rigid tensor categories over F, and any morphism of tensor functors between rigid tensor categories is an isomorphism by [Reference Deligne and Milne7, Proposition 1.13]. Note that tensor products and duals in $\textbf {Mod}^{\varphi ,\nabla }_R$ are defined as in [Reference Tsuzuki22, Section 3.1], and the identity object is $(R,\varphi ,\partial )$ .

We put

$$ \begin{align*}\operatorname{\mathrm{\boldsymbol\mu}} := \operatorname{\mathrm{\boldsymbol\mu}} (\varphi,t):= \partial(\varphi(t)).\end{align*} $$

Let $\Omega _R^1 := \Omega _{R/K}^1$ be the free R-module generated by the symbol $dt$ , with the K-linear derivation $d\colon R \to \Omega ^1_R, f\mapsto \partial (f) dt$ . We also define a $\varphi $ -linear endomorphism

$$ \begin{align*} d\varphi\colon \Omega^1_R \operatorname{\mathrm{\longrightarrow}} \Omega^1_R,\;\;\;\;\; fdt \longmapsto \operatorname{\mathrm{\boldsymbol\mu}}\varphi(f)dt. \end{align*} $$

Given a finite-dimensional representation $\rho \colon G\to \operatorname {\mathrm {GL}}_V$ , we have a morphism $\operatorname {\mathrm {Lie}}(\rho ) \colon \mathfrak {g}\to \operatorname {\mathrm {\mathfrak {gl}}}_V$ of Lie algebras, and hence a morphism $ \mathfrak {g}_R \to \operatorname {\mathrm {\mathfrak {gl}}}_V\bigotimes R\cong \operatorname {\mathrm {End}}_R(V_R)$ of Lie algebras over R (which is injective if $\rho $ is a closed embedding). For any $X\in \mathfrak {g}_R$ , we denote by X the action of $\operatorname {\mathrm {Lie}}(\rho )(X)$ on $V_R$ (see Remark 2.8). We define the connection $\nabla _X$ of $V_R$ associated to X by

$$ \begin{align*} \nabla_X := \nabla_{X,V} \colon V_R &\operatorname{\mathrm{\longrightarrow}} V_R \bigotimes_R \Omega_R^1,\\ v\otimes f &\longmapsto (v\otimes 1)\otimes d(f) + X(v\otimes f)\otimes dt. \end{align*} $$

Because $fdt\mapsto f$ gives an isomorphism $\Omega _R^1 \cong R$ , we have an isomorphism $\iota \colon V_R\bigotimes _R \Omega ^1_R\to V_R$ . Let $\Theta _X := \Theta _{X,V}$ be the differential operator associated to $\nabla _X$ given by the following composition:

We have that $\Theta _X(v\otimes f)=v\otimes \partial (f)+X(v\otimes f)$ for all $v\otimes f\in V_R$ .

When $G=\operatorname {\mathrm {GL}}_V$ for some $V\in \operatorname {\mathrm {\textbf {Vec}}}_F$ , we may canonically associate to any G- $(\varphi ,\nabla )$ -module $\operatorname {\mathrm {I}}$ over R a $(\varphi ,\nabla )$ -module $(V_R,\Phi ,\nabla )$ over R, where $(V_R,\Phi ,\nabla ) := \operatorname {\mathrm {I}}(V,\rho )$ and $\rho \colon G\to G$ is the identity. Choose a basis $e_1,\ldots ,e_d$ of V, we define elements $g\in G(R)$ and $X\in \mathfrak {g}_R$ by setting $g(e_i\otimes 1) := \Phi (e_i\otimes 1)$ and $X(e_i\otimes 1) := \iota \circ \nabla (e_i \otimes 1)$ , respectively. We then have $\Phi =g\varphi $ and $\nabla =\nabla _X$ .

Lemma 4.3 Let $V,W \in \operatorname {\mathrm {\textbf {Rep}}}_F(G)$ , and let $\alpha \in \operatorname {\mathrm {Hom}}_G(V,W)$ . We then have

$$ \begin{align*}\alpha_R\circ\Theta_{X,V}=\Theta_{X,W}\circ\alpha_R,\;\;\;\;\text{and}\;\;\;\;\Theta_{X,V\otimes W}=\Theta_{X,V}\otimes \operatorname{\mathrm{Id}}_{W_{\mathcal{R}}}+ \operatorname{\mathrm{Id}}_{V_{\mathcal{R}}}\otimes \Theta_{X,W}.\end{align*} $$

Proof The first equality holds, because $\alpha _R$ commutes with X (see Remark 2.8), and the second one follows from a direct computation.▪

Construction 4.4 We consider the R-algebra morphism

$$ \begin{align*}\hat\partial\colon R\operatorname{\mathrm{\longrightarrow}} R[\varepsilon],~~~ r\longmapsto r+\partial(r)\varepsilon,\end{align*} $$

which induces a morphism $G(\hat \partial )\colon G(R)\to G(R[\varepsilon ])$ . Notice that $\pi _R\circ \hat \partial =\operatorname {\mathrm {Id}}_R$ ; we then have $G(\pi _R)\circ G(\hat \partial )= \operatorname {\mathrm {Id}}_{G(R)}$ , in particular, $G(\pi _R)\big (G(\hat \partial ) (g)\big )=g$ . Identifying g with its image in $G(R[\varepsilon ])$ induced by the inclusion $R\to R[\varepsilon ],r\mapsto r$ , we then have

$$ \begin{align*}G(\hat\partial)(g)g^{-1}\in \operatorname{\mathrm{Ker}} G(\pi_R)=\mathfrak{g}_R.\end{align*} $$

For $g\in G(R)$ , we define $\partial (g):= G(\hat \partial ) (g)\in G(R[\varepsilon ])$ , and put

$$ \begin{align*}\operatorname{\mathrm{dlog}}(g):= \partial(g)g^{-1}\in \mathfrak{g}_R.\end{align*} $$

Example 4.5 Let $G=\operatorname {\mathrm {GL}}_d$ for some $d\in \mathbb {N}$ , and let $B\in G(R)$ . We have that $\operatorname {\mathrm {dlog}}(B)=I_d+\varepsilon \partial (B)B^{-1}$ , where $I_d$ is the $d\times d$ identity matrix and $\partial $ acts on B entrywise. Using the isomorphism $\operatorname {\mathrm {Lie}}(G)(R)=\{I_d+ \varepsilon B \mid B\in \operatorname {\mathrm {Mat}}_{d,d}(R)\}\cong \{B\mid B\in \operatorname {\mathrm {Mat}}_{d,d}(R)\}$ , we may identify $\operatorname {\mathrm {dlog}}(B)$ with $\partial (B)B^{-1}$ .

Definition 4.6

  1. (i) We define the gauge transformation

    $$ \begin{align*}\Gamma_g\colon \mathfrak{g}_R \operatorname{\mathrm{\longrightarrow}} \mathfrak{g}_R,\;\;\;\;\; X \longmapsto \operatorname{\mathrm{Ad}}(g)(X)-\operatorname{\mathrm{dlog}}(g),\end{align*} $$
    where $\operatorname {\mathrm {Ad}}\colon G\to \operatorname {\mathrm {GL}}_{\mathfrak {g}}$ is the adjoint representation.
  2. (ii) We define $\textbf {B}^{\varphi ,\nabla }(G,R)$ to be the groupoid whose objects are $(g,X)\in G(R)\times \mathfrak {g}_R$ satisfying $X=\Gamma _g(\operatorname {\mathrm {\boldsymbol \mu }}\varphi (X))$ , and whose morphisms $(g,X) \to (g',X')$ are elements $x\in G(R)$ such that $g'=xg \varphi (x^{-1})$ and $X'=\Gamma _x(X)$ .

Lemma 4.7 Let $(g,X)\in \textbf {B}^{\varphi ,\nabla }(G,R)$ . Then, $(V_R,g\varphi ,\nabla _X)$ is a $(\varphi ,\nabla )$ -module over R for all $V\in \operatorname {\mathrm {\textbf {Rep}}}_F(G)$ .

Proof Choose a basis $e_1,\ldots ,e_d$ for V over F where $d=\dim _F V$ . Let $A=(a_{ij})_{i,j} \in \operatorname {\mathrm {GL}}_d(R)$ (resp. $N=(n_{ij})_{i,j} \in \operatorname {\mathrm {Mat}}_{n,n}(R)$ ) be the representing matrix of $\rho (g)$ (resp. X). For any $\textbf {v}=\sum \limits _{i=1}^d e_i\otimes f_i \in V_{\mathcal {R}}$ , we compute

$$ \begin{align*} g\varphi(\Theta_X(\textbf{v})) &= g\varphi \Big(\operatorname*{\mathrm{\sum}}\limits_{i=1}^d e_i \otimes \partial(f_i) + \operatorname*{\mathrm{\sum}}\limits_{j=1}^d e_j \otimes \operatorname*{\mathrm{\sum}}\limits_{i=1}^d n_{ji}f_i \Big)\\ &=\operatorname*{\mathrm{\sum}} \limits_{j=1}^d e_j \otimes \operatorname*{\mathrm{\sum}} \limits_{i=1}^d a_{ji}\varphi(\partial(f_i)) + \operatorname*{\mathrm{\sum}}\limits_{k=1}^d e_k \otimes \operatorname*{\mathrm{\sum}} \limits_{i=1}^d \operatorname*{\mathrm{\sum}} \limits_{j=1}^d a_{kj}\varphi(n_{ji} f_i), \end{align*} $$

and

$$ \begin{align*} \Theta_X (g\varphi(\textbf{v})) &=\Theta_X \Big( \operatorname*{\mathrm{\sum}}\limits_{j=1}^d e_j \otimes \operatorname*{\mathrm{\sum}}\limits_{i=1}^d a_{ji} \varphi(f_i) \Big) \\ &=\operatorname*{\mathrm{\sum}}\limits_{j=1}^d e_j \otimes \operatorname*{\mathrm{\sum}}\limits_{i=1}^d \partial(a_{ji}) \varphi(f_i) +\operatorname*{\mathrm{\sum}}\limits_{j=1}^d e_j \otimes \operatorname*{\mathrm{\sum}}\limits_{i=1}^d a_{ji}\partial( \varphi(f_i))\\ &\quad{}+ \operatorname*{\mathrm{\sum}}\limits_{k=1}^d e_k \otimes \operatorname*{\mathrm{\sum}} \limits_{i=1}^d \operatorname*{\mathrm{\sum}} \limits_{j=1}^d n_{kj} a_{ji} \varphi(f_i). \end{align*} $$

Because $\operatorname {\mathrm {\boldsymbol \mu }}\cdot \sum \limits _{j=1}^d e_j \otimes \sum \limits _{i=1}^d a_{ji}\varphi (\partial (f_i))= \sum \limits _{j=1}^d e_j \otimes \sum \limits _{i=1}^d a_{ji}\partial ( \varphi (f_i))$ , we have that $\operatorname {\mathrm {\boldsymbol \mu }}\cdot g\varphi \circ \Theta _X= \Theta _X \circ g\varphi $ if and only if $\mu A \varphi (N)=\partial (A)+NA$ , i.e., $N=\operatorname {\mathrm {\boldsymbol \mu }} A\varphi (N)A^{-1}-\partial (A)A^{-1}$ . The last equality holds because of the assumption $X=\Gamma _g\big (\operatorname {\mathrm {\boldsymbol \mu }}\varphi (X)\big )$ , which completes the proof.▪

As a consequence, we may define a functor

(7) $$ \begin{align} \textbf{B}^{\varphi,\nabla}(G,R) \operatorname{\mathrm{\longrightarrow}} \operatorname{\mathrm{G-\textbf{Mod}^{\varphi,\nabla}_R}}, \;\;\;\;\; (g,X)\longmapsto \operatorname{\mathrm{I}}(g,X), \end{align} $$

where $\operatorname {\mathrm {I}}(g,X)(V):= (V_R,g\varphi ,\nabla _X)$ . We next show that this functor is an isomorphism. To do this, we need the following elementary descent result.

Lemma 4.8 Fix a field k, and let A and B be finitely generated k-algebras. Let $\rho \colon X\to Y$ be a closed embedding of affine algebraic k-schemes for $X=\operatorname {\mathrm {Spec}} A$ and $Y=\operatorname {\mathrm {Spec}} B$ . Let $\iota \colon S\hookrightarrow \tilde S$ be an embedding in $\operatorname {\mathrm {\textbf {Alg}}}_k$ . Suppose that we are given an element $\tilde z\in X(\tilde S)$ such that $\rho (\tilde z)\in Y(\iota (S))$ , then there exists a unique element $z\in X(S)$ such that $\tilde z=X(\iota )(z)$ .

Proof We have a diagram

with the outer triangle commutative in which $\rho ^*$ is surjective. We prove the lemma by constructing a unique k-algebra morphism $\alpha \colon A\to S$ such that $\tilde z=\iota \circ \alpha $ , as follows. For any $a\in A$ , the surjectivity of $\rho ^*$ gives us some $b\in B$ such that $\rho ^*(b)=a$ . We define $\alpha (a):= \beta (b)$ . Because $\iota $ is injective, we have $\operatorname {\mathrm {Ker}} \rho ^* \subseteq \operatorname {\mathrm {Ker}} \beta $ , which implies that $\alpha $ is well-defined. We then have $\tilde z\circ \rho ^*=\iota \circ \beta =\iota \circ \alpha \circ \rho ^*$ , which implies that $\tilde z=\iota \circ \alpha $ , because $\rho ^*$ is surjective. Moreover, $\alpha $ is a k-algebra morphism, because $\iota $ is injective and both $\iota $ and $\tilde z=\iota \circ \alpha $ are k-algebra morphisms. Finally, we see that $\alpha $ is unique, again because $\iota $ is injective.▪

Proposition 4.9 The functor $\textbf {B}^{\varphi ,\nabla }(G,R) \to \operatorname {\mathrm {G-\textbf {Mod}^{\varphi ,\nabla }_R}}$ defined in (7) is an isomorphism of categories.

Proof The proof is similar to that of [Reference Dat, Orlik and Rapoport6, Lemma 9.1.4]. We first show that the functor is fully faithful. Let $(g,X), (g',X')\in \textbf {B}^{\varphi ,\nabla }(G,R)$ . Then, any morphism $\eta \colon \operatorname {\mathrm {I}}(g,x) \to \operatorname {\mathrm {I}}(g',X')$ is an isomorphism according to [Reference Deligne and Milne7, Proposition 1.13] (see also Remark 4.2). By composing $\eta $ with the forgetful functor, we then have an automorphism of the fiber functor $\omega ^G\otimes R$ . By Corollary 2.5, this automorphism is given by a unique element $x\in G(R)$ , which then gives an isomorphism between $(g,X)$ and $(g',X')$ , as desired.

It remains to show that, for any $\operatorname {\mathrm {I}}\in \operatorname {\mathrm {G-\textbf {Mod}^{\varphi ,\nabla }_R}}$ , there exists a unique $(g,X)\in \textbf {B}^{\varphi ,\nabla }(G,R)$ such that $\operatorname {\mathrm {I}}=\operatorname {\mathrm {I}}(g,X)$ . For any $(V,\rho )\in \operatorname {\mathrm {\textbf {Rep}}}_F(G)$ , we write $\operatorname {\mathrm {I}}(V,\rho _V)=(V_R,\Phi _V,\nabla _V)$ for a $\varphi $ -linear map $\Phi _V$ and a connection $\nabla _V$ on $V_R$ . The proof consists of two steps.

Step 1: There exists a unique $X\in \mathfrak {g}_R$ such that $\nabla _V=\nabla _X$ . We write $\Theta _V$ for the composition of

where $\iota $ is induced by $fdt\mapsto f$ , and put $\theta _V:= \Theta _V-\operatorname {\mathrm {Id}}_V\otimes \partial $ . It is clear that

, where

denotes the trivial representation. Lemma 4.3 then implies that the family

$$ \begin{align*} \big\{\theta_V\colon V_R \to V_R \mid (V,\rho_V)\in \operatorname{\mathrm{\textbf{Rep}}}_F(G) \big\} \end{align*} $$

of R-linear endomorphisms satisfies conditions (i–iii) in Corollary 2.9. We thus find a unique $X\in \mathfrak {g}_R$ such that $\theta _V=\operatorname {\mathrm {Lie}}(\rho _V)(X)$ for all $(V,\rho _V)\in \operatorname {\mathrm {\textbf {Rep}}}_F(G)$ , which implies that $\nabla _V=\nabla _X$ .

Step 2: There exists a unique $g\in G(R)$ such that $\Phi _V=g\varphi $ . We put $\tilde \Phi _V := \Phi _V\otimes \varphi $ , where $\varphi $ is the Frobenius lift on $\operatorname {\mathrm {\tilde {\mathcal {R}}}}$ (in particular, $\operatorname {\mathrm {\tilde {\mathcal {R}}}}$ is viewed as an $\mathcal {R}$ -module via the $\varphi $ -equivariant embedding $\psi $ described in Section 2.3). The family

$$ \begin{align*}\big\{\lambda_V:= \tilde\Phi_V \circ (\operatorname{\mathrm{Id}}_V\otimes \varphi^{-1})\colon V_{\operatorname{\mathrm{\tilde{\mathcal{R}}}}}\to V_{\operatorname{\mathrm{\tilde{\mathcal{R}}}}} \mid V\in \operatorname{\mathrm{\textbf{Rep}}}_F(G) \big \}\end{align*} $$

of $\operatorname {\mathrm {\tilde {\mathcal {R}}}}$ -linear endomorphisms satisfies conditions (i–) in Theorem 2.4, which provides a unique element $\tilde g\in G(\operatorname {\mathrm {\tilde {\mathcal {R}}}})$ such that $\lambda _V=\rho _V(\tilde g)$ for all $(V,\rho _V) \in \operatorname {\mathrm {\textbf {Rep}}}_F(G)$ . We next reduce $\tilde g$ to an element in $G(\mathcal {R})$ . We compute

$$ \begin{align*}\tilde\Phi_V\circ (\operatorname{\mathrm{Id}}_V\otimes \varphi^{-1})(v\otimes f) =\tilde\Phi_V(v\otimes \varphi^{-1}(f))=\rho_V(\tilde g)(v\otimes f),\end{align*} $$

which implies that $\tilde \Phi _V(v\otimes f)=\rho _V(\tilde g)(v\otimes \varphi (f))$ , and hence, $\tilde \Phi _V=\tilde g \varphi $ . We now fix a d-dimensional faithful representation $(V,\rho _V)$ , and an F-basis $e_1,\ldots ,e_d$ for V. Suppose that $\Phi _V(e_i)=\sum \limits _{j=1}^d a_{ji}e_j$ , where $a_{ij}\in R$ for all $1\leq i,j \leq d$ . Put $A=(a_{ij})_{i,j} \in \operatorname {\mathrm {GL}}_d(R)$ . Then, $\psi (A)=(\psi (a_{ij}))_{i,j}\in \operatorname {\mathrm {GL}}_d(\operatorname {\mathrm {\tilde {\mathcal {R}}}})$ describes the $\varphi $ -linear action of $\tilde \Phi _V$ as well as the $\operatorname {\mathrm {\tilde {\mathcal {R}}}}$ -linear action $\rho (\tilde g)$ in the basis $e_1\otimes 1,\ldots , e_d \otimes 1$ . By replacing X with G, Y with $\operatorname {\mathrm {GL}}_d$ , S with R, $\tilde S$ with $\operatorname {\mathrm {\tilde {\mathcal {R}}}}$ , and $\iota $ with $\psi $ in Lemma 4.8, we find a unique element $g\in G(R)$ such that $\psi (g)=\tilde g$ . It follows that $\Phi _V=g\varphi $ , as desired.▪

Example 4.10 Let $d\in \mathbb {N}$ . The affine algebraic F-group $\operatorname {\mathrm {SL}}_d$ is defined by

$$ \begin{align*}\operatorname{\mathrm{SL}}_d(S)=\{ A\in \operatorname{\mathrm{GL}}_d(S) \mid \det(A)=1\}\end{align*} $$

for all $S\in \operatorname {\mathrm {\textbf {Alg}}}_F$ , whose Lie algebra $\operatorname {\mathrm {\mathfrak {sl}}}_d$ consists of $d\times d$ matrices with entries in F and with trace zero.

  1. (i) We claim that any pair $(A,N) \in \operatorname {\mathrm {SL}}_d(\mathcal {R}) \times \operatorname {\mathrm {Mat}}_{d,d}(\mathcal {E}^{\dagger })$ satisfying $N=\operatorname {\mathrm {\boldsymbol \mu }} A \varphi (N) A^{-1}-\partial (A)A^{-1}$ is already an object in $\textbf {B}^{\varphi ,\nabla }(\operatorname {\mathrm {SL}}_d,\mathcal {R})$ . It is equivalent to showing that the trace $\operatorname {\mathrm {Tr}}(N)$ of N is zero. Recall that the Frobenius lift $\varphi $ on $\mathcal {E}^{\dagger }$ is given by $\varphi \big (\sum \limits _{i\in \mathbb {Z}} c_it^i \big )=\sum \limits _{i\in \mathbb {Z}} \varphi (c_i) u^i$ , where $u=\varphi (t)$ satisfies $|u-t^q|_1<1$ . If we write $u=\sum \limits _{i\in \mathbb {Z}} u_it^i, u_i\in K$ , we then have $|u_j|<1$ for all $j \neq q$ and $|u_q|=1$ . It follows that $|\operatorname {\mathrm {\boldsymbol \mu }}|_1=|\partial (u)|_1=|\sum \limits _{i\in \mathbb {Z}} i u_i t^{i-1}|_1 <1$ . On the other hand, we have $\operatorname {\mathrm {Tr}}\big (\partial (A)A^{-1} \big )=0$ , because $\partial (A)A^{-1} \in \operatorname {\mathrm {\mathfrak {sl}}}_{d,\mathcal {R}}$ (see Construction 4.4). Assume, to the contrary, that $\operatorname {\mathrm {Tr}}(N)\neq 0$ , we have

    $$ \begin{align*} |\operatorname{\mathrm{Tr}}(N)|_1= |\operatorname{\mathrm{\boldsymbol\mu}} \operatorname{\mathrm{Tr}}(\varphi(N))|_1 = |\operatorname{\mathrm{\boldsymbol\mu}} \varphi(\operatorname{\mathrm{Tr}}(N))|_1 < |\varphi(\operatorname{\mathrm{Tr}}(N))|_1= |\operatorname{\mathrm{Tr}}(N)|_1, \end{align*} $$
    a contradiction (we have the last equality, because $\varphi $ preserves the $1$ -Gauss norm on $\mathcal {E}^{\dagger }$ ).
  2. (ii) We use the Bessel isocrystal as described in [Reference Kedlaya12, Example 20.2.1] (see also [Reference Kedlaya9, Section 1.5] and [Reference Tsuzuki24, Example 6.2.6]) to construct an object in $\textbf {B}^{\varphi ,\nabla }(\operatorname {\mathrm {SL}}_2,\mathcal {R})$ . We first briefly recall the Bessel isocrystal. In Hypothesis 2.1, we let $q=p$ be an odd prime, $\kappa =\mathbb {F}_p$ , and $F=\mathbb {Q}_p(\pi )$ , where $\pi $ is a $(p-1)$ st root of $-p$ in $\bar {\mathbb {Q}}_p$ . Then, the (p-power) Frobenius on $K=F$ is the identity. Let $\varphi $ be the Frobenius lift on $\mathcal {R}$ given by $\varphi (t)=t^p$ . Then, [Reference Kedlaya12, Example 20.2.1] gives rise to a pair $(A_0,N_0)\in \operatorname {\mathrm {GL}}_2(\mathcal {R}) \times \operatorname {\mathrm {Mat}}_{2,2}(\mathcal {E}^{\dagger })$ with $\det (A_0)=p$ satisfying the gauge compatibility condition, in which $N_0= \left ( \begin {smallmatrix} 0 & t^{-1}\\ \pi ^2 t^{-2} & 0 \end {smallmatrix}\right ) \in \operatorname {\mathrm {\mathfrak {sl}}}_{2,\mathcal {E}^{\dagger }}.$ We now assume that $p \equiv 1 (\operatorname {\mathrm {mod}} 4)$ , and $\textbf {i}$ is a square root of $-1$ in $\mathbb {Q}_p$ . Because $p-1$ is even, we may set $\alpha := \frac {\textbf {i}}{\pi ^{(p-1)/2}}\in F^{\times }$ . We then have $\alpha ^2=p^{-1}=\det (A_0)^{-1}$ . Put $D_0 = \left ( \begin {smallmatrix} 0 & 1\\ \alpha & 0 \end {smallmatrix}\right ) \in \operatorname {\mathrm {GL}}_2(F).$ Then, $D_0A_0D_0\in \operatorname {\mathrm {SL}}_2(\mathcal {R})$ . Moreover, we see that $D_0N_0D_0^{-1}=D_0^{-1}N_0 D_0 \in \operatorname {\mathrm {\mathfrak {sl}}}_{2,\mathcal {E}^{\dagger }}$ . Put $A := D_0A_0 D_0$ and $N := D_0 N_0 D_0^{-1}$ . Then, a straightforward verification shows $N=\operatorname {\mathrm {\boldsymbol \mu }} A \varphi (N)A^{-1}-\partial (A)A^{-1}$ (noting that $\varphi (D_0)=D_0$ and $\partial (D_0)=0$ ). We thus have $(A,N)\in \textbf {B}^{\varphi ,\nabla }(\operatorname {\mathrm {SL}}_2,\mathcal {R})$ , as desired.

  3. (iii) Let $(A,N) \in \textbf {B}^{\varphi ,\nabla }(\operatorname {\mathrm {SL}}_d,\mathcal {R})$ . We show that $(A,N)$ is“ $\operatorname {\mathrm {SL}}_d$ -quasi-unipotent” (as described in the introduction) by modifying the classical monodromy as follows. By the classical pLMT, we find a finite separable extension L of $\kappa (\!( t )\!)$ and $B\in \operatorname {\mathrm {GL}}_d(\mathcal {R}_L)$ such that $BNB^{-1}-\partial (B)B^{-1}$ has trace zero being an upper-triangular block matrix with zero blocks in the diagonal. We wish to replace B with an element in $\operatorname {\mathrm {SL}}_d(\mathcal {R}_L)$ . To this end, we deduce first that $\operatorname {\mathrm {Tr}}(\partial (B)B^{-1})= \operatorname {\mathrm {Tr}}(BNB^{-1})=\operatorname {\mathrm {Tr}}(N)=0$ . It then follows from Jacobi’s formula that $\partial (\det (B))= \det (B) \cdot \operatorname {\mathrm {Tr}}(B^{-1} \partial (B))=0$ . Put $D := \operatorname {\mathrm {Diag}}( \det (B)^{-1},1,\ldots ,1)$ . Then, $DB\in \operatorname {\mathrm {SL}}_d(\mathcal {R}_L)$ and $\partial (D)=0$ . We then have

    $$ \begin{align*} (DB) N (DB)^{-1}- \partial(DB)(DB)^{-1} = D \big( B N B^{-1} -\partial(B)B^{-1} \big) D^{-1}, \end{align*} $$
    which is an upper-triangular block matrix with zero blocks, and the sizes of the blocks are the same as those in $B N B^{-1} -\partial (B)B^{-1}$ (the said properties are preserved under conjugation by a diagonal matrix). Hence, $DB$ is a desired replacement of B and we are done.

Example 4.11 For any matrix X, we denote by $X^{\operatorname {\mathrm {T}}}$ its transpose, and by $X^{-\operatorname {\mathrm {T}}}$ the inverse of transpose if X is invertible.

We fix the skew-symmetric matrix $J=\left (\begin {smallmatrix} & & & 1 \\ & & -1 & \\ & 1 & &\\ -1 & & & \end {smallmatrix}\right )$ . The affine algebraic F-group $\operatorname {\mathrm {Sp}}_4$ is defined by

$$ \begin{align*} \operatorname{\mathrm{Sp}}_4(S) := \{ A\in \operatorname{\mathrm{GL}}_4(S) \mid A^{-1}=J^{-1} A^{\operatorname{\mathrm{T}}} J\}, \end{align*} $$

for all $S\in \operatorname {\mathrm {\textbf {Alg}}}_F$ . We denote by $\operatorname {\mathrm {\mathfrak {sp}}}_4$ the Lie algebra of $\operatorname {\mathrm {Sp}}_4$ . For any $S\in \operatorname {\mathrm {\textbf {Alg}}}_F$ , we then have $\operatorname {\mathrm {\mathfrak {sp}}}_{4,S}=\{X\in \operatorname {\mathrm {Mat}}_{4,4}(S) \mid X=J X^{\operatorname {\mathrm {T}}} J\}$ . We remark that the specific choice of J preserves Borel subgroups under conjugation, which will be useful in the monodromy considered below.

Given any $(\varphi ,\nabla )$ -module over $\mathcal {R}$ of rank $2$ , e.g., the Bessel isocrystal described above, we obtain a pair $(A_0,N_0)\in \operatorname {\mathrm {GL}}_2(\mathcal {R}) \times \operatorname {\mathrm {Mat}}_{2,2}(\mathcal {R})$ satisfying $N_0=\operatorname {\mathrm {\boldsymbol \mu }} A_0\varphi (N_0)A_0^{-1}-\partial (A_0)A_0^{-1}$ . Put

$$ \begin{align*} A := \begin{pmatrix} A_0 & 0 \\ 0 & \left(\begin{smallmatrix} & 1\\ -1 & \end{smallmatrix}\right)^{-1} A_0^{-\operatorname{\mathrm{T}}} \left(\begin{smallmatrix} & 1\\ -1 & \end{smallmatrix}\right) \end{pmatrix} ~~~\text{and}~~~ N := \begin{pmatrix} N_0 &0 \\ 0 & \left(\begin{smallmatrix} & 1\\ -1 & \end{smallmatrix}\right) N_0^{\operatorname{\mathrm{T}}} \left(\begin{smallmatrix} & 1\\ -1 & \end{smallmatrix}\right) \end{pmatrix}. \end{align*} $$

A straightforward verification shows that $A\in \operatorname {\mathrm {Sp}}_4(\mathcal {R}),N\in \operatorname {\mathrm {\mathfrak {sp}}}_{4,\mathcal {R}}$ , and, moreover, $N=\operatorname {\mathrm {\boldsymbol \mu }} A \varphi (N)A^{-1}-\partial (A)A^{-1}$ (noting that $\left (\begin {smallmatrix} 0 & 1\\ -1 & 0 \end {smallmatrix}\right )^{-1} =-\left (\begin {smallmatrix} 0 & 1\\ -1 & 0\end {smallmatrix} \right )$ ), which implies that $(A,N) \in \textbf {B}^{\varphi ,\nabla }(\operatorname {\mathrm {Sp}}_4,\mathcal {R})$ .

We next show that $(A,N)$ is “ $\operatorname {\mathrm {Sp}}_4$ -quasi-unipotent.” By the classical pLMT, we find a finite separable extension L of $\kappa (\!( t )\!)$ and $B_0 \in \operatorname {\mathrm {GL}}_2(\mathcal {R}_L)$ such that

$$ \begin{align*} B_0 N_0 B_0^{-1}-\partial(B_0)B_0^{-1}= \left( \begin{smallmatrix} 0 & n\\ 0 &0 \end{smallmatrix}\right), \end{align*} $$

for some $n\in \mathcal {R}_L$ (n could be $0$ ). Put

$$ \begin{align*} B := \begin{pmatrix} B_0 & 0 \\ 0 & \left(\begin{smallmatrix} & 1\\ -1 & \end{smallmatrix}\right)^{-1} B_0^{-\operatorname{\mathrm{T}}} \left(\begin{smallmatrix} & 1\\ -1 & \end{smallmatrix}\right) \end{pmatrix}. \end{align*} $$

We then have $B\in \operatorname {\mathrm {Sp}}_4(\mathcal {R}_L)$ , and

$$ \begin{align*} BNB^{-1}-\partial(B)B^{-1}= \left(\begin{smallmatrix} 0 & n & 0 & 0\\ 0 &0 & 0 & 0\\ 0 &0 & 0 & n\\ 0 &0 & 0 & 0 \end{smallmatrix}\right), \end{align*} $$

again by straightforward computations.

4.2 The pushforward functor

Let $R\in \{\mathcal {E}^{\dagger },\mathcal {R}\}$ . For any $g\in G(R)$ and $n\in \mathbb {N}$ , we define

$$ \begin{align*}[n]_*(g):= g\varphi(g)\cdots\varphi^{n-1}(g)\in G(R),\end{align*} $$

the n-pushforward of g. Notice that $[n]_*(g)\varphi ^n=(g\varphi )^n\in G(R)\rtimes \operatorname {\mathrm {\langle }}\varphi \operatorname {\mathrm {\rangle }}$ for all $n\in \mathbb {N}$ .

We define the n-pushforward functor by

$$ \begin{align*} [n]_* \colon \textbf{B}^{\varphi,\nabla}(G,R) \operatorname{\mathrm{\longrightarrow}} {\textbf{B}}^{\varphi^n,\nabla}(G,R), \;\;\;\;\; (g,X) \longmapsto \big([n]_*(g), X \big),\end{align*} $$

and $[n]_*(x)=x$ for all morphisms $x\in \textbf {B}^{\varphi ,\nabla }(G,R)$ . The following lemma shows that this functor is well-defined (in particular, faithful).

Lemma 4.12 Let $(g,X)\in \textbf {B}^{\varphi ,\nabla }(G,R)$ . We then have $\big ([n]_*(g), X \big )\in \textbf {B}^{\varphi ^n,\nabla }(G,R)$ for all $n\in \mathbb {N}$ .

Proof We show by induction on n that

$$ \begin{align*}X+\operatorname{\mathrm{dlog}}\big([n]_*(g)\big)=\operatorname{\mathrm{\boldsymbol\mu}}(\varphi^n,t)\operatorname{\mathrm{Ad}}\big([n]_*(g)\big)\big(\varphi^{n}(X)\big).\end{align*} $$

There is nothing to show when $n=1$ . We now assume by the induction hypothesis that

$$ \begin{align*}X+\operatorname{\mathrm{dlog}}\big([n-1]_*(g)\big)=\operatorname{\mathrm{\boldsymbol\mu}}(\varphi^{n-1},t)\operatorname{\mathrm{Ad}}\big([n-1]_*(g)\big)\big( \varphi^{n-1}(X)\big),\end{align*} $$

We notice that $\operatorname {\mathrm {\boldsymbol \mu }}(\varphi ^{n-1},t)=\operatorname {\mathrm {\boldsymbol \mu }}\varphi (\operatorname {\mathrm {\boldsymbol \mu }})\cdots \varphi ^{n-2}(\operatorname {\mathrm {\boldsymbol \mu }})$ , and hence,

$$ \begin{align*}\partial(\varphi^{n-1}(f))=\operatorname{\mathrm{\boldsymbol\mu}}\varphi(\operatorname{\mathrm{\boldsymbol\mu}})\cdots\varphi^{n-2}(\operatorname{\mathrm{\boldsymbol\mu}})\varphi^{n-1}(\partial(f))=\operatorname{\mathrm{\boldsymbol\mu}}(\varphi^{n-1},t)\varphi^{n-1}(\partial(f)),~~~~~\forall f\in R,\end{align*} $$

which implies that

$$ \begin{align*}\operatorname{\mathrm{dlog}}(\varphi^{n-1}(g))=\operatorname{\mathrm{\boldsymbol\mu}}(\varphi^{n-1},t) \varphi^{n-1}(\operatorname{\mathrm{dlog}}(g)).\end{align*} $$

On the other hand, because $X+\operatorname {\mathrm {dlog}}(g)=\operatorname {\mathrm {\boldsymbol \mu }} \operatorname {\mathrm {Ad}}(g)(\varphi (X))$ , we have

$$ \begin{align*}\varphi^{n-1}(X)+\varphi^{n-1}(\operatorname{\mathrm{dlog}}(g))= \varphi^{n-1}(\operatorname{\mathrm{\boldsymbol\mu}})\operatorname{\mathrm{Ad}}\big(\varphi^{n-1}(g)\big)\big(\varphi^n(X)\big).\end{align*} $$

We now compute

$$ \begin{align*} X+\operatorname{\mathrm{dlog}}\big([n]_*(g)\big) &=X+\operatorname{\mathrm{dlog}}\big([n-1]_*(g)\big)+\operatorname{\mathrm{Ad}}\big([n-1]_*(g)\big)\big(\operatorname{\mathrm{dlog}}(\varphi^{n-1}(g))\big)\\ &=\operatorname{\mathrm{\boldsymbol\mu}}(\varphi^{n-1},t)\operatorname{\mathrm{Ad}}\big([n-1]_*(g)\big) \big(\varphi^{n-1}(X)\big)\\ &\;\;\;\; +\operatorname{\mathrm{\boldsymbol\mu}}(\varphi^{n-1},t)\operatorname{\mathrm{Ad}}\big([n-1]_*(g)\big) \big(\varphi^{n-1}(\operatorname{\mathrm{dlog}}(g)) \big)\\ &=\operatorname{\mathrm{\boldsymbol\mu}}(\varphi^{n-1},t) \operatorname{\mathrm{Ad}}\big([n-1]_*(g)\big) \big(\varphi^{n-1}(X)+\varphi^{n-1}(\operatorname{\mathrm{dlog}}(g)) \big)\\ &= \operatorname{\mathrm{\boldsymbol\mu}}(\varphi^{n-1},t)\operatorname{\mathrm{Ad}}\big([n-1]_*(g)\big) \big(\varphi^{n-1}(\operatorname{\mathrm{\boldsymbol\mu}})\operatorname{\mathrm{Ad}}\big(\varphi^{n-1}(g)\big)\big(\varphi^n(X)\big)\big)\\ &=\operatorname{\mathrm{\boldsymbol\mu}}(\varphi^n,t) \operatorname{\mathrm{Ad}}\big([n]_*(g)\big)\big(\varphi^{n}(X)\big), \end{align*} $$

which proves the lemma.▪

In connection with the pushforward functor on $\varphi $ -modules as recalled in Section 2.3, we state the following lemma resulting from [Reference Kedlaya11, Lemma 1.6.3 and Remark 1.7.2], which will not be explicitly used in the sequel.

Lemma 4.13 Let $g\in G(R)$ . Then, $(V_R,g\varphi )$ is pure of slope $\mu $ if and only if $(V_R,[n]_*(g)\varphi ^n)$ is pure of slope $n\mu $ for all $n\in \mathbb {N}$ . Moreover, if $(V_R,g\varphi )$ has jumps $\mu _1,\ldots ,\mu _l$ , then $(V_R,[n]_*(g)\varphi ^n)$ has jumps $n\mu _1,\ldots ,n\mu _l$ .

4.3 G- $\varphi $ -modules attached to splittings

Let $g\in G(\mathcal {R})$ . We fix a splitting $\xi _g$ of $\operatorname {\mathrm {{HN}}}_g$ by Theorem 3.12.

Construction 4.14 Let $(V_{\mathcal {R}},g\varphi ,\nabla _X)$ be a $(\varphi ,\nabla )$ -module over $\mathcal {R}$ with the slope filtration

$$ \begin{align*}0\subseteq V_{\mathcal{R}}^{\mu_1} \subseteq \cdots \subseteq V_{\mathcal{R}}^{\mu_l}=V_{\mathcal{R}},\end{align*} $$

with jumps $\mu _1<\cdots <\mu _l$ . Then, $\xi _g(V)$ is the decomposition

$$ \begin{align*} V_{\mathcal{R}}=\operatorname*{\mathrm{\bigoplus}}\limits_{i=1}^l V_{\mathcal{R},\mu_i}\end{align*} $$

of $\mathcal {R}$ -modules such that $\bigoplus \limits _{i=1}^j V_{\mathcal {R},\mu _i} =V_{\mathcal {R}}^{\mu _j}$ for $j=1,\ldots ,l$ .

  1. (i) For any $1\leq j\leq l$ and $\textbf {v}\in V_{\mathcal {R},\mu _j}$ , we have $\Phi _g(\textbf {v})\in V_{\mathcal {R}}^{\mu _j}$ , whence a unique expression $\Phi _g(\textbf {v})=\operatorname *{\mathrm {\sum }}\limits _{i=1}^j \textbf {v}_i$ with $\textbf {v}_i\in V_{\mathcal {R},\mu _i}$ . We thus have a $\varphi $ -linear map

    $$ \begin{align*}\Phi_{g,\mu_j} \colon V_{\mathcal{R},\mu_j} \operatorname{\mathrm{\longrightarrow}} V_{\mathcal{R},\mu_j},~~~~ \textbf{v} \longmapsto \textbf{v}_j.\end{align*} $$
    We then define $\Phi _g':= \bigoplus \limits _{i=1}^l \Phi _{g,{\mu _i}}$ . We define
    $$ \begin{align*}\operatorname{\mathrm{I}}'(g)(V):= (V_{\mathcal{R}},\Phi_g').\end{align*} $$
    For a morphism $\alpha \colon V\to W$ of finite-dimensional G-modules, we define $\operatorname {\mathrm {I}}'(g)(\alpha ):= \alpha _{\mathcal {R}}$ .
  2. (ii) Similarly, for any $1\leq j\leq l$ and $\textbf {v}\in V_{\mathcal {R},\mu _j}$ , we have $\Theta _X(\textbf {v})\in V_{\mathcal {R}}^{\mu _j}$ , whence a unique expression $\Theta _X(\textbf {v})=\sum _{i=1}^j \textbf {v}_i$ with $\textbf {v}_i\in V_{\mathcal {R},\mu _i}$ . We thus have a K-linear differential operator

    $$ \begin{align*}\Theta_{X,\mu_j} \colon V_{\mathcal{R},\mu_j} \operatorname{\mathrm{\longrightarrow}} V_{\mathcal{R},\mu_j},~~~~ \textbf{v} \longmapsto \textbf{v}_j.\end{align*} $$
    We then define $\Theta _X':= \bigoplus \limits _{i=1}^l \Theta _{X,{\mu _i}}$ .

Notice that $\big ( V_{\mathcal {R},\mu _1},\Phi _{g,{\mu _1}}\big )=\big ( V_{\mathcal {R}}^{\mu _1},\Phi _g|_{V_{\mathcal {R}}^{\mu _1}} \big )$ , and $\big ( V_{\mathcal {R},\mu _i},\Phi _{g,{\mu _i}}\big )$ is isomorphic to $V_{\mathcal {R}}^{\mu _i}/V_{\mathcal {R}}^{\mu _{i-1}}$ as $\varphi $ -modules for $2\leq i\leq l$ . Similarly, we have $\big ( V_{\mathcal {R},\mu _1},\Theta _{X,{\mu _1}} \big )=\big ( V_{\mathcal {R}}^{\mu _1},\Theta _X|_{V_{\mathcal {R}}^{\mu _1}} \big )$ , and $\big ( V_{\mathcal {R},\mu _i},\Theta _{X,\mu _i}\big )$ is isomorphic to $V_{\mathcal {R}}^{\mu _i}/V_{\mathcal {R}}^{\mu _{i-1}}$ as a differential module for $2\leq i\leq l$ .

The remainder of this subsection is devoted to the consequences of Construction 4.14 (i). We will continue to discuss (ii) in Section 4.4; we will show, in particular, that both constructions assemble to give a G- $(\varphi ,\nabla )$ -module over $\mathcal {R}$ .

Lemma 4.15 $\operatorname {\mathrm {I}}'(g) \colon \operatorname {\mathrm {\textbf {Rep}}}_F(G) \to \varphi \text {-}\operatorname {\mathrm {\textbf {Mod}}}_{\mathcal {R}}$ is a G- $\varphi $ -module over $\mathcal {R}$ .

Proof By Definition 3.1, it amounts to show that $\operatorname {\mathrm {I}}'(g)$ is an exact faithful F-linear tensor functor. In this proof, we fix $V,W\in \operatorname {\mathrm {\textbf {Rep}}}_F(G)$ , and suppose the slope filtration of $(V_{\mathcal {R}},g\varphi )$ (resp. of $(W_{\mathcal {R}},g\varphi )$ ) has jumps $\mu _1<\cdots <\mu _{l_V}$ (resp. $\nu _1<\cdots <\nu _{l_W}$ ).

We first check the functoriality of $\operatorname {\mathrm {I}}'(g)$ (the exactness, faithfulness, and F-linearity will follow immediately). Given $\alpha \in \operatorname {\mathrm {Hom}}_G(V,W)$ , we need to show that

$$ \begin{align*}\alpha_{\mathcal{R}}\circ \Phi_g'=\Phi_g'\circ \alpha_{\mathcal{R}}.\end{align*} $$

For any fixed $1\leq l \leq {l_V}$ , we have that $\alpha _{\mathcal {R}}(V_{\mathcal {R},\mu _l})\subseteq W_{\mathcal {R},\mu _l}$ by Theorem 3.12. Notice that $W_{\mathcal {R},\mu _l}=W_{\mathcal {R},\nu _s}$ if $\mu _l=\nu _s$ for some $1\leq s\leq {l_W}$ , and $W_{\mathcal {R},\mu _l}=0$ otherwise. In the latter case, it is clear that $\alpha _{\mathcal {R}}\circ \Phi _g'=\Phi _g'\circ \alpha _{\mathcal {R}}=0$ on $V_{\mathcal {R},\mu _l}$ , and we are done. Suppose now we are in the former case. Let $\textbf {v}$ be a nonzero element in $V_{\mathcal {R},\mu _l}$ . We then have $\Phi _g(\textbf {v})\in V_{\mathcal {R}}^{\mu _l}$ and $\alpha _{\mathcal {R}}(\textbf {v})\in W_{\mathcal {R},\nu _s}$ . We have unique expressions

$$ \begin{align*} \Phi_g(\textbf{v})= \operatorname*{\mathrm{\sum}}\limits_{i=1}^l \textbf{v}_i,~~~~~\textbf{v}_i\in V_{\mathcal{R},\mu_i},\end{align*} $$

and

$$ \begin{align*} \alpha_{\mathcal{R}}\circ\Phi_g(\textbf{v})=\operatorname*{\mathrm{\sum}}\limits_{i=1}^s \textbf{w}_i,~~~~~\textbf{w}_i\in W_{\mathcal{R},\nu_i};\end{align*} $$

therefore $\alpha _{\mathcal {R}}(\textbf {v}_l)=\textbf {w}_s$ . We also write

$$ \begin{align*} \Phi_g\circ \alpha_{\mathcal{R}}(\textbf{v})=\operatorname*{\mathrm{\sum}}\limits_{i=1}^s \textbf{w}_i', \;\;\;\textbf{w}_i'\in W_{\mathcal{R},\nu_i};\end{align*} $$

we then have $\textbf {w}_i=\textbf {w}_i'$ for $i=1,\ldots ,s$ , as $\alpha _{\mathcal {R}}\circ \Phi _g=\Phi _g\circ \alpha _{\mathcal {R}}$ . We thus have $\alpha _{\mathcal {R}}\circ \Phi _{g,{\mu _l}}(\textbf {v})=\alpha _{\mathcal {R}}(\textbf {v}_l)=\textbf {w}_s$ and $\Phi _{g,{\nu _s}}\circ \alpha _{\mathcal {R}}(\textbf {v})= \textbf {w}_s'=\textbf {w}_s$ , which implies that $\alpha _{\mathcal {R}}\circ \Phi _{g,{\mu _l}}=\Phi _{g,{\nu _s}}\circ \alpha _{\mathcal {R}}$ , as desired.

It remains to show that $\operatorname {\mathrm {I}}'(g)$ preserves tensor products. Because $\tau _g$ is a tensor functor, the $(\mu _l+\nu _s)$ th graded piece of $\tau _g(V\otimes W)$ is then

$$ \begin{align*} \big(V\bigotimes\limits_F W \big)_{\mathcal{R},\mu_l+\nu_s}= \operatorname*{\mathrm{\bigoplus}}\limits_{\scriptstyle \begin{array}{c}\mu_i+\nu_j=\mu_l+\nu_s \\ 1\leq i\leq {l_V}, 1\leq j\leq {l_W}\end{array}} \big( V_{\mathcal{R},\mu_i}\bigotimes\limits_{\mathcal{R}} W_{\mathcal{R},\nu_j } \big),\end{align*} $$

for all $1\leq l\leq l_V$ and $1\leq s \leq l_W$ . It then follows from Construction 4.14(i) that

$$ \begin{align*} \Phi^{\prime}_{g,\mu_l+\nu_s}=\operatorname*{\mathrm{\bigoplus}}\limits_{\scriptstyle \begin{array}{c} \mu_i+\nu_j=\mu_l+\nu_s \\ 1\leq i\leq {l_V},1\leq j\leq {l_W}\end{array}} \big(\Phi^{\prime}_{g,\mu_i} \otimes \Phi^{\prime}_{g,\nu_j}\big),\end{align*} $$

which implies that $\operatorname {\mathrm {I}}'(g)(V\bigotimes W)$ coincides with $\operatorname {\mathrm {I}}'(g)(V)\bigotimes \operatorname {\mathrm {I}}'(g)(W)$ on all $(V\bigotimes W)_{\mathcal {R},\mu _l+\nu _s}$ , whence on $(V\bigotimes W)_{\mathcal {R}}$ . This completes the proof.▪

With Lemma 4.15, we imitate Step 2 in the proof of Proposition 4.9 and have the following proposition.

Proposition 4.16 There exists a unique element $z\in G(\mathcal {R})$ such that $\operatorname {\mathrm {I}}'(g)=\operatorname {\mathrm {I}}(z)$ .

4.4 G- $(\varphi ,\nabla )$ -modules attached to splittings

We fix $(g,X)\in \textbf {B}^{\varphi ,\nabla }(G,\mathcal {R})$ . We also fix a splitting $\xi _g$ of $\operatorname {\mathrm {{HN}}}_g$ given by Theorem 3.12.

We now look back at Construction 4.14(ii). We claim that $\Theta _X'-\operatorname {\mathrm {Id}}_V\otimes \partial \colon V_R \to V_R$ is R-linear for all $(V,\rho _V)\in \operatorname {\mathrm {\textbf {Rep}}}_F(G)$ . Let $1\leq j\leq l$ and let $v\otimes f\in V_{\mathcal {R},\mu _j}$ . Suppose that $\Theta _X(v\otimes f)=\operatorname *{\mathrm {\sum }}_{i=1}^j \textbf {v}_i$ with $\textbf {v}_i\in V_{\mathcal {R},\mu _i}$ . Then, $\Theta _X'(v\otimes f)=\textbf {v}_j$ by construction. Let $f'\in R$ . We compute

$$ \begin{align*} \Theta^{\prime}_X(v\otimes ff') &=v\otimes \partial(f)f'+v\otimes f\partial(f')+X (v\otimes ff')\\ &= \big( v\otimes \partial(f)+ X(v\otimes f) \big)f'+v\otimes f\partial(f')\\ &= \Theta^{\prime}_X(v\otimes f)f' +v\otimes f\partial(f')\\ &=f' \operatorname*{\mathrm{\sum}}_{i=1}^j \textbf{v}_i+v\otimes f\partial(f'), \end{align*} $$

which implies that $\Theta ^{\prime }_X(v\otimes ff')=f'\textbf {v}_j+v\otimes f\partial (f')$ . We thus have

$$ \begin{align*} (\Theta_X'-\operatorname{\mathrm{Id}}_V\otimes \partial)(v\otimes ff') &=f'\textbf{v}_j+v\otimes f\partial(f')- v\otimes \partial(ff')\\ &= f'\textbf{v}_j+v\otimes f\partial(f')-v\otimes \partial(f)f'-v\otimes f\partial(f')\\ &=f' (\textbf{v}_j-v\otimes \partial(f))\\ &= f' (\Theta_X'-\operatorname{\mathrm{Id}}_V\otimes \partial)(v\otimes f), \end{align*} $$

as desired.

The following proposition (and its proof) is analogous to Lemma 4.15.

Proposition 4.17 There exists a unique element $X_0\in \mathfrak {g}_{\mathcal {R}}$ such that $\Theta _X'=\Theta _{X_0}$ .

Proof For any $(V,\rho _V)\in \operatorname {\mathrm {\textbf {Rep}}}_F(G)$ , we define $\theta _V:= \Theta _X'-\operatorname {\mathrm {Id}}_V\otimes \partial $ . We claim that the family

$$ \begin{align*} \big\{\theta_V\colon V_{\mathcal{R}}\to V_{\mathcal{R}} \mid (V,\rho_V)\in \operatorname{\mathrm{\textbf{Rep}}}_F(G) \big\} \end{align*} $$

of R-linear endomorphisms satisfies conditions (i–iii) in Corollary 2.9. The lemma will follow immediately.

It is clear that $\theta _V=0$ if $V=F$ is the trivial G-representation. For the remainder of the proof, we fix $(V,\rho _V),(W,\rho _W)\in \operatorname {\mathrm {\textbf {Rep}}}_F(G)$ , and suppose the slope filtration of $(V_{\mathcal {R}},g\varphi )$ (resp. of $(W_{\mathcal {R}},g\varphi )$ ) has jumps $\mu _1<\cdots <\mu _{l_V}$ (resp. $\nu _1<\cdots <\nu _{l_W}$ ). Let $\alpha \in \operatorname {\mathrm {Hom}}_G(V,W)$ . To show that $\theta _V\circ \alpha _{\mathcal {R}} =\alpha _{\mathcal {R}}\circ \theta _W$ , it suffices to show that $\Theta _X'\circ \alpha _{\mathcal {R}} =\alpha _{\mathcal {R}} \circ \Theta _X'$ . Notice that $\alpha _{\mathcal {R}}$ respects gradings. Replacing $\Phi _g$ with $\Theta _X$ (possibly with proper decorations) in the second paragraph of the proof of Lemma 4.15, we have the desired result.

It remains to show that

$$ \begin{align*}\theta_{V\otimes W} =\theta_V\otimes \operatorname{\mathrm{Id}}_{W_{\mathcal{R}}} +\operatorname{\mathrm{Id}}_{V_{\mathcal{R}}} \otimes \theta_W.\end{align*} $$

Because $\tau _g$ is a tensor functor, the $(\mu _l+\nu _s)$ th graded piece of $\tau _g(V\bigotimes W)$ is then

$$ \begin{align*} \big(V \bigotimes W \big)_{\mathcal{R},\mu_l+\nu_s}= \operatorname*{\mathrm{\bigoplus}}\limits_{\scriptstyle \begin{array}{c}\mu_i+\nu_j=\mu_l+\nu_s \\ 1\leq i\leq {l_V}, 1\leq j\leq {l_W}\end{array}} \big( V_{\mathcal{R},\mu_i}\bigotimes_{\mathcal{R}} W_{\mathcal{R},\nu_j } \big),\end{align*} $$

for all $1\leq l\leq l_V$ and $1\leq s \leq l_W$ . It follows from Lemma 4.3 and Construction 4.14 that

$$ \begin{align*} \Theta^{\prime}_{X,\mu_l+\nu_s}=\operatorname*{\mathrm{\bigoplus}}\limits_{\scriptstyle \begin{array}{c}\mu_i+\nu_j=\mu_l+\nu_s \\ 1\leq i\leq {l_V}, 1\leq j\leq {l_W}\end{array}} \big(\Theta^{\prime}_{X,\mu_i}\otimes \operatorname{\mathrm{Id}}_{W_{\mathcal{R},\nu_j }}+ \operatorname{\mathrm{Id}}_{V_{\mathcal{R},\mu_i }} \otimes \Theta^{\prime}_{X,\nu_j} \big).\end{align*} $$

Let $v\otimes f\otimes w\otimes f'\in V_{\mathcal {R},\mu _i}\bigotimes _{\mathcal {R}} W_{\mathcal {R},\nu _j }$ . We compute

$$ \begin{align*} & \big( \theta_V\otimes \operatorname{\mathrm{Id}}_{W_{\mathcal{R}}} +\operatorname{\mathrm{Id}}_{V_{\mathcal{R}}} \otimes \theta_W \big)(v\otimes f\otimes w\otimes f')\\ =& \big(\Theta^{\prime}_{X,\mu_i}(v\otimes f)-v\otimes \partial(f)\big) \otimes w\otimes f'+ v\otimes f \otimes \big( \Theta^{\prime}_{X,\nu_j}(w\otimes f')-w\otimes\partial(f') \big)\\ =& \big( \Theta^{\prime}_{X,\mu_i}\otimes \operatorname{\mathrm{Id}}_{W_{\mathcal{R},\nu_j }}+ \operatorname{\mathrm{Id}}_{V_{\mathcal{R},\mu_i }} \otimes \Theta^{\prime}_{X,\nu_j}\big)(v\otimes f\otimes w\otimes f')- v\otimes 1\otimes w\otimes \partial(ff') \\ =& \big(\Theta^{\prime}_{X,\mu_l+\nu_s}-\operatorname{\mathrm{Id}}_{V\otimes W}\otimes \partial \big)(v\otimes w\otimes ff')\\ =& \theta_{V\otimes W} (v\otimes w\otimes ff'), \end{align*} $$

which completes the proof.▪

We now summarize what we have shown thus far. The splitting $\xi _g$ of $\operatorname {\mathrm {{HN}}}_g$ gives a unique element $z\in G(\mathcal {R})$ such that $\operatorname {\mathrm {I}}'(g)=\operatorname {\mathrm {I}}(z)$ by Proposition 4.16, and a unique element $X_0\in \mathfrak {g}_{\mathcal {R}}$ such that $\Theta ^{\prime }_X=\Theta _{X_0}$ by Proposition 4.17. These two elements are related as in Proposition 4.19 below.

We next recall some notions from [Reference Conrad, Gabber and Prasad4, Section 2.1]. Let k be a commutative ring with $1$ , and let $\mathfrak {G}$ be a reductive k-group. Hereupon, we denote by $\kappa (s)$ the residue field of s and $\bar {\kappa }(s)$ an algebraic closure of $\kappa (s)$ , for all $s\in \operatorname {\mathrm {Spec}} k$ . A subgroup $\mathfrak {P}$ of $\mathfrak {G}$ is a parabolic (resp. Borel) subgroup if $\mathfrak {P}$ is smooth and $\mathfrak {P}_{\bar {\kappa }(s)}$ is a parabolic (resp. Borel) subgroup of $\mathfrak {G}_{\bar {\kappa }(s)}$ , for all $s\in \operatorname {\mathrm {Spec}} k$ .

Suppose we have a cocharacter $\lambda \colon \mathbb {G}_m\to \mathfrak {G}$ over k. For any k-algebra R, we let $\mathbb {G}_{m,R}$ act on $\mathfrak {G}_R$ via the conjugation

$$ \begin{align*} \mathbb{G}_{m,R}(S) \times \mathfrak{G}_R(S) \operatorname{\mathrm{\longrightarrow}} \mathfrak{G}_R(S),\;\;\;\;(t,x)\longmapsto t.x:= \lambda(t)x\lambda(t)^{-1} \end{align*} $$

for all R-algebra S. For any $x\in \mathfrak {G}(R)$ , we have an orbit map $\alpha _x\colon \mathbb {G}_{m,R} \to \mathfrak {G}_R$ given by

$$ \begin{align*}\alpha_x\colon \mathbb{G}_{m,R}(S) \operatorname{\mathrm{\longrightarrow}} \mathfrak{G}_R(S),\;\;\;\; t \longmapsto t.x\end{align*} $$

for all R-algebras S. Let $\mathbb {A}^1$ be the affine k-line. We say that the limit $\lim \limits _{t \to 0} t.x$ exists if $\alpha _x$ extends (necessarily uniquely) to a morphism $\tilde \alpha _x\colon \mathbb {A}^1_R \to \mathfrak {G}_R$ of affine R-schemes, and put $\lim \limits _{t \to 0} t.x := \tilde \alpha _x(0)\in \mathfrak {G}_R(R)$ . We define

$$ \begin{align*}P_{\mathfrak{G}}(\lambda) (R):= \big\{x\in \mathfrak{G}(R) \mid \lim\limits_{t \to 0} t.x~\text{exists} \big\},\end{align*} $$
$$ \begin{align*}U_{\mathfrak{G}}(\lambda) (R):= \big\{x\in \mathfrak{G}(R) \mid \lim\limits_{t \to 0} t.x=1 \big\},\end{align*} $$

and

$$ \begin{align*}Z_{\mathfrak{G}}(\lambda) (R):= P_{\mathfrak{G}}(\lambda)(R)\cap P_{\mathfrak{G}}(-\lambda)(R),\end{align*} $$

where $-\lambda $ is the reciprocal of $\lambda $ . Then, $P_{\mathfrak {G}}(\lambda )$ is a closed k-subgroup of $\mathfrak {G}$ [Reference Conrad, Gabber and Prasad4, Lemma 2.1.4], $U_{\mathfrak {G}}(\lambda )$ is an affine algebraic k-normal subgroup of $P_{\mathfrak {G}}(\lambda )$ , and $Z_{\mathfrak {G}}(\lambda )$ is the centralizer of the $\mathbb {G}_m$ -action in $\mathfrak {G}$ [Reference Conrad, Gabber and Prasad4, Lemma 2.1.5]. By [Reference Conrad, Gabber and Prasad4, Proposition 2.1.8(3)], these subgroups are smooth, because $\mathfrak {G}$ is smooth.

It follows from the definitions that the formations of $P_{\mathfrak {G}}(\lambda ),U_{\mathfrak {G}}(\lambda )$ , and $Z_{\mathfrak {G}}(\lambda )$ commute with any base extension on k. In particular, for every $s\in \operatorname {\mathrm {Spec}} k$ , we have $P_{\mathfrak {G}}(\lambda )_{\bar {\kappa }(s)} =P_{\mathfrak {G}_{\bar {\kappa }(s)}}(\lambda _{\bar {\kappa }(s)})$ , which is a parabolic subgroup of $\mathfrak {G}_{\bar {\kappa }(s)}$ by [Reference Springer20, Proposition 8.4.5]. Hence, $P_{\mathfrak {G}}(\lambda )$ is a parabolic k-group.

By [Reference Conrad, Gabber and Prasad4, Proposition 2.1.8(2)], the multiplication map gives an isomorphism

$$ \begin{align*}U_{\mathfrak{G}}(\lambda) \rtimes Z_{\mathfrak{G}}(\lambda) \operatorname{\mathrm{\longrightarrow}} P_{\mathfrak{G}}(\lambda)\end{align*} $$

of affine algebraic k-groups.

Now, let $\mathbb {G}_m$ act on $\mathfrak {g}=\operatorname {\mathrm {Lie}}(\mathfrak {G})(k)$ through the adjoint representation. We then have $\mathfrak {g}=\operatorname *{\mathrm {\bigoplus }}\limits _{n\in \mathbb {Z}} \mathfrak {g}_n$ , where $\mathfrak {g}_n=\{X\in \mathfrak {g} \mid t.X=t^nX,\forall t\in \mathbb {G}_m\}$ for all $n\in \mathbb {Z}$ . We have $\operatorname {\mathrm {Lie}}\big (Z_{\mathfrak {G}}(\lambda ) \big ) =\mathfrak {g}_0$ (which is the centralizer of the $\mathbb {G}_m$ -action on $\mathfrak {g}$ ), $\operatorname {\mathrm {Lie}}\big (U_{\mathfrak {G}}(\lambda ) \big ) =\operatorname *{\mathrm {\bigoplus }}\limits _{n>0} \mathfrak {g}_n$ , and $\operatorname {\mathrm {Lie}}\big (P_{\mathfrak {G}}(\lambda ) \big ) =\operatorname *{\mathrm {\bigoplus }}\limits _{n\geq 0} \mathfrak {g}_n$ . In particular, we have the following decomposition:

(8) $$ \begin{align} \operatorname{\mathrm{Lie}}\big(P_{\mathfrak{G}}(\lambda) \big)=\operatorname{\mathrm{Lie}}\big(Z_{\mathfrak{G}}(\lambda) \big)\textstyle\operatorname*{\mathrm{\bigoplus}} \operatorname{\mathrm{Lie}}\big(U_{\mathfrak{G}}(\lambda) \big). \end{align} $$

Lemma 4.18 With the notion above, we have

$$ \begin{align*}Z-\operatorname{\mathrm{Ad}}(u)(Z)\in \operatorname{\mathrm{Lie}}\big( U_{\mathfrak{G}}(\lambda)\big),\end{align*} $$

for all $u\in U_{\mathfrak {G}}(\lambda )(k)$ and $Z\in \operatorname {\mathrm {Lie}}\big ( Z_{\mathfrak {G}}(\lambda )\big )$ .

Proof Recall that $Z\in Z_{\mathfrak {G}}(\lambda )(k[\varepsilon ])$ by definition; we may also view u as an element in $U_{\mathfrak {G}}(\lambda )(k[\varepsilon ])$ via the inclusion $\iota \colon k\hookrightarrow k[\varepsilon ]$ . By the definition of the adjoint representation, we have

$$ \begin{align*}Z-\operatorname{\mathrm{Ad}}(u)(Z)=Z(uZu^{-1})^{-1}=ZuZ^{-1}u^{-1} \in P_{\mathfrak{G}}(\lambda)(k[\varepsilon]).\end{align*} $$

Because $U_{\mathfrak {G}}(\lambda )$ is normal in $P_{\mathfrak {G}}(\lambda )$ , we have that $ZuZ^{-1}\in U_{\mathfrak {G}}(\lambda )(k[\varepsilon ])$ , and so is $ZuZ^{-1}u^{-1}$ . Consider the following commutative diagram:

Because both Z and $uZ^{-1}u^{-1}$ lie in the kernel of the right vertical map, so does their product $ZuZ^{-1}u^{-1}$ . Hence, $ZuZ^{-1}u^{-1}\in U_{\mathfrak {G}}(\lambda )(k[\varepsilon ])$ lies in the kernel of the left vertical map. The lemma then follows.▪

Proposition 4.19 Let $z\in G(\mathcal {R})$ and $X_0\in \mathfrak {g}_{\mathcal {R}}$ be the unique elements given by Propositions 4.16 and 4.17, respectively. We have $X_0=\Gamma _z \big (\operatorname {\mathrm {\boldsymbol \mu }}\varphi (X_0)\big )$ . In particular, $\operatorname {\mathrm {I}}(z,X_0)$ is a G- $(\varphi ,\nabla )$ -module over $\mathcal {R}$ .

Proof The second assertion follows from the first assertion and Lemma 4.7. For the first assertion, we need to show

(9) $$ \begin{align} X_0=\operatorname{\mathrm{\boldsymbol\mu}} \cdot \operatorname{\mathrm{Ad}}(z)\big(\varphi(X_0)\big)-\operatorname{\mathrm{dlog}}(z). \end{align} $$

It suffices to show (3) with both sides understood as elements in $\operatorname {\mathrm {End}}_{\mathcal {R}}(V_{\mathcal {R}})$ for some faithful representation $(V,\rho )\in \operatorname {\mathrm {\textbf {Rep}}}_F(G)$ . Suppose that $\dim _F V=d$ , and suppose that $\nu _g(V)$ is the decomposition $V_{\mathcal {R}}=\bigoplus \limits _{i=1}^l V_{\mathcal {R},\mu _i}$ . We choose for each graded-piece $V_{\mathcal {R},\mu _i}$ a basis. They altogether give a basis $\textbf {v}_1,\ldots ,\textbf {v}_d$ of $V_{\mathcal {R}}$ , in which $\Phi _g$ acts via a block-upper-triangular matrix

$$ \begin{align*} A= \left(\begin{smallmatrix} A_1 & &\\ & A_2 &~~~~{\ast}\\ && \ddots \\ && &A_l \end{smallmatrix}\right) \in\operatorname{\mathrm{GL}}_d(\mathcal{R}), \end{align*} $$

where each $A_i$ is an $m_i$ by $m_i$ invertible matrix with $m_i$ the multiplicity of $\mu _i$ . Then, $\Phi _z$ acts in this basis via $Z:= \operatorname {\mathrm {Diag}}(A_1,\ldots ,A_l)$ . Likewise, $\Theta _X$ acts in the basis $\textbf {v}_1,\ldots ,\textbf {v}_d$ via a block-upper-triangular matrix

$$ \begin{align*} N= \left(\begin{smallmatrix} N_1 & &\\ & N_2 &~~~~{{\ast}} \\ && \ddots \\ && &N_l \end{smallmatrix}\right) \in\operatorname{\mathrm{Mat}}_{d,d}(\mathcal{R}), \end{align*} $$

where each $N_i$ is an $m_i$ by $m_i$ matrix, and $\Theta _{X_0}$ acts via $\overline N:= \operatorname {\mathrm {Diag}}(N_1,\ldots ,N_l)$ . Write $A=ZU$ for $U\in \operatorname {\mathrm {GL}}_d(\mathcal {R})$ , and $N=\overline N+N_+$ for $N_+\in \operatorname {\mathrm {Mat}}_{d,d}(\mathcal {R})$ . Because $X=\Gamma _g\big (\operatorname {\mathrm {\boldsymbol \mu }}\varphi (X)\big )$ , we have $N=\operatorname {\mathrm {\boldsymbol \mu }}\cdot A \varphi (N) A^{-1}-\partial (A)A^{-1}$ , and then

$$ \begin{align*} \overline N+N_+ =& \operatorname{\mathrm{\boldsymbol\mu}}\cdot (UZ)(\varphi(\overline N+N_+))(UZ)^{-1}-\partial(UZ)(UZ)^{-1}\\ =& \operatorname{\mathrm{\boldsymbol\mu}}\cdot(\!UZ)\varphi(\overline N)Z^{-1}U^{-1}\!+\operatorname{\mathrm{\boldsymbol\mu}}\cdot(\!UZ)\varphi(N_+)Z^{-1}U^{-1}\!-\partial(\!U)U^{-1}\!-U\partial(Z)Z^{-1}U^{-1}. \end{align*} $$

Applying $\operatorname {\mathrm {Ad}}(U^{-1})$ on both sides, we then have

$$ \begin{align*} &\operatorname{\mathrm{\boldsymbol\mu}}\cdot Z\varphi(\overline N)Z^{-1}-\partial(Z)Z^{-1}+\operatorname{\mathrm{\boldsymbol\mu}}\cdot Z\varphi(N_+)Z^{-1}-U^{-1}\partial(U)\\ =& U^{-1}\overline NU+U^{-1}N_+U = \overline N -(\overline N-U^{-1}\overline NU-U^{-1}N_+U). \end{align*} $$

We claim that $\operatorname {\mathrm {\boldsymbol \mu }}\cdot Z\varphi (\overline N)Z^{-1}-\partial (Z)Z^{-1}=\overline N$ . Put $\lambda _{\rho ,g}:= \rho \circ \lambda _g \colon \mathbb {G}_{m,\mathcal {R}} \to \operatorname {\mathrm {GL}}_{V,\mathcal {R}}$ , where $\lambda _g\colon \mathbb {G}_{m,\mathcal {R}} \to G_{\mathcal {R}}$ is the slope morphism defined in Construction 3.15. Identifying $\operatorname {\mathrm {GL}}_{V,\mathcal {R}}$ with $\operatorname {\mathrm {GL}}_{d,\mathcal {R}}$ via the basis $\textbf {v}_1,\ldots ,\textbf {v}_d$ given in the preceding paragraph, and letting $\mathfrak {G}=\operatorname {\mathrm {GL}}_{d,\mathcal {R}}$ , we then have an isomorphism

$$ \begin{align*}U_{\mathfrak{G}}(-\lambda_{\rho,g}) \rtimes Z_{\mathfrak{G}}(-\lambda_{\rho,g}) \cong P_{\mathfrak{G}}(-\lambda_{\rho,g})\end{align*} $$

of affine algebraic $\mathcal {R}$ -groups. Because $\mu _1< \cdots < \mu _l$ , we have

$$ \begin{align*}A\in P_{\mathfrak{G}}(-\lambda_{\rho,g})(\mathcal{R}),~ U\in U_{\mathfrak{G}}(-\lambda_{\rho,g})(\mathcal{R}),~ Z\in Z_{\mathfrak{G}}(-\lambda_{\rho,g})(\mathcal{R});\end{align*} $$

$$ \begin{align*}N\in \operatorname{\mathrm{Lie}}\big( P_{\mathfrak{G}}(-\lambda_{\rho,g})\big),~ N_+\in \operatorname{\mathrm{Lie}}\big( U_{\mathfrak{G}}(-\lambda_{\rho,g})\big),~\overline N\in \operatorname{\mathrm{Lie}}\big( Z_{\mathfrak{G}}(-\lambda_{\rho,g})\big).\end{align*} $$

It follows from Lemma 4.18 that $\overline N-U^{-1}\overline NU \in \operatorname {\mathrm {Lie}} \big ( U_{\mathfrak {G}}(-\lambda _{\rho ,g})\big )$ . In particular, we have $\overline N-U^{-1}\overline NU-U^{-1}N_+U \in \operatorname {\mathrm {Lie}} \big ( U_{\mathfrak {G}}(-\lambda _{\rho ,g})\big )$ . On the other hand, it is clear that $\operatorname {\mathrm {\boldsymbol \mu }}\cdot Z\varphi (\overline N)Z^{-1}-\partial (Z)Z^{-1}\in \operatorname {\mathrm {Lie}}\big ( Z_{\mathfrak {G}}(-\lambda _{\rho ,g})\big )$ and $\operatorname {\mathrm {\boldsymbol \mu }}\cdot Z\varphi (N_+)Z^{-1}-U^{-1}\partial (U)\in \operatorname {\mathrm {Lie}} \big ( U_{\mathfrak {G}}(-\lambda _{\rho ,g})\big )$ . By decomposition (2), we have $\operatorname {\mathrm {\boldsymbol \mu }}\cdot Z\varphi (\overline N)Z^{-1}-\partial (Z)Z^{-1}=\overline N$ , and the desired equality (3) follows.▪

Recall that the least common denominator $d_g$ of g is constructed in Construction 3.8, and $\lambda _g \colon \mathbb {G}_{m,\mathcal {R}} \to G_{\mathcal {R}}$ is the slope morphism (see Construction 3.15). We next reduce the G- $(\varphi ,\nabla )$ -module $(z,X_0)$ over $\mathcal {R}$ to a unit-root one by applying the pushforward functor $[d_g]_*$ and twisting by $\lambda _g(\varpi ^{-1})$ .

Corollary 4.20 $\operatorname {\mathrm {I}} \big (\lambda _g(\varpi ^{-1})[d_g]_*(z),X_0 \big )$ is a unit-root G- $(\varphi ^{d_g},\nabla )$ -module over $\mathcal {R}$ .

Proof For any $V\in \operatorname {\mathrm {\textbf {Rep}}}_F(G)$ , it suffices to show that $(V_{\mathcal {R}},[d_g]_*(z)\varphi ^{d_g},\nabla _{X_0})$ is unit-root. By Proposition 4.19 and Lemma 4.12, $(V_{\mathcal {R}},[d_g]_*(z)\varphi ^{d_g},\nabla _{X_0})$ is a $(\varphi ^{d_g},\nabla )$ -module over $\mathcal {R}$ . Equivalently, we have $\Theta _{X_0}\circ \Phi _z^{d_g}= \operatorname {\mathrm {\boldsymbol \mu }}\cdot \Phi _z^{d_g}\circ \Theta _{X_0}$ . Suppose that $(V_{\mathcal {R}},g\varphi )$ has jumps $\mu _1,\ldots ,\mu _l$ , then $\big ( V_{\mathcal {R}},[d_g]_*(z)\varphi ^{d_g} \big )$ has jumps $d_g\mu _1,\ldots ,d_g\mu _l$ by Lemma 4.13. For any $1\leq i \leq l$ , $\rho (\lambda _g(\varpi ^{-1}))$ acts via multiplication by $\varpi ^{-d_g\mu _i}\in K$ on the graded-piece $V_{\mathcal {R},\mu _i}$ , which implies that $\big ( V_{\mathcal {R},\mu _i},\lambda _g(\varpi ^{-1})[d_g]_*(z)\varphi ^{d_g} \big )$ is unit-root. It follows from [Reference Kedlaya10, Proposition 4.6.3(a)] that $\big ( V_{\mathcal {R}},\lambda _g(\varpi ^{-1})[d_g]_*(z)\varphi ^{d_g} \big )$ is unit-root. Moreover, because $\Theta _{X_0}$ is K-linear, we have

$$ \begin{align*}\Theta_{X_0}\circ \rho(\lambda_g(\varpi^{-1}))\circ \Phi_z^{d_g}= \rho(\lambda_g(\varpi^{-1}))\circ \Theta_{X_0}\circ \Phi_z^{d_g}= \operatorname{\mathrm{\boldsymbol\mu}}\cdot \rho(\lambda_g(\varpi^{-1}))\circ \Phi_z^{d_g}\circ \Theta_{X_0},\end{align*} $$

which completes the proof.▪

4.5 A G-version of the p-adic local monodromy theorem

Let L be a finite separable extension of $\kappa (\!( t )\!)$ , and let $\mathcal {E}^{\dagger }_L$ be the unique unramified extension of $\mathcal {E}^{\dagger }$ with residue field L. We put $\mathcal {R}_L := \mathcal {R}\bigotimes _{\mathcal {E}^{\dagger }} \mathcal {E}^{\dagger }_L$ .

We put

$$ \begin{align*}\mathcal{E}^{\dagger,\operatorname{\mathrm{nr}}}:= \varinjlim_L \mathcal{E}^{\dagger}_L, \;\;\;\text{and} \;\;\; \mathcal{B}_0:= \varinjlim_L \mathcal{R}_L \cong \mathcal{R}\bigotimes_{\mathcal{E}^{\dagger}} \mathcal{E}^{\dagger,\operatorname{\mathrm{nr}}}, \end{align*} $$

where L runs through all finite separable extensions of $\kappa (\!( t )\!)$ . In fact, $\mathcal {E}^{\dagger ,\operatorname {\mathrm {nr}}}$ is the maximal unramified extension of $\mathcal {E}^{\dagger }$ with residue field $\kappa (\!( t )\!)^{\operatorname {\mathrm {sep}}}$ , the separable closure of $\kappa (\!( t )\!)$ .

The main result of this paper is the following theorem.

Theorem 4.21 Let G be a connected reductive F-group, and let $(g,X)\in \textbf {B}^{\varphi ,\nabla }(G,\mathcal {R})$ . Then, there exist a finite separable extension L of $\kappa (\!( t )\!)$ and an element $b\in G(\mathcal {R}_L)$ such that $\Gamma _b(X)\in \operatorname {\mathrm {Lie}} \big (U_{G_{\mathcal {R}}}(-\lambda _g)\big )_{\mathcal {R}_L}$ .

We will make use of the following lemma, which is often mentioned as Steinberg’s theorem. The theory of fields of cohomological dimension $\leq 1$ can be found in, e.g., [Reference Serre19, Chapter II, Section 3]; for us, the most important example will be a Henselian discretely valued field of characteristic $0$ with algebraically closed residue field (see [Reference Serre19, Chapter II, Section 3.3]).

Lemma 4.22 ([Reference Steinberg21, Theorem 1.9])

Suppose that k is a field of cohomological dimension $\leq 1$ and $\mathfrak {G}$ is a connected reductive k-group, then $H^1(k,\mathfrak {G})=1$ .

Proof of Theorem 4.21 Let $z\in G(\mathcal {R})$ and $X_0\in \mathfrak {g}_{\mathcal {R}}$ be the unique elements given by Propositions 4.16 and 4.17, respectively.

Let $(V,\rho )$ be a d-dimensional G-representation (not necessarily faithful). Suppose the slope filtration of $(V_{\mathcal {R}},g\varphi )$ has jumps $\mu _1,\ldots ,\mu _l$ . Suppose that $\xi _g(V)=\bigoplus \limits _{i=1}^l V_{\mathcal {R},\mu _i}$ , we put $d_i:= \operatorname {\mathrm {rk}}_{\mathcal {R}} (V_{\mathcal {R},\mu _i})$ for all i. In the proof of Corollary 4.20, we see that $\big (V_{\mathcal {R},\mu _i},\lambda _g(\varpi ^{-1})[d_g]_*(z)\varphi ^{d_g},\nabla _{X_0} \big )$ is a unit-root $(\varphi ,\nabla )$ -module over $\mathcal {R}$ for all $1\leq i\leq l$ . Let $\Phi _z=z\varphi $ , and let $\Theta _{X_0}\colon V_{\mathcal {R}}\to V_{\mathcal {R}}$ be the differential operator associated to $\nabla _{X_0}$ . Then, $\Phi _z$ (resp. $\Theta _{X_0}$ ) may be extended to $V\bigotimes _F \mathcal {B}_0$ , which is still denoted by $\Phi _z$ (resp. $\Theta _{X_0}$ ). By the unit-root pLMT [Reference Kedlaya9, Theorem 6.11], we find:

  1. (i) a finite separable extension $L(V)$ of $\kappa (\!( t )\!)$ ;

  2. (ii) for each $1\leq i\leq l$ , a basis $\textbf {w}_1^{(i)},\ldots ,\textbf {w}_{d_i}^{(i)}$ for $V_{\mathcal {R},\mu _i}\bigotimes _{\mathcal {R}} \mathcal {R}_{L(V)}$ over $\mathcal {R}_{L(V)}$ such that $\Theta _{X_0}(\textbf {w}_j^{(i)})=0$ , for all $1\leq j\leq d_i$ .

Then, for each $1\leq i\leq l$ , we have that

$$ \begin{align*}W_i:= (V_{\mathcal{R},\mu_i}\bigotimes\limits_{\mathcal{R}} \mathcal{B}_0)^{\Theta_{X_0}=0}= \big\{ x\in V_{\mathcal{R},\mu_i}\bigotimes\limits_{\mathcal{R}} \mathcal{B}_0 \mid \Theta_{X_0}(x)=0 \big\}\end{align*} $$

is a $d_i$ -dimensional $K^{\operatorname {\mathrm {nr}}}$ -vector space spanned by $\textbf {w}_1^{(i)},\ldots ,\textbf {w}_{d_i}^{(i)}$ . In particular, we have

$$ \begin{align*} (V_{\mathcal{B}_0} )^{\Theta_{X_0}=0}= \big\{ x\in V_{\mathcal{B}_0} \mid \Theta_{X_0}(x)=0 \big\}=\operatorname*{\mathrm{\bigoplus}}\limits_{i=1}^l W_i,\end{align*} $$

which is a $d_i$ -dimensional $K^{\operatorname {\mathrm {nr}}}$ -vector space.

We now have two $K^{\operatorname {\mathrm {nr}}}$ -valued fiber functors

$$ \begin{align*}\omega_1=\omega^G \otimes K^{\operatorname{\mathrm{nr}}}\colon \operatorname{\mathrm{\textbf{Rep}}}_F(G) \operatorname{\mathrm{\longrightarrow}} \operatorname{\mathrm{\textbf{Vec}}}_{K^{\operatorname{\mathrm{nr}}}},\;\;\;\; V \longmapsto V\otimes K^{\operatorname{\mathrm{nr}}},\end{align*} $$

and

$$ \begin{align*}\omega_2 \colon \operatorname{\mathrm{\textbf{Rep}}}_F(G) \operatorname{\mathrm{\longrightarrow}} \operatorname{\mathrm{\textbf{Vec}}}_{K^{\operatorname{\mathrm{nr}}}},\;\;\;\; V\longmapsto (V_{\mathcal{B}_0} )^{\Theta_{X_0}=0}.\end{align*} $$

Moreover, we have an action

$$ \begin{align*}\operatorname{\mathrm{\underline{Isom}}}^{\otimes}(\omega_1,\omega_2) \times \operatorname{\mathrm{\underline{Aut}}}^{\otimes}(\omega_1)\operatorname{\mathrm{\longrightarrow}} \operatorname{\mathrm{\underline{Isom}}}^{\otimes}(\omega_1,\omega_2)\end{align*} $$

of $\operatorname {\mathrm {\underline {Aut}}}^{\otimes }(\omega _1)$ on $\operatorname {\mathrm {\underline {Isom}}}^{\otimes }(\omega _1,\omega _2)$ , given by precomposition. We note that $\operatorname {\mathrm {\underline {Aut}}}^{\otimes }(\omega _1)=\operatorname {\mathrm {\underline {Aut}}}^{\otimes }(\omega ^G\otimes K^{\operatorname {\mathrm {nr}}})\cong G_{K^{\operatorname {\mathrm {nr}}}}$ ,Footnote 2 so $\operatorname {\mathrm {\underline {Isom}}}^{\otimes }(\omega _1,\omega _2)$ may be viewed as a $G_{K^{\operatorname {\mathrm {nr}}}}$ -torsor over $K^{\operatorname {\mathrm {nr}}}$ . By Lemma 4.22, we have $H^1(K^{\operatorname {\mathrm {nr}}},G_{K^{\operatorname {\mathrm {nr}}}})=1$ . Thus, $\operatorname {\mathrm {\underline {Isom}}}^{\otimes }(\omega _1,\omega _2)$ is isomorphic to the trivial $G_{K^{\operatorname {\mathrm {nr}}}}$ -torsor over $K^{\operatorname {\mathrm {nr}}}$ , i.e., we have $\operatorname {\mathrm {\underline {Isom}}}^{\otimes }(\omega _1,\omega _2)_{K^{\operatorname {\mathrm {nr}}}} \cong G_{K^{\operatorname {\mathrm {nr}}}}$ .

On the other hand, we have an isomorphism $\gamma \colon \omega _2\otimes \mathcal {B}_0 \to \omega _1 \otimes \mathcal {B}_0$ of tensor functors, induced by the $\mathcal {B}_0$ -linear extension of the inclusion

for all $(V,\rho )\in \operatorname {\mathrm {\textbf {Rep}}}_F(G)$ . We now fix $\beta \in \operatorname {\mathrm {\underline {Isom}}}^{\otimes }(\omega _1,\omega _2)(K^{\operatorname {\mathrm {nr}}})$ ; we then have an element $\tilde \beta := \gamma \circ \beta _{\mathcal {B}_0} \in \operatorname {\mathrm {\underline {Aut}}}^{\otimes }(\omega _1\otimes \mathcal {B}_0)(\mathcal {B}_0)=G(\mathcal {B}_0)$ . Let $b\in G(\mathcal {B}_0)$ be the inverse of the image of $\tilde \beta $ under the following isomorphism:

$$ \begin{align*}\operatorname{\mathrm{\underline{Aut}}}^{\otimes}(\omega_1\otimes \mathcal{B}_0)(\mathcal{B}_0) \operatorname{\mathrm{\longrightarrow}} G(\mathcal{B}_0).\end{align*} $$

Because $F[G]$ is finitely presented over F, the functor $\operatorname {\mathrm {Hom}}_{\operatorname {\mathrm {\textbf {Alg}}}_F}(F[G],\rule {2mm}{0.15mm})$ commutes with colimits. We have

$$ \begin{align*}G(\mathcal{B}_0)=G(\varinjlim_L \mathcal{R}_L)=\varinjlim_L G(\mathcal{R}_L),\end{align*} $$

where L runs over all finite separable extensions of $\kappa (\!( t )\!)$ ; we thus find a finite separable extension L of $\kappa (\!( t )\!)$ such that $b\in G(\mathcal {R}_L)$ .

For any $(V,\rho )\in \operatorname {\mathrm {\textbf {Rep}}}_F(G)$ , it follows from the construction of b that the automorphism $\rho (b^{-1})\colon V_{\mathcal {B}_0} \to V_{\mathcal {B}_0}$ factors through $(V_{\mathcal {B}_0})^{\Theta _{X_0}=0} \otimes \mathcal {B}_0$ . Notice that $\Theta _{X_0}$ and $X_0$ agree on $\omega _1(V)=V_{K^{\operatorname {\mathrm {nr}}}}$ . Therefore, we have

(10) $$ \begin{align} \rho(b)X_0\rho(b^{-1})-\partial(\rho(b))\rho(b^{-1})=0. \end{align} $$

We now fix a faithful representation $(V,\rho )$ . The equality (4) then implies

$$ \begin{align*}\Gamma_{b}(X_0)=0.\end{align*} $$

Put $X_1:= X-X_0\in \mathfrak {g}_{\mathcal {R}}$ ; we then have

$$ \begin{align*} \Gamma_{b}(X) &=\operatorname{\mathrm{Ad}}(b)(X_0+X_1)-\operatorname{\mathrm{dlog}}(b)\\ &= \operatorname{\mathrm{Ad}}(b)(X_0)-\operatorname{\mathrm{dlog}}(b)+\operatorname{\mathrm{Ad}}(b)(X_1)\\ &=\Gamma_{b}(X_0)+\operatorname{\mathrm{Ad}}(b)(X_1)\\ &=\operatorname{\mathrm{Ad}}(b)(X_1). \end{align*} $$

Conserving the notation as in the second paragraph, $\Theta _X=\rho (b) X_1 \rho (b^{-1})$ acts in the basis $\textbf {w}_1^{(1)},\ldots ,\textbf {w}_{d_1}^{(1)},\ldots ,\textbf {w}_1^{(l)},\ldots ,\textbf {w}_{d_l}^{(l)}$ via a matrix of the form

$$ \begin{align*} \left(\begin{smallmatrix} 0 & &\\ && 0 &~~~~{{\ast}} \\ &&& \ddots \\ &&& & 0 \end{smallmatrix}\right) \in\operatorname{\mathrm{Mat}}_{d,d}(\mathcal{R}_L). \end{align*} $$

Here, the ith $0$ in the diagonal denotes the zero matrix of size $d_i\times d_i$ . Hence, $\Gamma _{b}(X)\in \operatorname {\mathrm {Lie}} \big (U_{G_{\mathcal {R}_L}}(-\lambda _{g,\mathcal {R}_L})\big )=\operatorname {\mathrm {Lie}}\big ( U_{G_{\mathcal {R}}}(-\lambda _g)_{\mathcal {R}_L} \big )=\operatorname {\mathrm {Lie}} \big (U_{G_{\mathcal {R}}}(-\lambda _g)\big )_{\mathcal {R}_L}$ , as desired.▪

Acknowledgment

The content of this paper is part of the author’s Ph.D. thesis carried out at Humboldt-Universität zu Berlin. He owes a deep gratitude to his supervisor Elmar Grosse-Klönne for providing him this problem, and for all the helpful discussions. He would like to thank the external examiners of the thesis for their valuable feedback. He is also indebted to Claudius Heyer for many constructive suggestions. Finally, he would like to thank the anonymous referees for very helpful comments and suggestions which have greatly improved the presentation of this paper.

Footnotes

This paper is partially supported by a research grant from Shanghai Key Laboratory of PMMP 18dz2271000.

1 By an algebra, we always mean a commutative algebra with $1$ .

2 For this isomorphism, we refer to the discussion above Proposition 3.11.

References

André, Y., Filtrations de type Hasse–Arf et monodromie $p$ -adique . Invent. Math. 148(2002), 285317.Google Scholar
Anschütz, J., Reductive group schemes over the Fargues–Fontaine curve. Math. Ann. 374(2019), 12191260.Google Scholar
Berger, L., Représentations  $p$ -adiques et équations différentielles . Invent. Math. 148(2002), 219284.CrossRefGoogle Scholar
Conrad, B., Gabber, O., and Prasad, G., Pseudo-reductive groups. 2nd ed., New Mathematical Monographs, 26, Cambridge University Press, Cambridge, 2015.CrossRefGoogle Scholar
Crew, R., Finiteness theorems for the cohomology of an overconvergent isocrystal on a curve. Ann. Sci. École Norm. Sup. 31(1998), 717763.CrossRefGoogle Scholar
Dat, J.-F., Orlik, S., and Rapoport, M., Period domains over finite and p-adic fields. Cambridge Tracts in Mathematics, 183, Cambridge University Press, Cambridge, 2010.CrossRefGoogle Scholar
Deligne, P. and Milne, J., Tannakian categories. In: Hodge cycles, motives, and Shimura varieties, Lecture Notes in Mathematics, 900, Springer, 1982, pp. 101228.CrossRefGoogle Scholar
Demazure, M. and Gabriel, P., Groupes algébriques. Tome I: Géométrie algébrique, généralités, groupes commutatifs. Masson, Paris-Amsterdam, 1970.Google Scholar
Kedlaya, K. S., A  $p$ -adic local monodromy theorem . Ann. Math. 160(2004), 93184.CrossRefGoogle Scholar
Kedlaya, K. S., Slope filtrations revisited. Doc. Math. 10(2005), 447525.Google Scholar
Kedlaya, K. S., Slope filtrations for relative Frobenius. Astérisque 317(2008), 259301.Google Scholar
Kedlaya, K. S., p-Adic differential equations. Cambridge Studies in Advanced Mathematics, 125, Cambridge University Press, Cambridge, 2010.CrossRefGoogle Scholar
Kedlaya, K. S., Notes on isocrystals. Preprint, 2018. arXiv:1606.01321.Google Scholar
Kottwitz, R., Isocrystals with additional structures. Compos. Math. 56(1985), 201220.Google Scholar
Liu, R., Slope filtrations in families. J. Inst. Math. Jussieu 12(2013), 249296.CrossRefGoogle Scholar
Marmora, A., Facteurs epsilon  $p$ -adiques . Compos. Math. 144(2008), 439483.Google Scholar
Mebkhout, Z., Analogue  $p$ -adique du théorème de turrittin et le théorème de la monodromie  $p$ -adique . Invent. Math. 148(2002), 319351.CrossRefGoogle Scholar
Milne, J., Algebraic groups. Cambridge Studies in Advanced Mathematics, 170, Cambridge University Press, Cambridge, 2017.CrossRefGoogle Scholar
Serre, J.-P., Galois cohomology. Corrected second printing edition, Springer Monographs in Mathematics, Springer-Verlag, Berlin-Heidelberg, 2002.Google Scholar
Springer, T. A., Linear algebraic groups. 2nd ed., Modern Birkhäuser Classics, Birkhäuser, Boston, 1998.CrossRefGoogle Scholar
Steinberg, R., Regular elements of semisimple algebraic groups. Inst. Hautes Études Sci. Publ. Math. 25(1965), 4980.CrossRefGoogle Scholar
Tsuzuki, N., The overconvergence of morphisms of etale  $\varphi \hbox{--} \nabla$ -spaces on a local field . Compos. Math. 103(1996), 227239.Google Scholar
Tsuzuki, N., Finite local monodromy of overconvergent unit-root  $F$ -isocrystals on a curve . Amer. J. Math. 120(1998), 11651190.CrossRefGoogle Scholar
Tsuzuki, N., Slope filtration of quasi-unipotent overconvergent  $F$ -isocrystals . Ann. Inst. Fourier 48(1998), 379412.CrossRefGoogle Scholar
Ziegler, P., Graded and filtered fiber functors on Tannakian categories. J. Inst. Math. Jussieu 14(2015), 87130.CrossRefGoogle Scholar