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

A Jacquet-Langlands functor for pp-adic locally analytic representations

Gabriel Dospinescu  and  Juan Esteban Rodríguez Camargo
Abstract.

We study the locally analytic theory of infinite level local Shimura varieties. As a main result, we prove that in the case of a duality of local Shimura varieties, the locally analytic vectors of different period sheaves at infinite level are independent of the actions of the pp-adic Lie groups GG and GbG_{b} of the two towers; this generalizes a result of Pan for the Lubin-Tate and Drinfeld spaces for 𝐆𝐋2\mathbf{GL}_{2}. We apply this theory to show that the pp-adic Jacquet-Langlands functor of Scholze commutes with the passage to locally analytic vectors, and is compatible with central characters of Lie algebras. We also prove that the compactly supported de Rham cohomology of the two towers are isomorphic as smooth representations of G×GbG\times G_{b}.

1991 Mathematics Subject Classification:
11F77, 11G18 ,14G35, 22E50

1. Introduction

Let pp be a prime number. The main objective of this work is to give some new insights in the locally analytic incarnation of the pp-adic local Langlands correspondence, cf. [Bre10] [CDP14]. The objects of study in this paper are the infinite level Lubin-Tate and Drinfeld spaces (or more generally local Shimura varieties) and the locally analytic vectors for the action of the associated pp-adic Lie groups on different period sheaves. To motivate our main results let us recall the perfectoid geometry of the Lubin-Tate and Drinfeld spaces.

Let FF be a finite extension of p\mathbb{Q}_{p} with ring of integers 𝒪\mathcal{O}, pseudo-uniformizer ϖ\varpi and residue field 𝔽\mathbb{F}. Fix 𝔽¯\overline{\mathbb{F}} an algebraic closure of 𝔽\mathbb{F} and let 𝒪˘\breve{\mathcal{O}} be the completion of the maximal unramified extension of 𝒪\mathcal{O} with residue field 𝔽¯\overline{\mathbb{F}}. Write F˘=𝒪˘[1ϖ]\breve{F}=\breve{\mathcal{O}}[\frac{1}{\varpi}]. Consider the group 𝐆𝐋n,F\mathbf{GL}_{n,F}, let μ\mu be the cocharacter (1,0,,0)(1,0,\cdots,0) with n1n-1 occurrences of 0 and let 𝕏b\mathbb{X}_{b} be a formal 𝒪\mathcal{O}-module over 𝔽¯\overline{\mathbb{F}} of dimension 11 and 𝒪\mathcal{O}-height nn. Denote by DD the central division algebra over FF with invariant 1/n1/n. Let Def𝕏b\operatorname{\scriptsize Def}_{\mathbb{X}_{b}} be the formal scheme over 𝒪˘\breve{\mathcal{O}} parametrizing deformations of the formal 𝒪\mathcal{O}-module 𝕏b\mathbb{X}_{b}. The Lubin-Tate space of 𝐆𝐋n,F\mathbf{GL}_{n,F} at level 𝐆𝐋n(𝒪)\mathbf{GL}_{n}(\mathcal{O}) is the rigid generic fiber 𝒯F˘\mathcal{LT}_{\breve{F}} of Def𝕏b\operatorname{\scriptsize Def}_{\mathbb{X}_{b}}. The space 𝒯F˘\mathcal{LT}_{\breve{F}} has a dual given by the Drinfeld space ΩF˘F˘n1\Omega_{\breve{F}}\subset\mathbb{P}^{n-1}_{\breve{F}} defined as the complement of the FF-rational hyperplanes of F˘n1\mathbb{P}^{n-1}_{\breve{F}}. These are particular examples of Rapoport-Zink spaces [RZ96] which are themselves special cases of local Shimura varieties [RV14] [SW20].

The spaces 𝒯F˘\mathcal{LT}_{\breve{F}} and ΩF˘\Omega_{\breve{F}} are intimately related via perfectoid geometry in a very clean way: let 𝒯,F˘\mathcal{LT}_{\infty,\breve{F}} be the Lubin-Tate space at infinite level obtained by trivializing the Tate module of the universal deformation of 𝕏b\mathbb{X}_{b}. It was shown in [SW13] that 𝒯,F˘\mathcal{LT}_{\infty,\breve{F}} has a natural structure of a perfectoid space which, by construction, is a proétale 𝐆𝐋n(𝒪)\mathbf{GL}_{n}(\mathcal{O})-torsor over 𝒯F˘\mathcal{LT}_{\breve{F}}. Furthermore, a Hodge-Tate period map is constructed in loc. cit.

πHT:𝒯,F˘F˘n1.\pi_{\operatorname{\scriptsize HT}}:\mathcal{LT}_{\infty,\breve{F}}\to\mathbb{P}^{n-1}_{\breve{F}}.

The image of the Hodge-Tate period map is the Drinfeld space ΩF˘\Omega_{\breve{F}}, and the map πHT:𝒯,F˘ΩF˘\pi_{\operatorname{\scriptsize HT}}:\mathcal{LT}_{\infty,\breve{F}}\to\Omega_{\breve{F}} is a proétale D×D^{\times}-torsor. When composing 𝒯,F˘𝒯F˘\mathcal{LT}_{\infty,\breve{F}}\to\mathcal{LT}_{\breve{F}} with the Grothendieck-Messing period map 𝒯F˘F˘n1\mathcal{LT}_{\breve{F}}\to\mathbb{P}^{n-1}_{\breve{F}}, the morphism

πGM:𝒯,F˘F˘n1\pi_{\operatorname{\scriptsize GM}}:\mathcal{LT}_{\infty,\breve{F}}\to\mathbb{P}^{n-1}_{\breve{F}}

becomes a proétale 𝐆𝐋n(F)\mathbf{GL}_{n}(F)-torsor.

Summarizing, we have a perfectoid space 𝒯,F˘\mathcal{LT}_{\infty,\breve{F}} endowed with an action of 𝐆𝐋n(F)×D×\mathbf{GL}_{n}(F)\times D^{\times} fitting in an equivariant diagram

𝒯,F˘{\mathcal{LT}_{\infty,\breve{F}}}F˘n1{\mathbb{P}^{n-1}_{\breve{F}}}ΩF˘{\Omega_{\breve{F}}}πHT\scriptstyle{\pi_{\operatorname{\scriptsize HT}}}πGM\scriptstyle{\pi_{\operatorname{\scriptsize GM}}} (1.1)

such that:

  • πHT\pi_{\operatorname{\scriptsize HT}} is a 𝐆𝐋n(F)\mathbf{GL}_{n}(F)-equivariant D×D^{\times}-torsor for the natural action of 𝐆𝐋n(F)\mathbf{GL}_{n}(F) on ΩF˘\Omega_{\breve{F}}.

  • πGM\pi_{\operatorname{\scriptsize GM}} is a D×D^{\times}-equivariant 𝐆𝐋n(F)\mathbf{GL}_{n}(F)-torsor where D×D^{\times} acts on F˘n1\mathbb{P}^{n-1}_{\breve{F}} via its embedding into 𝐆𝐋n(F˘)\mathbf{GL}_{n}(\breve{F}).

  • The diagram carries a suitable Weil descent over FF. Thus, its base change to p\mathbb{C}_{p} carries an action of the Weil group WFW_{F}.

The diagram (1.1) actually encodes the isomorphism of the Lubin-Tate and Drinfeld towers which was previously established by Fargues in [Far08] and envisioned by Faltings in [Fal02].

In particular, there is an action of 𝐆𝐋n(F)×D××WF\mathbf{GL}_{n}(F)\times D^{\times}\times W_{F} on the infinite level Lubin-Tate space 𝒯,p\mathcal{LT}_{\infty,\mathbb{C}_{p}} which makes natural the expectation that both the Jacquet-Langlands correspondence, relating 𝐆𝐋n(F)\mathbf{GL}_{n}(F) and D×D^{\times}-representations, and the Langlands correspondence, relating 𝐆𝐋n(F)\mathbf{GL}_{n}(F) and WFW_{F}-representations, can be realized in different cohomologies attached to 𝒯,p\mathcal{LT}_{\infty,\mathbb{C}_{p}}, see for example [Far08], [DLB17], [CDN20], [CDN21], [CDN23].

In [Sch18], Scholze used the diagram (1.1) to construct a pp-adic Jacquet-Langlands functor sending smooth admissible representations of 𝐆𝐋n(F)\mathbf{GL}_{n}(F) to smooth admissible representations of D×D^{\times}. Let us be more precise; let π\pi be an admissible representation of 𝐆𝐋n(F)\mathbf{GL}_{n}(F) over an artinian ring AA which is pp-power torsion. Since πGM\pi_{\operatorname{\scriptsize GM}} is a proétale 𝐆𝐋n(F)\mathbf{GL}_{n}(F)-torsor the representation π\pi gives rise an étale sheaf π\mathcal{F}_{\pi} on the rigid space pn1\mathbb{P}^{n-1}_{\mathbb{C}_{p}}. Furthermore, this étale sheaf descends to the vv-stack [pn1/D×][\mathbb{P}^{n-1}_{\mathbb{C}_{p}}/D^{\times}] and so its cohomology carries a natural action of D×D^{\times}. The Jacquet-Langlands functor 𝒥\mathcal{JL} is the functor sending a smooth admissible representation π\pi of 𝐆𝐋n(F)\mathbf{GL}_{n}(F) to the complex of smooth D×D^{\times}-representations over AA

𝒥(π)=RΓe´t(pn1,π).\mathcal{JL}(\pi)=R\Gamma_{{\rm\acute{e}t}}(\mathbb{P}^{n-1}_{\mathbb{C}_{p}},\mathcal{F}_{\pi}).

By [Sch18, Theorem 1.1] the cohomology groups 𝒥i(π)\mathcal{JL}^{i}(\pi) of 𝒥(π)\mathcal{JL}(\pi) are smooth admissible representations of D×D^{\times} over AA and 𝒥i(π)=0\mathcal{JL}^{i}(\pi)=0 for i>2(n1)i>2(n-1). In addition, [Sch18, Theorem 1.3] says that this construction satisfies a local-global compatibility for 𝐆𝐋2\mathbf{GL}_{2} (though the main ideas should hold for general 𝐆𝐋n\mathbf{GL}_{n}), justifying the compatibility with a more classical Jacquet-Langlands correspondence.

One can naturally extend the Jacquet-Langlands functor to unitary Banach representations and it is not hard to see that it also preserves admissible Banach representations, see Corollary 5.3.5. On the other hand, Schneider-Teitelbaum introduced a class of admissible locally analytic representations for pp-adic Lie groups in [ST03]. A natural question arises:

Question 1.1.

Is there a Jacquet-Langlands functor π𝒥(π)\pi\mapsto\mathcal{JL}(\pi) for admissible locally analytic representations of 𝐆𝐋n(F)\mathbf{GL}_{n}(F)? If so, is it compatible with the Jacquet-Langlands functor of Banach representations?

In this paper we give a partial answer to this question, namely, that the Jacquet-Langlands functor for admissible Banach representations is compatible with the passage to locally analytic vectors, see 1.8 for a more precise statement.

In a different direction, the works of Lue Pan [Pan22a, Pan22b] studying the locally analytic vectors of perfectoid modular curves use some special sheaves of locally analytic functions at infinite level. These sheaves encode, via the localization theory of Beilinson-Bernstein on the flag variety [BB81] and the Hodge-Tate period map, many aspects of the pp-adic Hodge theory of Shimura varieties. In [RC23, RC24b] some of these features have been generalized to arbitrary global Shimura varieties under the name of geometric Sen theory; part of the goals of this paper is to extend the results in geometric Sen theory from the global to the local set up. It is then natural to ask what additional properties local Shimura varieties acquire after taking locally analytic vectors, in particular one can ask the following question:

Question 1.2.

Let 𝒯,p\mathcal{LT}_{\infty,\mathbb{C}_{p}} be the infinite level Lubin-Tate space and let 𝒪^𝒯\widehat{\mathscr{O}}_{\mathcal{LT}} be its structural sheaf as a perfectoid space. Let U𝒯,pU_{\infty}\subset\mathcal{LT}_{\infty,\mathbb{C}_{p}} be an affinoid perfectoid and let KU𝐆𝐋n(F)K_{U}\subset\mathbf{GL}_{n}(F) and KU,DD×K_{U,D}\subset D^{\times} be the (open) stabilizers of UU_{\infty}. Do we have an equality of locally analytic vectors

𝒪^𝒯(U)KUla=𝒪^𝒯(U)KU,Dla\widehat{\mathscr{O}}_{\mathcal{LT}}(U_{\infty})^{K_{U}-la}=\widehat{\mathscr{O}}_{\mathcal{LT}}(U_{\infty})^{K_{U,D}-la}

as subspaces of 𝒪^𝒯(U)\widehat{\mathscr{O}}_{\mathcal{LT}}(U_{\infty})? Equivalently, are the locally analytic vectors of the structural sheaf at infinite level independent of the tower?

For the case of 𝐆𝐋2\mathbf{GL}_{2} this is proven by Pan in [Pan22b, Corollary 5.3.9] via explicit power series expansions. In this paper we prove a much more general result that holds for an arbitrary duality of local Shimura varieties and arbitrary period sheaves appearing in the affinoid charts of relative Fargues-Fontaine curves, see 1.2 for a precise statement. Then, the partial result towards 1.1 mentioned above will be a rather formal consequence of this independence of locally analytic vectors at infinite level, after applying enough technology coming from the theory of solid locally analytic representations [RJRC22, RJRC23]. We also apply this independence of locally analytic vectors to construct an equivariant isomorphism for the compactly supported de Rham cohomology between the two towers of a duality of local Shimura varieties, see 1.7.

In order to present the main results of this paper we have separated the introduction in different paragraphs, going from the general results on towers of rigid spaces, passing to the applications to local Shimura varieties, and finishing with the most specific applications to the Lubin-Tate and Drinfeld towers.

Main results

Cohomology of towers of rigid spaces

In this paragraph we explain the results of Section 5.1 about locally analytic vectors of period sheaves in towers of rigid spaces. Let Perfd\operatorname{Perfd} be the category of perfectoid spaces and PerfPerfd\operatorname{Perf}\subset\operatorname{Perfd} the full subcategory of perfectoid spaces in characteristic pp. Following [Sch22] we see Perfd\operatorname{Perfd} and Perf\operatorname{Perf} as sites endowed with the vv-topology. Let Perfϖ\operatorname{Perf}_{\varpi} be the category of perfectoid spaces in characteristic pp endowed with a fixed pseudo-uniformizer ϖ\varpi, and with maps preserving the pseudo-uniformizer. We can define different period sheaves as follows:

  • i.

    We have the structural sheaves 𝒪^\widehat{\mathscr{O}} and 𝒪^+\widehat{\mathscr{O}}^{+} mapping an affinoid perfectoid Spa(R,R+)Perfd\operatorname{Spa}(R,R^{+})\in\operatorname{Perfd} to 𝒪^(R,R+)=R\widehat{\mathscr{O}}(R,R^{+})=R and 𝒪^+(R,R+)=R+\widehat{\mathscr{O}}^{+}(R,R^{+})=R^{+} respectively.

  • ii.

    We have the tilted sheaves 𝒪\mathscr{O}^{\flat} and 𝒪,+\mathscr{O}^{\flat,+} mapping an affinoid perfectoid Spa(R,R+)Perfd\operatorname{Spa}(R,R^{+})\in\operatorname{Perfd} to 𝒪(R,R+)=R\mathscr{O}^{\flat}(R,R^{+})=R^{\flat} and 𝒪,+(R,R+)=R,+\mathscr{O}^{\flat,+}(R,R^{+})=R^{\flat,+} respectively.

  • iii.

    We have the period sheaf 𝔸inf\mathbb{A}_{\inf} mapping an affinoid perfectoid Spa(R,R+)Perfd\operatorname{Spa}(R,R^{+})\in\operatorname{Perfd} to 𝔸inf(R,R+)=W(R,+)\mathbb{A}_{\inf}(R,R^{+})=W(R^{\flat,+}) where W()W(-) is the functor of pp-typical Witt vectors.

  • iv.

    For I=[s,r](0,)I=[s,r]\subset(0,\infty) a compact interval with rational ends we define the period ring 𝔹I\mathbb{B}_{I} mapping an affinoid perfectoid Spa(R,R+)Perfϖ\operatorname{Spa}(R,R^{+})\in\operatorname{Perf}_{\varpi} in characteristic pp with fixed pseudo-uniformizer to the rational localization

    𝔹I(R,R+)=𝔸inf(R,R+)(p[ϖ]1/r,[ϖ]1/sp)[1[ϖ]].\mathbb{B}_{I}(R,R^{+})=\mathbb{A}_{\inf}(R,R^{+})\left(\frac{p}{[\varpi]^{1/r}},\frac{[\varpi]^{1/s}}{p}\right)[\frac{1}{[\varpi]}].

The period sheaves in (i)-(iii) above are standard in pp-adic Hodge theory. The sheaves in (iv) give rise to affinoid charts of families of Fargues-Fontaine curves as in [SW20] and [FS24].

Let KK be a perfectoid field in characteristic 0 containing the pp-th powers roots of unit and let XX be a qcqs smooth rigid space over KK. Let HH be a compact pp-adic Lie group acting on XX. Let GG be another compact pp-adic Lie group and suppose we are given with an HH-equivariant pro-finite-étale GG-torsor X~X\widetilde{X}\to X^{\lozenge} of the diamond attached to XX. In particular, X~\widetilde{X} is endowed with an action of the pp-adic Lie group G×HG\times H. The following theorem relates the locally analytic vectors of the vv-cohomologies of period sheaves at infinite level.

Theorem 1.3 (Theorem 5.1.1).

Let I(0,)I\subset(0,\infty) be a compact interval with rational ends. Then the GG-locally analytic vectors of the solid p\mathbb{Q}_{p}-linear representation RΓv(X~,𝔹I)R\Gamma_{v}(\widetilde{X},\mathbb{B}_{I}) are HH-locally analytic. More precisely, the natural map of solid G×HG\times H-representations

RΓv(X~,𝔹I)RG×HlaRΓv(X~,𝔹I)RGlaR\Gamma_{v}(\widetilde{X},\mathbb{B}_{I})^{RG\times H-la}\xrightarrow{\sim}R\Gamma_{v}(\widetilde{X},\mathbb{B}_{I})^{RG-la}

is an equivalence.

Remark 1.1.

1.3 holds for a larger class of 𝔹I\mathbb{B}_{I}-modules, including 𝔹I\mathbb{B}_{I} and 𝒪^\widehat{\mathscr{O}}-vector bundles, see Remark 5.1.5.

As a corollary, in the case when G=1G=1, we prove that proétale cohomologies of qcqs rigid varieties XX endowed with actions of pp-adic Lie groups HH tent to be locally analytic:

Corollary 1.1 (Corollary 5.1.6).

Let XX be a qcqs smooth rigid space endowed with the action of a pp-adic Lie group HH. Then for I(0,)I\subset(0,\infty) a compact interval with rational ends the solid HH-representation RΓv(X,𝔹I)R\Gamma_{v}(X,\mathbb{B}_{I}) is HH-locally analytic.

Remark 1.2.

1.1 implies that the cohomology groups of period sheaves on XX admit an action of the Lie algebra of HH obtained by derivations, we found this fact surprising since there is no finiteness or Hausdorff assumptions in the cohomology groups. This also suggests that there is a deeper structure in the period sheaves of rigid spaces that witness the locally analytic properties of their cohomologies. In a work in progress of Johannes Anschütz, Arthur-César le Bras, Peter Scholze and the second author we expect to give a conceptual explanation of these facts via the analytic prismatization.

Geometric Sen theory over local Shimura varieties

In the next paragraph we state the main results of Section 4 extending those of [Pan22a] and [RJRC22] about the Sen operators of local Shimura varieties. In order to be more precise we need to introduce some notation, we shall follow [SW20]. Let (𝐆,b,μ)(\mathbf{G},b,\mu) be a local Shimura datum as in Lecture XXIV of loc. cit., let EE be the field of definition of μ\mu and E˘\breve{E} the completion of the maximal unramified extension of EE. Let FL𝐆,μ,E\operatorname{FL}_{\mathbf{G},\mu,E} and FL𝐆,μ1,E\operatorname{FL}_{\mathbf{G},\mu^{-1},E} be the algebraic flag varieties parametrizing decreasing and increasing μ\mu-filtrations of the trivial 𝐆\mathbf{G}-torsor respectively. We let 𝐆,μ,E\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu,E} and 𝐆,μ1,E\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu^{-1},E} be the analytification of the flag varieties to adic spaces [Hub94]. For K𝐆(p)K\subset\mathbf{G}(\mathbb{Q}_{p}) open compact subgroup we let 𝐆,b,μ,K,E˘\mathcal{M}_{\mathbf{G},b,\mu,K,\breve{E}} be the local Shimura variety over E˘\breve{E} at level KK.

Let E/EE^{\prime}/E be a finite extension where the group 𝐆\mathbf{G} is split and let us fix once and for all a cocharacter μ:𝔾m,E𝐆E\mu:\mathbb{G}_{m,E^{\prime}}\to\mathbf{G}_{E^{\prime}} representing the conjugacy class of μ\mu. Let 𝐏μ\mathbf{P}_{\mu} and 𝐏μ1\mathbf{P}_{\mu^{-1}} be the parabolic subgroups parametrizing decreasing and increasing filtrations of μ\mu, let 𝐍μ𝐏μ\mathbf{N}_{\mu}\subset\mathbf{P}_{\mu} and 𝐍μ1𝐏μ1\mathbf{N}_{\mu^{-1}}\subset\mathbf{P}_{\mu^{-1}} be their unipotent radicals respectively, and let 𝐌=𝐌μ=𝐌μ1\mathbf{M}=\mathbf{M}_{\mu}=\mathbf{M}_{\mu^{-1}} be the centralizer of μ\mu (eq. of μ1\mu^{-1}) in 𝐆E\mathbf{G}_{E^{\prime}}. We have presentations for the flag varieties FL𝐆,μ,E=𝐆E/𝐏μ\operatorname{FL}_{\mathbf{G},\mu,E^{\prime}}=\mathbf{G}_{E^{\prime}}/\mathbf{P}_{\mu} and FL𝐆,μ1,E=𝐆E/𝐏μ1\operatorname{FL}_{\mathbf{G},\mu^{-1},E^{\prime}}=\mathbf{G}_{E^{\prime}}/\mathbf{P}_{\mu^{-1}}, these presentations give rise to an equivalence of 𝐆\mathbf{G}-equivariant quasi-coherent sheaves on FL𝐆,μ,E\operatorname{FL}_{\mathbf{G},\mu,E^{\prime}} and FL𝐆,μ1,E\operatorname{FL}_{\mathbf{G},\mu^{-1},E^{\prime}} and algebraic representations of 𝐏μ\mathbf{P}_{\mu} and 𝐏μ1\mathbf{P}_{\mu^{-1}} respectively.

Let 𝔫μ𝔭μ𝔤\mathfrak{n}_{\mu}\subset\mathfrak{p}_{\mu}\subset\mathfrak{g} be the Lie algebras of 𝐍μ𝐏μ𝐆E\mathbf{N}_{\mu}\subset\mathbf{P}_{\mu}\subset\mathbf{G}_{E^{\prime}} and 𝔪μ=𝔭μ/𝔫μ\mathfrak{m}_{\mu}=\mathfrak{p}_{\mu}/\mathfrak{n}_{\mu} the Lie algebra of the Levi quotient. We see these Lie algebras endowed with the adjoint action of 𝐏μ\mathbf{P}_{\mu} and consider their corresponding 𝐆\mathbf{G}-equivariant Lie algebroids 𝔫μ0𝔭μ0𝔤μ0=𝒪FL𝐆,μ,Ep𝔤\mathfrak{n}^{0}_{\mu}\subset\mathfrak{p}^{0}_{\mu}\subset\mathfrak{g}^{0}_{\mu}=\mathscr{O}_{\operatorname{FL}_{\mathbf{G},\mu,E^{\prime}}}\otimes_{\mathbb{Q}_{p}}\mathfrak{g} and 𝔪μ0=𝔭μ0/𝔫μ0\mathfrak{m}^{0}_{\mu}=\mathfrak{p}^{0}_{\mu}/\mathfrak{n}^{0}_{\mu} appearing in the localization theory of Beilinson-Bernstein [BB81]. We denote in the same way their pullbacks to vector bundles over the analytic flag varieties, and use similar notation for the Lie algebras of the opposite parabolic and their associated Lie algebroids in FL𝐆,μ1,E\operatorname{FL}_{\mathbf{G},\mu^{-1},E^{\prime}}.

Let 𝐆,b,μ,,E˘=limK𝐆,b,μ,K,E˘\mathcal{M}_{\mathbf{G},b,\mu,\infty,\breve{E}}^{\lozenge}=\varprojlim_{K}\mathcal{M}_{\mathbf{G},b,\mu,K,\breve{E}}^{\lozenge} be the infinite level local Shimura variety seen as a diamond over E˘\breve{E} and consider the Grothendieck-Messing and Hodge-Tate period maps

𝐆,b,μ,,E˘{\mathcal{M}_{\mathbf{G},b,\mu,\infty,\breve{E}}^{\lozenge}}𝐆,μ,E˘{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu,\breve{E}}^{\lozenge}}𝐆,μ1,E˘.{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu^{-1},\breve{E}}^{\lozenge}.}πHT\scriptstyle{\pi_{\operatorname{\scriptsize HT}}}πGM\scriptstyle{\pi_{\operatorname{\scriptsize GM}}}

Let G~b\widetilde{G}_{b} be the group of automorphisms of the constant 𝐆\mathbf{G}-torsor b\mathcal{E}_{b} over the curve, see [FS24, III.5.1]. Then 𝐆,b,μ,,E˘\mathcal{M}_{\mathbf{G},b,\mu,\infty,\breve{E}}^{\lozenge} is endowed with an action of 𝐆(p)×G~b\mathbf{G}(\mathbb{Q}_{p})\times\widetilde{G}_{b} for which both maps πGM\pi_{\operatorname{\scriptsize GM}} and πHT\pi_{\operatorname{\scriptsize HT}} are equivariant in a suitable sense (see Section 3 for more details).

By [SW20, Corollary 23.3.2], when bb is basic there is a dual local Shimura datum (𝐆ˇ,bˇ,μˇ)(\check{\mathbf{G}},\check{b},\check{\mu}), an isomorphism Gb=G~b=𝐆ˇ(p)G_{b}=\widetilde{G}_{b}=\check{\mathbf{G}}(\mathbb{Q}_{p}) and a 𝐆(p)×𝐆ˇ(p)\mathbf{G}(\mathbb{Q}_{p})\times\check{\mathbf{G}}(\mathbb{Q}_{p})-equivariant isomorphism of infinite level local Shimura varieties

𝐆,b,μ,,E˘𝐆ˇ,bˇ,μˇ,,E˘\mathcal{M}_{\mathbf{G},b,\mu,\infty,\breve{E}}\cong\mathcal{M}_{\check{\mathbf{G}},\check{b},\check{\mu},\infty,\breve{E}}

that exchanges the Grothendieck-Messing and Hodge-Tate period maps (see also Proposition 3.2.3).

On the other hand, the map πK:𝐆,b,μ,,E˘𝐆,b,μ,K,E˘\pi_{K}:\mathcal{M}_{\mathbf{G},b,\mu,\infty,\breve{E}}^{\lozenge}\to\mathcal{M}_{\mathbf{G},b,\mu,K,\breve{E}}^{\lozenge} is a proétale KK-torsor. Thus, for any ind-system V=``limi"ViV=``\varinjlim_{i}"V_{i} of pp-adically complete continuous representations of KK we can construct a vv-sheaf V\mathcal{F}_{V} on 𝐆,b,μ,K,E˘\mathcal{M}_{\mathbf{G},b,\mu,K,\breve{E}}^{\lozenge} by first constructing the pp-complete vv-sheaves Vi\mathcal{F}_{V_{i}} via descent ViV_{i} along πK\pi_{K} and then by extending by colimits V=limiVi\mathcal{F}_{V}=\varinjlim_{i}\mathcal{F}_{V_{i}} (see Definition 3.3.3). In particular, for VV an algebraic representation of 𝐆\mathbf{G} we have automorphic local systems V\mathcal{F}_{V}, and for π\pi a smooth admissible representation of 𝐆(p)\mathbf{G}(\mathbb{Q}_{p}) over a pp-power torsion ring AA the sheaf π\mathcal{F}_{\pi} is the étale local system considered in [Sch18] for the pp-adic Jacquet-Langlands functor.

Let 𝔤\mathcal{F}_{\mathfrak{g}^{\vee}} be the local system over the local Shimura tower (𝐆,b,μ,K,E˘)K𝐆(p)(\mathcal{M}_{\mathbf{G},b,\mu,K,\breve{E}^{\prime}})_{K\subset\mathbf{G}(\mathbb{Q}_{p})} attached to the dual of the adjoint representation of 𝐆\mathbf{G}. Let us write by 𝒪\mathscr{O}_{\mathcal{M}} for the structural sheaf of a finite level local Shimura variety and let Ω1\Omega^{1}_{\mathcal{M}} be its cotangent bundle. We now state the first theorem concerning the computation of the geometric Sen operators of local Shimura varieties extending [RC24b, Theorem 5.2.5].

Theorem 1.4 (Theorem 4.3.1).

The geometric Sen operator θ:𝔤p𝒪^Ω1𝒪𝒪^(1)\theta_{\mathcal{M}}:\mathcal{F}_{\mathfrak{g}^{\vee}}\otimes_{\mathbb{Q}_{p}}\widehat{\mathscr{O}}\to\Omega^{1}_{\mathcal{M}}\otimes_{\mathscr{O}_{\mathcal{M}}}\widehat{\mathscr{O}}(-1) of the tower (𝐆,b,μ,K,E˘)K𝐆(p)(\mathcal{M}_{\mathbf{G},b,\mu,K,\breve{E}^{\prime}})_{K\subset\mathbf{G}(\mathbb{Q}_{p})} in the sense of [RC23, Theorem 3.3.4] is given by the pullback along πHT\pi_{\operatorname{\scriptsize HT}} of the 𝐆\mathbf{G}-equivariant map of vector bundles on 𝐆,μ1,E\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu^{-1},E^{\prime}}

𝔤0,𝔫μ10,\mathfrak{g}^{0,\vee}\to\mathfrak{n}^{0,\vee}_{\mu^{-1}}

where the identification πHT(𝔫μ10,)Ω1𝒪𝒪^(1)\pi^{*}_{\operatorname{\scriptsize HT}}(\mathfrak{n}^{0,\vee}_{\mu^{-1}})\cong\Omega^{1}_{\mathcal{M}}\otimes_{\mathscr{O}_{\mathcal{M}}}\widehat{\mathscr{O}}(-1) is through the opposite of the Kodaira-Spencer isomorphism, see Section 4.1. Here (n)\mathscr{F}(n) is the nn-th Hodge-Tate twist of \mathscr{F} by the nn-th power of the cyclotomic character.

Remark 1.3.

The previous theorem was stated for the base change of flag varieties and local Shimura varieties to EE^{\prime}. This base change can be avoided if one works without fixing a Hodge cocharacter μ\mu, indeed, the flag varieties and the Hodge-Tate period maps are already defined over EE. Moreover, the Lie algebroids 𝔫μ10\mathfrak{n}^{0}_{\mu^{-1}}, 𝔭μ10\mathfrak{p}^{0}_{\mu^{-1}} and 𝔪μ10\mathfrak{m}^{0}_{\mu^{-1}} are also defined over EE, see Remark 2.5.1.

A first consequence of the computation of the geometric Sen operator is the vanishing of the higher locally analytic vectors of the structural sheaf 𝒪^\widehat{\mathscr{O}} at infinite level, as well as the computation of the arithmetic Sen operator in terms of representation theory. Let C/EC/E^{\prime} be a completed algebraically closed extension and consider the CC-base change 𝐆,b,μ,K,C\mathcal{M}_{\mathbf{G},b,\mu,K,C} of the local Shimura varieties. Let 𝒱K=Cla(K,p)1\mathcal{V}_{K}=C^{la}(K,\mathbb{Q}_{p})_{\star_{1}} be the space of locally analytic functions of KK endowed with the left regular action and let 𝒱K\mathcal{F}_{\mathcal{V}_{K}} be the vv-sheaf over 𝐆,b,μ,K,C\mathcal{M}_{\mathbf{G},b,\mu,K,C} obtained by descent from infinite level. We have the following theorem, analogue to Proposition 6.2.8, Corollary 6.2.12 and Theorem 6.3.5 of [RC24b].

Theorem 1.5 (Theorem 4.3.3).

Let U𝐆,b,μ,K,CU\subset\mathcal{M}_{\mathbf{G},b,\mu,K,C} be an open affinoid subspace admitting an étale map to a product of tori 𝕋Cd\mathbb{T}^{d}_{C} that factors as a finite composition of rational localizations and finite étale maps. Let U𝐆,b,μ,,CU_{\infty}\subset\mathcal{M}_{\mathbf{G},b,\mu,\infty,C}^{\lozenge} be the pullback of UU, then the vv-cohomology

RΓv(U,𝒪^^p𝒱K)R\Gamma_{v}(U,\widehat{\mathscr{O}}\widehat{\otimes}_{\mathbb{Q}_{p}}\mathcal{F}_{\mathcal{V}_{K}}) (1.2)

sits in degree 0 and is equal to the locally analytic vectors 𝒪^(U)Gla\widehat{\mathscr{O}}(U_{\infty})^{G-la} of 𝒪^(U)\widehat{\mathscr{O}}(U_{\infty}). Here the completed tensor product is a filtered colimit of pp-completed tensor products obtained by writing 𝒱K\mathcal{F}_{\mathcal{V}_{K}} as a colimit of Banach sheaves (equivalently a solid tensor product as in [AM24]).

Furthermore, the action of 𝔤μ10=𝒪𝐆,μ1,Cp𝔤\mathfrak{g}^{0}_{\mu^{-1}}=\mathscr{O}_{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu^{-1},C}}\otimes_{\mathbb{Q}_{p}}\mathfrak{g} on 𝒪^(U)Gla\widehat{\mathscr{O}}(U_{\infty})^{G-la} by derivations kills 𝔫μ10\mathfrak{n}^{0}_{\mu^{-1}}. Thus, we have an horizontal action of 𝔪μ10\mathfrak{m}^{0}_{\mu^{-1}} on 𝒪^(U)Gla\widehat{\mathscr{O}}(U_{\infty})^{G-la}. Moreover, the space 𝒪^(U)Gla\widehat{\mathscr{O}}(U_{\infty})^{G-la} has an arithmetic Sen operator as in [RC24b, Theorem 6.3.5] given by the opposite of the derivative of the Hodge cocharacter θμ=θμ1𝔪μ10-\theta_{\mu}=\theta_{\mu^{-1}}\in\mathfrak{m}^{0}_{\mu^{-1}}.

Locally analytic vectors of local Shimura varieties

In this paragraph we apply 1.3 to prove the independence of locally analytic vectors for a duality of local Shimura varieties, generalizing a theorem of Pan for the Lubin-Tate tower of 𝐆𝐋2\mathbf{GL}_{2} [Pan22b, Corollary 5.3.9]. Let C/EC/E be a complete algebraically closed field, we write G=𝐆(p)G=\mathbf{G}(\mathbb{Q}_{p}) and let GbG_{b} be the profinite quotient of G~b\widetilde{G}_{b}.

Let 𝒪^\widehat{\mathscr{O}}_{\mathcal{M}} be the restriction of the structural sheaf 𝒪^\widehat{\mathscr{O}} in the vv-site of 𝐆,μ,b,,C\mathcal{M}_{\mathbf{G},\mu,b,\infty,C}^{\lozenge} to the underlying topological space |𝐆,μ,b,,C||\mathcal{M}_{\mathbf{G},\mu,b,\infty,C}^{\lozenge}|. Let 𝒪Gla𝒪^\mathscr{O}^{G-la}_{\mathcal{M}}\subset\widehat{\mathscr{O}}_{\mathcal{M}} be the subsheaf whose values in a qcqs open subspace U𝐆,μ,b,,CU_{\infty}\subset\mathcal{M}_{\mathbf{G},\mu,b,\infty,C}^{\lozenge} are given by the G=𝐆(p)G=\mathbf{G}(\mathbb{Q}_{p})-locally analytic sections of 𝒪^\widehat{\mathscr{O}}_{\mathcal{M}}, namely, given by

𝒪Gla(U)=𝒪^(U)KUla\mathscr{O}^{G-la}_{\mathcal{M}}(U_{\infty})=\widehat{\mathscr{O}}(U_{\infty})^{K_{U_{\infty}}-la}

where KUGK_{U_{\infty}}\subset G is the stabilizer of UU_{\infty}. We have the following corollary:

Corollary 1.2 (5.1.9).

For any pp-adic Lie group HG~bH\subset\widetilde{G}_{b} and any qcqs open subspace U𝐆,b,μ,,CU_{\infty}\subset\mathcal{M}^{\lozenge}_{\mathbf{G},b,\mu,\infty,C} the natural map

𝒪Gla(U)RHla𝒪Gla(U)\mathscr{O}^{G-la}_{\mathcal{M}}(U_{\infty})^{RH-la}\xrightarrow{\sim}\mathscr{O}^{G-la}_{\mathcal{M}}(U_{\infty})

from the derived HH-locally analytic vectors is an equivalence. In particular, if bb is basic we have an equality of subsheaves of 𝒪^\widehat{\mathscr{O}}_{\mathcal{M}}

𝒪Gla=𝒪Gbla.\mathscr{O}^{G-la}_{\mathcal{M}}=\mathscr{O}^{G_{b}-la}_{\mathcal{M}}.

More generally, for bb basic and I(0,)I\subset(0,\infty) a compact interval with rational ends, we have an equivalence of derived solid locally analytic representations of G×GbG\times G_{b}

RΓv(U,𝔹I)RGblaRΓv(U,𝔹I)RG×GblaRΓv(U,𝔹I)RGla.R\Gamma_{v}(U_{\infty},\mathbb{B}_{I})^{RG_{b}-la}\xrightarrow{\sim}R\Gamma_{v}(U_{\infty},\mathbb{B}_{I})^{RG\times G_{b}-la}\xleftarrow{\sim}R\Gamma_{v}(U_{\infty},\mathbb{B}_{I})^{RG-la}.

From now on we shall focus in the case when bb is basic. We will identify the GG and GbG_{b}-locally analytic vectors of the structural sheaf 𝒪^𝒲\widehat{\mathscr{O}}_{\mathcal{W}} at infinite level and simply write 𝒪𝒲la\mathscr{O}^{la}_{\mathcal{W}}. We can then identify the horizontal actions of the Levi Lie algebras of 1.5. For this, we need some additional notation.

By Corollary 3.3.7 we have a natural G×GbG\times G_{b}-equivariant isomorphism of 𝒪^\widehat{\mathscr{O}}_{\mathcal{M}}-vector bundles on 𝐆,b,μ,,C\mathcal{M}^{\lozenge}_{\mathbf{G},b,\mu,\infty,C}

𝔪μ10𝒪𝐆,μ1,C𝒪^𝒪^𝒪𝐆,μ,C𝔪μ0.\mathfrak{m}^{0}_{\mu^{-1}}\otimes_{\mathscr{O}_{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu^{-1},C}}}\widehat{\mathscr{O}}_{\mathcal{M}}\cong\widehat{\mathscr{O}}_{\mathcal{M}}\otimes_{\mathscr{O}_{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu,C}}}\mathfrak{m}^{0}_{\mu}.

By taking locally analytic vectors we obtain an 𝒪la\mathscr{O}^{la}_{\mathcal{M}}-vector bundle which we shall denote as 𝔪0,la\mathfrak{m}^{0,la}, endowed with G×GbG\times G_{b}-equivariant isomorphisms

𝔪μ10𝒪𝐆,μ1,C𝒪la𝔪0,la𝒪la𝒪𝐆,μ,C𝔪μ0.\mathfrak{m}^{0}_{\mu^{-1}}\otimes_{\mathscr{O}_{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu^{-1},C}}}\mathscr{O}^{la}_{\mathcal{M}}\cong\mathfrak{m}^{0,la}\cong\mathscr{O}^{la}_{\mathcal{M}}\otimes_{\mathscr{O}_{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu,C}}}\mathfrak{m}^{0}_{\mu}. (1.3)

We have the following theorem:

Theorem 1.6 (Theorem 5.1.10).

The actions of 𝔫μ0\mathfrak{n}_{\mu}^{0} and 𝔫μ10\mathfrak{n}_{\mu^{-1}}^{0} on 𝒪la\mathscr{O}^{la}_{\mathcal{M}} vanish. Furthermore, the actions of 𝔪μ0\mathfrak{m}^{0}_{\mu} and 𝔪μ10\mathfrak{m}^{0}_{\mu^{-1}} on 𝒪la\mathscr{O}^{la}_{\mathcal{M}} by derivations are identified via (1.3). In particular, the central character of the actions of 𝔪μ0\mathfrak{m}^{0}_{\mu} and 𝔪μ10\mathfrak{m}^{0}_{\mu^{-1}} on 𝒪la\mathscr{O}^{la}_{\mathcal{M}} agree under the natural isomorphism of the center of the enveloping algebras 𝒵(𝔪μ,C)𝒵(𝔪μ1,C)\mathcal{Z}(\mathfrak{m}_{\mu,C})\cong\mathcal{Z}(\mathfrak{m}_{\mu^{-1},C}), where 𝔪μ,C\mathfrak{m}_{\mu,C} and 𝔪μ1,C\mathfrak{m}_{\mu^{-1},C} are the Levi subalgebras of LieGbpC\operatorname{Lie}G_{b}\otimes_{\mathbb{Q}_{p}}C and LieGpC\operatorname{Lie}G\otimes_{\mathbb{Q}_{p}}C respectively.

De Rham cohomology of towers of local Shimura varieties

Our next result is the comparison between compactly supported de Rham cohomologies of the two towers in a duality of local Shimura varieties. This theorem has been also independently obtained by Guido Bosco, Wiesława Nizioł and the first author. Let (𝐆,b,μ)(\mathbf{G},b,\mu) be a local Shimura datum with bb basic and let (𝐆ˇ,bˇ,μˇ)(\check{\mathbf{G}},\check{b},\check{\mu}) be the dual local Shimura datum. Consider the towers of rigid spaces (𝐆,b,μ,K,E˘)K𝐆(p)(\mathcal{M}_{\mathbf{G},b,\mu,K,\breve{E}})_{K\subset\mathbf{G}(\mathbb{Q}_{p})} and (𝐆ˇ,bˇ,μˇ,Kˇ,E˘)K𝐆ˇ(p)(\mathcal{M}_{\check{\mathbf{G}},\check{b},\check{\mu},\check{K},\breve{E}})_{K\subset\check{\mathbf{G}}(\mathbb{Q}_{p})}. We have the following theorem:

Theorem 1.7 (Theorem 5.2.2).

There is a natural 𝐆(p)×𝐆ˇ(p)\mathbf{G}(\mathbb{Q}_{p})\times\check{\mathbf{G}}(\mathbb{Q}_{p})-equivariant isomorphism of compactly supported de Rham cohomology groups

limK𝐆(p)HdR,ci(𝐆,b,μ,K,C)limKˇ𝐆ˇ(p)HdR,ci(𝐆ˇ,bˇ,μˇ,Kˇ,C).\varinjlim_{K\subset\mathbf{G}(\mathbb{Q}_{p})}H^{i}_{dR,c}(\mathcal{M}_{\mathbf{G},b,\mu,K,C})\cong\varinjlim_{\check{K}\subset\check{\mathbf{G}}(\mathbb{Q}_{p})}H^{i}_{dR,c}(\mathcal{M}_{\check{\mathbf{G}},\check{b},\check{\mu},\check{K},C}).
Remark 1.4.

1.7 should be seen as an evidence of the fact that there is a well defined analytic de Rham stack (in the sense of [RC24a]) for the infinite level Shimura variety, together with 𝐆(p)×𝐆ˇ(p)\mathbf{G}(\mathbb{Q}_{p})\times\check{\mathbf{G}}(\mathbb{Q}_{p})-equivariant equivalences

limK𝐆,b,μ,K,E˘dR=𝐆,b,μ,,E˘,dR𝐆ˇ,bˇ,μˇ,,E˘,dR=limKˇ𝐆ˇ,bˇ,μˇ,Kˇ,E˘dR.\varprojlim_{K}\mathcal{M}_{\mathbf{G},b,\mu,K,\breve{E}}^{\operatorname{dR}}=\mathcal{M}^{\lozenge,\operatorname{dR}}_{\mathbf{G},b,\mu,\infty,\breve{E}}\cong\mathcal{M}^{\lozenge,\operatorname{dR}}_{\check{\mathbf{G}},\check{b},\check{\mu},\infty,\breve{E}}=\varprojlim_{\check{K}}\mathcal{M}_{\check{\mathbf{G}},\check{b},\check{\mu},\check{K},\breve{E}}^{\operatorname{dR}}.

Indeed, as it was explained by Scholze to the second author, one can prove sufficient descent for the formation of the analytic de Rham stack to be well defined for (suitable nice) diamonds, where the previous equivalence holds as analytic stacks. It is likely that purely motivic techniques as those appearing in [Vez19] are enough to show the equivalence of the de Rham cohomologies for the two towers, see Proposition 4.5 in loc. cit.; we thanks Arthur-César le Bras for this observation.

Locally analytic Jacquet-Langlands functor in the Lubin Tate case

We finish the presentation of the main results with the principal motivation that initiated this project, that is, the pp-adic Jacquet-Langlands functor of the Lubin-Tate tower treated in [Sch18]. We shall keep the notation of the beginning of the introduction regarding the Lubin-Tate and Drinfeld towers. We have the following compatibility with the passage to locally analytic vectors:

Theorem 1.8 (Theorem 5.3.6).

Let π\pi be an admissible Banach representation of 𝐆𝐋n(F)\mathbf{GL}_{n}(F) and let Π=(π[1p])𝐆𝐋n(F)la\Pi=(\pi[\frac{1}{p}])^{\mathbf{GL}_{n}(F)-la} be the space of locally analytic vectors seen as a colimit of Banach spaces. Let Π\mathcal{F}_{\Pi} be the proétale sheaf over pn1\mathbb{P}^{n-1}_{\mathbb{C}_{p}} constructed via descent along πGM\pi_{\operatorname{\scriptsize GM}} of the continuous representation Π\Pi. There is a natural equivalence of solid locally analytic HH-representations

(𝒥(π)[1π])RHlaRΓproe´t(pn1,Π).(\mathcal{JL}(\pi)[\frac{1}{\pi}])^{RH-la}\cong R\Gamma_{\operatorname{\scriptsize pro\acute{e}t}}(\mathbb{P}^{n-1}_{\mathbb{C}_{p}},\mathcal{F}_{\Pi}).

Furthermore, this equivalence induces an isomorphism of cohomology groups:

(𝒥i(π)[1p])HlaHproe´ti(pn1,Π).(\mathcal{JL}^{i}(\pi)[\frac{1}{p}])^{H-la}\cong H^{i}_{\operatorname{\scriptsize pro\acute{e}t}}(\mathbb{P}^{n-1}_{\mathbb{C}_{p}},\mathcal{F}_{\Pi}).

As a corollary of Theorems 1.6 and 1.8 we can show that the Jacquet-Langlands functor preserves central characters for the locally analytic vectors of admissible Banach representations.

Corollary 1.3 (Corollary 5.3.7).

Let π\pi be an admissible Banach representation of 𝐆𝐋n(L)\mathbf{GL}_{n}(L) over a finite extension of p\mathbb{Q}_{p} and suppose that Π=π𝐆𝐋n(L)la\Pi=\pi^{\mathbf{GL}_{n}(L)-la} has central character χ\chi. Then, for all ii\in\mathbb{Z}, the locally analytic HH-representation 𝒥i(π)Hla\mathcal{JL}^{i}(\pi)^{H-la} has central character χ\chi under the natural identification 𝒵(LieH)𝒵(LieG)\mathcal{Z}(\operatorname{Lie}H)\cong\mathcal{Z}(\operatorname{Lie}G).

Outline of the paper

Section 2 is a preliminary section where we introduce the main objects and tools used in the paper. In Section 2.1 we recall the definition of period sheaves and the construction of the different incarnations of families of Fargues-Fontaine curves following [SW20, §11.2] and [FS24, §II.1]. In 2.2 we briefly recall the construction of the categories of solid almost quasi-coherent sheaves on diamonds of [Man22b], and its relation with smooth representation theory of profinite groups. In 2.3 we recall the construction of the décalage operator of [BMS18] which will be relevant to perform a technical dévisage in the proof of 1.3. Then, in Section 2.4 we briefly recall the basics of the theory of solid locally analytic representations of [RJRC22, RJRC23], in particular we state the locally analytic criterion of Lemma 2.4.1 which is key in the proof of 1.3. Finally, in Section 2.5 we briefly summarize the relationship between representations of reductive groups and equivariant sheaves over flag varieties, making special emphasis in the Lie algebroids appearing in the localization theory of Beilinson-Bernstein [BB81].

We continue with Section 3 which concerns the definition of the local Shimura varieties and some basic Hodge theoretic features of them, we follow [SW20, Lecture XXIII] and [FS24, §III.4 and 5]. In Section 3.1 we discuss some facts about torsors on families of Fargues-Fontaine curves. In Section 3.2 we recall the definition of the moduli space of shtukas of one leg as well as the construction of the Grothendieck-Messing and Hodge-Tate period maps. Then, in Section 3.3 we specialize the previous construction to the situation of local Shimura varieties where we deduce from the general theory of [SW20, Lecture XXIII] a pp-adic Riemann-Hilbert correspondence for automorphic local systems in Proposition 3.3.4; this formulation of the theory of Scholze-Weinstein will be useful in the computation of the geometric Sen operator of the next section.

Next, in Section 4 we compute the geometric and arithmetic Sen operators for local Shimura varieties. In Section 4.1 we explain the purely representation theoretic construction of the Kodaira-Spencer isomorphism for Shimura varieties which is essentially a reinterpretation of the anchor map of the reductive group acting on the flag variety. This point of view of the Kodaira-Spencer map will allow us to compute the pullback of equivariant sheaves of flag varieties via the Hodge-Tate period maps in terms of automorphic vector bundles and the Faltings extension in Section 4.2, see Theorem 4.2.1. Finally, we use this description of the pullbacks of automorphic vector bundles to compute the geometric and arithmetic Sen operators of Theorems 1.4 and 1.5. The analogue of these theorems for global Shimura varieties were achieved in [RC23, RC24b], and the proofs in the local situation follow exactly the same line of arguments.

We conclude with Section 5 were most of the main theorems stated in the introduction are proven. In Section 5.1 we prove 1.3; the strategy of the proof is to use the locally analytic criterion of Lemma 2.4.1. For this one has to implement a long dévisage up to the point where one is reduced to showing that the proétale (eq. vv-) cohomology of 𝒪+/p\mathscr{O}^{+}/p on a qcqs smooth rigid space is small after applying a décalage operator LηpεL\eta_{p^{\varepsilon}} for some ε>0\varepsilon>0, see 4. As an immediate consequence we obtain 1.1. We conclude this section with the application of 1.3 to local Shimura varieties; we first obtain the independence of locally analytic vectors at infinite level for bb basic of 1.2, then, with a more careful study of the horizontal actions arising from the two towers, we prove 1.6. In Section 5.2 we compare the de Rham cohomology of the two towers proving 1.7; here the strategy is to relate the de Rham complexes of each tower with a suitable de Rham complex of the sheaf 𝒪la\mathscr{O}^{la}_{\mathcal{M}} arising from the derivations of both groups GG and GbG_{b}. Finally, in Section 5.3 we prove the compatibility of Scholze’s Jacquet-Langlands functor with the passage to locally analytic vectors for admissible Banach representations proving 1.8; here the key strategy is to rewrite the proétale cohomology of the sheaf π\mathcal{F}_{\pi} in terms of period sheaves 𝔹I\mathbb{B}_{I} and then to exploit the independence of locally analytic vectors at infinite level of 1.2 in order to jump between towers. Finally, using the proof of 1.8 and the compatibility of the horizontal characters for the sheaf 𝒪la\mathscr{O}^{la}_{\mathcal{M}} of 1.6, we obtain the compatibility of central characters of the Jacquet-Langlands functor of 1.3.

Conventions

In this paper we use the vv-site of perfectoid spaces as introduced in [Sch22]. We use the theory of solid almost quasi-coherent sheaves of [Man22b], [AM24] and [AMLB]; the use of these cohomology theories is important in order to properly keep track to the condensed or topological structure of cohomology complexes. In particular, this work heavily depends on the theory of condensed mathematics of Clausen and Scholze [CS19, CS20], and in higher category theory for which we refer to [Lur09, Lur17]. A different reason to use condensed mathematics is to have access to the theory of solid locally analytic representations of [RJRC22, RJRC23]. This is important since, even though most of the main theorems involve classical topological representations, the proofs will make appear very general solid representations which are not classical.

Acknowledgements

We thank Johannes Anschütz, Guido Bosco, Arthur-César le Bras, Pierre Colmez, Wiesława Nizioł, Lue Pan and Peter Scholze for enlightening conversations in different stages of this work. We also thank Johannes Anschütz, Guido Bosco and Arthur-César le Bras for corrections and comments in a first draft of the paper. Part of this project was done during the trimester program in Bonn: “The Arithmetic of The Langlands Program” during the summer of 2023, we heartily thank the organizers and the Hausdorff Research Institute for Mathematics for the excellent environment for mathematical discussions and exchanges. The second author wants to thank Columbia University and the Simons Society of Fellows for the wonderful working conditions and support as a postdoc and Junior Fellow.

2. Preliminaries

In this section we introduce the main objects and techniques used in the paper. In Section 2.1 we recall the definition of families of Fargues-Fontaine curves following [SW20, §11.2] and [FS24, §II.1]. Some period sheaves associated to affinoid charts of the curves will be explicitly introduced for reference in later sections. Then in Section 2.2, we recall the definition of the derived categories of solid almost quasi-coherent sheaves of [Man22b]; we shall not need all the power of the six functor formalism, only the existence of these categories and their relation with smooth representations after [Man22b, §3.4]. We continue in Section 2.3 with some basic properties of the décalage operator LηL\eta_{\mathcal{I}} of [BMS18, §6]; their importance for us will be to kill some small enough torsion in order to apply a locally analytic criterion discussed in the next section. In Section 2.4 we briefly introduce the theory of solid locally analytic representations of [RJRC22, RJRC23], in particular we recall the criterion of [RJRC23, Proposition 3.3.3] for a solid representation of a pp-adic Lie group to be locally analytic. Finally, in Section 2.5 we state the classical dictionary between representation theory and equivariant quasi-coherent sheaves on flag varieties; in particular we make emphasis in the Lie algebroids over the flag variety appearing in the localization theory of [BB81].

Apart (but not disjoint!) from condensed mathematics, we also use different aspects of pp-adic Hodge theory. The main objects we study are period sheaves on diamonds [RC23] and the computation of the geometric Sen operators of Shimura varieties [RC24b]. Finally, in order to realize our cohomology groups as honest solid abelian groups we use the categories of solid quasi-coherent sheaves of diamonds of Mann [Man22b] though the full six functor formalism will not be necessary.

2.1. The Fargues-Fontaine curve and sheaves of periods

Let Perfd\operatorname{Perfd} be the category of perfectoid spaces over p\mathbb{Z}_{p} and let PerfPerfd\operatorname{Perf}\subset\operatorname{Perfd} be the full subcategory of perfectoid spaces over 𝔽p\mathbb{F}_{p}. Following [Sch22], we consider the vv-site Perfdv\operatorname{Perfd}_{v} of perfectoid spaces. Let 𝔽p((ϖ1/p))\mathbb{F}_{p}((\varpi^{1/p^{\infty}})) be the perfectoid field parametrizing pseudo-uniformizers in Perf\operatorname{Perf} and let Perfϖ\operatorname{Perf}_{\varpi} be the slice category Perf/Spa𝔽p((ϖ1/p))\operatorname{Perf}_{/\operatorname{Spa}\mathbb{F}_{p}((\varpi^{1/p^{\infty}}))}, equivalently, Perfϖ\operatorname{Perf}_{\varpi} is the category of perfectoid spaces in characteristic pp with fixed pseudo-uniformizer ϖ\varpi, and with maps preserving the pseudo-uniformizer. We let 𝒪^\widehat{\mathscr{O}} and 𝒪^+\widehat{\mathscr{O}}^{+} be the vv-sheaves on Perfd\operatorname{Perfd} mapping an affinoid perfectoid Spa(R,R+)\operatorname{Spa}(R,R^{+}) to RR and R+R^{+} respectively. Similarly, we let 𝒪\mathscr{O}^{\flat} and 𝒪,+\mathscr{O}^{\flat,+} be the vv-sheaves mapping Spa(R,R+)\operatorname{Spa}(R,R^{+}) to RR^{\flat} and R,+R^{\flat,+} respectively. Given S=Spa(R,R+)PerfdS=\operatorname{Spa}(R,R^{+})\in\operatorname{Perfd} we let 𝔸inf(S):=𝔸inf(R+):=W(R,+)\mathbb{A}_{\inf}(S):=\mathbb{A}_{\inf}(R^{+}):=W(R^{\flat,+}) be the period ring of Fontaine, and denote by []:R,+𝔸inf(S)[-]:R^{\flat,+}\to\mathbb{A}_{\inf}(S) the Teichmüller lift.

For S=Spa(R,R+)PerfϖS=\operatorname{Spa}(R,R^{+})\in\operatorname{Perf}_{\varpi} consider the sous-perfectoid analytic adic space [SW20, Proposition 11.2.1]

𝒴SFF:={|[ϖ]|0}Spa(𝔸inf(S)),\mathcal{Y}_{S}^{\operatorname{\scriptsize FF}}:=\{|[\varpi]|\neq 0\}\subset\operatorname{Spa}(\mathbb{A}_{\inf}(S)),

consisting on the locus where [ϖ][\varpi] is a pseudo-uniformizer, we call this adic space the 𝒴FF\mathcal{Y}^{\operatorname{\scriptsize FF}}-curve over SS.

The space 𝒴SFF\mathcal{Y}_{S}^{\operatorname{\scriptsize FF}} has pseudo-uniformizer [ϖ][\varpi]. Furthermore, the following properties hold (see [SW20, §11] and [FS24, §II.1])

  • We have a natural equivalence of diamonds

    𝒴SFF,=S×Spdp\mathcal{Y}_{S}^{\operatorname{\scriptsize FF},\lozenge}=S\times\operatorname{Spd}\mathbb{Z}_{p}

    where Spdp\operatorname{Spd}\mathbb{Z}_{p} is the diamond parametrizing untilts of perfectoid spaces. In particular the formation of 𝒴SFF\mathcal{Y}_{S}^{\operatorname{\scriptsize FF}} is independent of the pseudo-uniformizer of SS. Moreover, 𝒴SFF\mathcal{Y}_{S}^{\operatorname{\scriptsize FF}} has a natural Frobenius automorphism φS\varphi_{S} lifting the Frobenius of SS.

  • Let |𝒴SFF||\mathcal{Y}_{S}^{\operatorname{\scriptsize FF}}| be the underlying topological space of the adic space. There is a (unique) continuous radius map

    rad:|𝒴SFF|[0,)\operatorname{rad}:|\mathcal{Y}_{S}^{\operatorname{\scriptsize FF}}|\to[0,\infty)

    sending a rank 11 point xx to

    rad(x)=logp|[ϖ]|x/logp|p|x=log|p|x|[ϖ]|x.\operatorname{rad}(x)=\log_{p}|[\varpi]|_{x}/\log_{p}|p|_{x}=\log_{|p|_{x}}|[\varpi]|_{x}.

    The radius map and the Frobenius endomorphism are related by the formula

    radφS=prad.\operatorname{rad}\circ\varphi_{S}=p\operatorname{rad}.
  • For I=[s,r][0,)I=[s,r]\subset[0,\infty) a compact interval with rational ends one defines affinoid subspaces

    𝒴S,IFF=𝒴SFF(p[ϖ]1/r,[ϖ]1/sp).\mathcal{Y}_{S,I}^{\operatorname{\scriptsize FF}}=\mathcal{Y}_{S}^{\operatorname{\scriptsize FF}}(\frac{p}{[\varpi]^{1/r}},\frac{[\varpi]^{1/s}}{p}).

    One has rad(|𝒴S,IFF|)I\operatorname{rad}(|\mathcal{Y}_{S,I}^{\operatorname{\scriptsize FF}}|)\subset I but the inclusion 𝒴S,IFFrad1(I)\mathcal{Y}^{\operatorname{\scriptsize FF}}_{S,I}\subsetneq\operatorname{rad}^{-1}(I) is strict.

  • For I(0,)I\subset(0,\infty) an open interval we let 𝒴S,IFF𝒴SFF\mathcal{Y}_{S,I}^{\operatorname{\scriptsize FF}}\subset\mathcal{Y}_{S}^{\operatorname{\scriptsize FF}} be the inverse image of II by rad\operatorname{rad}. This space can be also described as

    𝒴S,I=JI𝒴S,JFF\mathcal{Y}_{S,I}=\bigcup_{J\subset I}\mathcal{Y}_{S,J}^{\operatorname{\scriptsize FF}}

    where JJ runs over all the compact subintervals of II with rational ends.

  • The relative Fargues-Fontaine curve over SS is the sous-perfectoid space

    𝒳SFF=𝒴S,(0,)FF/φS.\mathcal{X}_{S}^{\operatorname{\scriptsize FF}}=\mathcal{Y}_{S,(0,\infty)}^{\operatorname{\scriptsize FF}}/\varphi_{S}^{\mathbb{Z}}.

    We shall call this space the 𝒳FF\mathcal{X}^{\operatorname{\scriptsize FF}}-curve over SS.

Convention 2.1.1.

From now on all closed rational intervals I=[s,r][0,)I=[s,r]\subset[0,\infty) will be assumed to have rational ends.

Lemma 2.1.2.

Let S,SPerfϖS,S^{\prime}\in\operatorname{Perf}_{\varpi} be affinoid perfectoids of characteristic pp with fixed pseudo-uniformizer ϖ\varpi, let f:SSf:S^{\prime}\to S be a map in Perfϖ\operatorname{Perf}_{\varpi} and let fFF:𝒴SFF𝒴SFFf^{\operatorname{\scriptsize FF}}:\mathcal{Y}^{\operatorname{\scriptsize FF}}_{S}\to\mathcal{Y}^{\operatorname{\scriptsize FF}}_{S^{\prime}} be the corresponding map of 𝒴FF\mathcal{Y}^{\operatorname{\scriptsize FF}}-curves.

  1. (1)

    We have equivalences of sites 𝒴S,e´tFF𝒴S,e´tFF,\mathcal{Y}_{S,{\rm\acute{e}t}}^{\operatorname{\scriptsize FF}}\cong\mathcal{Y}_{S,{\rm\acute{e}t}}^{\operatorname{\scriptsize FF},\lozenge} and 𝒴S,fe´tFF𝒴S,fe´tFF,\mathcal{Y}^{\operatorname{\scriptsize FF}}_{S,\operatorname{\scriptsize f\acute{e}t}}\cong\mathcal{Y}^{\operatorname{\scriptsize FF},\lozenge}_{S,\operatorname{\scriptsize f\acute{e}t}}.

  2. (2)

    fFFf^{\operatorname{\scriptsize FF}} is an open immersion if and only if ff is so. If ff is a rational localization then so is fFFf^{\operatorname{\scriptsize FF}}.

  3. (3)

    For I[0,)I\subset[0,\infty) a closed or open interval, the map ff induces a cartesian square

    𝒴S,IFF{\mathcal{Y}_{S^{\prime},I}^{\operatorname{\scriptsize FF}}}𝒴S,IFF{\mathcal{Y}_{S,I}^{\operatorname{\scriptsize FF}}}𝒴SFF{\mathcal{Y}_{S^{\prime}}^{\operatorname{\scriptsize FF}}}𝒴SFF.{\mathcal{Y}_{S}^{\operatorname{\scriptsize FF}}.}
Proof.

Part (1) is a particular case of [Sch22, Lemma 15.6]. The first assertion of (2) follows from (1) and the fact that open immersions of analytic adic spaces are the same as étale maps f:YXf:Y\to X such that the diagonal YY×XYY\to Y\times_{X}Y is an equivalence. The second assertion of (2) follows from the fact that if S=S(f1,,fng)S^{\prime}=S\left(\frac{f_{1},\ldots,f_{n}}{g}\right) then

𝒴SFF=𝒴SFF([f1],,[fn][g]).\mathcal{Y}^{\operatorname{\scriptsize FF}}_{S^{\prime}}=\mathcal{Y}^{\operatorname{\scriptsize FF}}_{S}\left(\frac{[f_{1}],\ldots,[f_{n}]}{[g]}\right).

For part (3) it suffices to deal with the case where II is closed, then it follows from the definition of 𝒴S,IFF\mathcal{Y}_{S,I}^{\operatorname{\scriptsize FF}} as it is a rational localization involving only pp and [ϖ][\varpi]. ∎

Some of the main players in this paper are the sheaves of periods defined by the affinoid subspaces 𝒴S,IFF\mathcal{Y}^{\operatorname{\scriptsize FF}}_{S,I} for II closed.

Definition 2.1.3.

Let I[0,)I\subset[0,\infty) be a compact interval. We define the vv-sheaves 𝔹I\mathbb{B}_{I} and 𝔹I+\mathbb{B}_{I}^{+} on Perfϖ\operatorname{Perf}_{\varpi} to be the vv-sheafification of the presheaves mapping an affinoid perfectoid S=Spa(R,R+)PerfϖS=\operatorname{Spa}(R,R^{+})\in\operatorname{Perf}_{\varpi} to the rings

𝔹I+(S)=𝔹I+(R,R+):=𝒪+(𝒴S,IFF),\displaystyle\mathbb{B}_{I}^{+}(S)=\mathbb{B}_{I}^{+}(R,R^{+}):=\mathscr{O}^{+}(\mathcal{Y}_{S,I}^{\operatorname{\scriptsize FF}}),
𝔹I(S)=𝔹I(R,R+):=𝒪(𝒴S,IFF).\displaystyle\mathbb{B}_{I}(S)=\mathbb{B}_{I}(R,R^{+}):=\mathscr{O}(\mathcal{Y}_{S,I}^{\operatorname{\scriptsize FF}}).

We also define the vv-sheaf 𝔸inf\mathbb{A}_{\inf} as the sheaf mapping SS to 𝔸inf(S)\mathbb{A}_{\inf}(S).

In the following lemma we consider almost mathematics with respect to the (p,[ϖ])(p,[\varpi])-completion of the 𝔸inf\mathbb{A}_{\inf}-ideal generated by ([ϖ]1/pn)n([\varpi]^{1/p^{n}})_{n\in\mathbb{N}}.

Lemma 2.1.4.

Let S=Spa(R,R+)PerfϖS=\operatorname{Spa}(R,R^{+})\in\operatorname{Perf}_{\varpi} be affinoid perfectoid.

  1. (1)

    We have an almost equivalence of derived (p,[ϖ])(p,[\varpi])-adically complete complexes

    𝔸inf(R+)=aRΓv(S,𝔸inf).\mathbb{A}_{\inf}(R^{+})=^{a}R\Gamma_{v}(S,\mathbb{A}_{\inf}).
  2. (2)

    Let I=[0,r]I=[0,r], then we have an isomorphism of vv-sheaves

    𝔹I+/([ϖ1/r])=𝒪^+/(ϖ1/r)[T]\mathbb{B}_{I}^{+}/([\varpi^{1/r}])=\widehat{\mathscr{O}}^{+}/(\varpi^{1/r})[T]

    where the RHS term is a polynomial algebra with TT the residue class of p/[ϖ]1/rp/[\varpi]^{1/r}.

  3. (3)

    Suppose that I=[0,r]I=[0,r], we have a natural almost equivalence of derived [ϖ][\varpi]-adically complete complexes

    𝔹I+(S)=aRΓv(S,𝔹I+).\mathbb{B}_{I}^{+}(S)=^{a}R\Gamma_{v}(S,\mathbb{B}_{I}^{+}).
  4. (4)

    For I[0,)I\subset[0,\infty) a compact interval we have a natural equivalence

    𝔹I(S)=RΓv(S,𝔹I)\mathbb{B}_{I}(S)=R\Gamma_{v}(S,\mathbb{B}_{I})
Proof.

By [Sch22, Proposition 8.8] we have a natural almost equivalence R+=aRΓv(S,𝒪+)R^{+}=^{a}R\Gamma_{v}(S,\mathscr{O}^{+}) and so an almost equivalence modulo any pseudo-uniformizer. Since 𝔸inf/(p,[ϖ])=a𝒪^+/ϖ\mathbb{A}_{\inf}/(p,[\varpi])=^{a}\widehat{\mathscr{O}}^{+}/\varpi as almost vv-sheaves, by derived Nakayama’s lemma we have an almost equivalence of derived (p,[ϖ])(p,[\varpi])-complexes

𝔸inf(R+)=aRΓv(S,𝔸inf)\mathbb{A}_{\inf}(R^{+})=^{a}R\Gamma_{v}(S,\mathbb{A}_{\inf})

proving (1).

Suppose that I=[0,r]I=[0,r], we have a short exact sequence of (p,[ϖ])(p,[\varpi])-adically complete vv-sheaves

0𝔸infT[ϖ]1/rTp𝔸infT𝔹I+00\to\mathbb{A}_{\inf}\langle T\rangle\xrightarrow{[\varpi]^{1/r}T-p}\mathbb{A}_{\inf}\langle T\rangle\to\mathbb{B}_{I}^{+}\to 0 (2.1)

where 𝔸infT\mathbb{A}_{\inf}\langle T\rangle is the (p,[ϖ])(p,[\varpi])-adic completion of the polynomial algebra over 𝔸inf\mathbb{A}_{\inf}. Indeed, this follows from the fact that one has the presentation for the ring 𝔹I+(S)\mathbb{B}_{I}^{+}(S) for any affinoid perfectoid S=Spa(R,R+)S=\operatorname{Spa}(R,R^{+}):

𝔹I+(S)={n[an]pn[ϖ]n/r:anR+ and |an|0 as n},\mathbb{B}_{I}^{+}(S)=\{\sum_{n\in\mathbb{N}}[a_{n}]\frac{p^{n}}{[\varpi]^{n/r}}\colon a_{n}\in R^{+}\mbox{ and }|a_{n}|\to 0\mbox{ as }n\to\infty\}, (2.2)

endowed with the [ϖ][\varpi]-adic topology, see [SW20, proof of Proposition 11.2.1]. The equation (2.2) shows that [ϖ]1/r[\varpi]^{1/r} is a regular element of 𝔹I+\mathbb{B}^{+}_{I}, and taking quotients in (2.1) by [ϖ]1/r[\varpi]^{1/r} yields an isomorphism of sheaves

𝔹[0,r]+/([ϖ1/r])=𝒪^+/ϖ1/r[T]\mathbb{B}^{+}_{[0,r]}/([\varpi^{1/r}])=\widehat{\mathscr{O}}^{+}/\varpi^{1/r}[T]

proving (2).

Then, part (2), the almost acyclicity of 𝒪^+\widehat{\mathscr{O}}^{+} and derived Nakayama’s lemma imply that

𝔹I+(S)=aRΓv(S,𝔹I+)\mathbb{B}_{I}^{+}(S)=^{a}R\Gamma_{v}(S,\mathbb{B}_{I}^{+})

proving (3).

Finally, part (4) for I=[0,r]I=[0,r] follows from (3) by inverting pseudo-uniformizers. Moreover, as 𝔹[0,r]+T\mathbb{B}^{+}_{[0,r]}\langle T\rangle is the ϖ\varpi-adic completion of 𝔹[0,r]+[T]\mathbb{B}^{+}_{[0,r]}[T], part (3) also implies that

𝔹[0,r]+(S)T=aRΓv(S,𝔹[0,r]+T)\mathbb{B}^{+}_{[0,r]}(S)\langle T\rangle=^{a}R\Gamma_{v}(S,\mathbb{B}_{[0,r]}^{+}\langle T\rangle)

and 𝔹[0,r](S)T=RΓv(S,𝔹[0,r]T)\mathbb{B}_{[0,r]}(S)\langle T\rangle=R\Gamma_{v}(S,\mathbb{B}_{[0,r]}\langle T\rangle). Let us now consider I=[s,r]I=[s,r] with s>0s>0. By [KHH+19, Lemma 1.8.2], for all affinoid perfectoid Spa(A,A+)Perfϖ\operatorname{Spa}(A,A^{+})\in\operatorname{Perf}_{\varpi} we have a short exact sequence

0𝔹[0,r](A,A+)TpT[ϖ]1/s𝔹[0,r](A,A+)T𝔹[r,s](A,A+)0.0\to\mathbb{B}_{[0,r]}(A,A^{+})\langle T\rangle\xrightarrow{pT-[\varpi]^{1/s}}\mathbb{B}_{[0,r]}(A,A^{+})\langle T\rangle\to\mathbb{B}_{[r,s]}(A,A^{+})\to 0.

This gives rise to a short exact sequence of vv-sheaves

0𝔹[0,r]TpT[ϖ]1/s𝔹[0,r]T𝔹[s,r]0.0\to\mathbb{B}_{[0,r]}\langle T\rangle\xrightarrow{pT-[\varpi]^{1/s}}\mathbb{B}_{[0,r]}\langle T\rangle\to\mathbb{B}_{[s,r]}\to 0. (2.3)

Part (4) follows from part (3) after taking vv-cohomology of (2.3). ∎

Lemma 2.1.5.

Let XX be a locally spatial diamond over an affinoid perfectoid space SS in characteristic pp. Let I[0,)I\subset[0,\infty) be a compact interval and ϖ\varpi a fixed pseudo-uniformizer of SS. Then 𝔹I+/[ϖ]\mathbb{B}^{+}_{I}/[\varpi] arises from an étale sheaf of XX via the fully faithful embedding Sh(Xe´t,Λ)Sh(Xv,Λ){\operatorname{Sh}}(X_{{\rm\acute{e}t}},\Lambda)\to{\operatorname{Sh}}(X_{v},\Lambda) of [Sch22, Proposition 14.10] with Λ=𝔹I+(S)/[ϖ]\Lambda=\mathbb{B}^{+}_{I}(S)/[\varpi].

Proof.

This follows from [Sch22, Theorem 14.12] since Λ=𝔹I+/[ϖ]\Lambda=\mathbb{B}^{+}_{I}/[\varpi] is clearly an étale sheaf on perfectoid spaces. Indeed, it suffices to show that 𝔹I+/[ϖ]b\mathbb{B}_{I}^{+}/[\varpi]^{b} is an étale sheaf for a suitable bb. If I=[0,r]I=[0,r] this follows from Lemma 2.1.4 (2). For I=[s,r]I=[s,r] with s>0s>0 consider the short exact sequence of vv-sheaves

0𝔹[0,r]TpT[ϖ]1/s𝔹[0,r]T𝔹[s,r]0.0\to\mathbb{B}_{[0,r]}\langle T\rangle\xrightarrow{pT-[\varpi]^{1/s}}\mathbb{B}_{[0,r]}\langle T\rangle\to\mathbb{B}_{[s,r]}\to 0.

Then, 𝔹[s,r]+/[ϖ]\mathbb{B}^{+}_{[s,r]}/[\varpi] is a subquotient of 𝔹[0,r]T/[ϖ]𝔹[0,r]+T\mathbb{B}_{[0,r]}\langle T\rangle/[\varpi]\mathbb{B}_{[0,r]}^{+}\langle T\rangle which is étale by the previous case, proving that 𝔹[s,r]+/[ϖ]\mathbb{B}^{+}_{[s,r]}/[\varpi] is étale itself. ∎

2.2. Solid almost quasi-coherent sheaves

In this paper we shall work with cohomologies of Banach sheaves on locally spatial diamonds such as 𝔹I\mathbb{B}_{I}. However, the sheaves we shall consider are not arbitrary; they are actually solid quasi-coherent sheaves over the 𝒴FF\mathcal{Y}^{FF}-curve in the sense of [AMLB]. This promotion to solid sheaves helps to naturally endow their vv-cohomologies with the structure of solid abelian groups as in [AM24, §4]. Since the sheaves we shall consider will be generic fibers of completed sheaves, it will be enough to use the formalism of solid almost quasi-coherent sheaves with torsion coefficients of [Man22b, §3] that we briefly recall in this section.

Let Perfϖ\operatorname{Perf}_{\varpi} be the category of perfectoids in characteristic pp with fixed pseudo-uniformizer ϖ\varpi. Let I=[0,r][0,)I=[0,r]\subset[0,\infty) be a compact interval and b>0b>0 a positive rational number. Consider the sheaf of coefficients 𝔹I,b+:=𝔹I+/([ϖ]b)\mathbb{B}^{+}_{I,b}:=\mathbb{B}^{+}_{I}/([\varpi]^{b}) on Perfϖ\operatorname{Perf}_{\varpi} with almost structure generated by ([ϖ]1/pk)k([\varpi]^{1/p^{k}})_{k\in\mathbb{N}}.

Definition 2.2.1.

Let XX be a small vv-stack with fixed pseudo-uniformizer ϖ\varpi. The \infty-category of solid almost quasi-coherent 𝔹I,b+\mathbb{B}_{I,b}^{+}-modules 𝒟a(X,𝔹I,b+)\mathscr{D}_{\square}^{a}(X,\mathbb{B}^{+}_{I,b}) is the hypercompletion of the functor mapping an affinoid perfectoid Spa(R,R+)Xv\operatorname{Spa}(R,R^{+})\in X_{v} to the almost category of solid 𝔹I,b+(R,R+)\mathbb{B}^{+}_{I,b}(R,R^{+})-modules 𝒟a(𝔹I,b+(R,R+))\mathscr{D}^{a}(\mathbb{B}^{+}_{I,b}(R,R^{+})_{\square}).

The category of almost solid modules satisfies strong descent properties:

Proposition 2.2.2.

Let X=Spa(R,R+)X=\operatorname{Spa}(R,R^{+}) be a totally disconnected perfectoid space. Then the natural map

𝒟a(𝔹I,b+(R,R+))𝒟a(X,𝔹I,b+)\mathscr{D}^{a}(\mathbb{B}^{+}_{I,b}(R,R^{+})_{\square})\to\mathscr{D}_{\square}^{a}(X,\mathbb{B}^{+}_{I,b})

is an equivalence of \infty-categories.

Proof.

This follows essentially from [Man22b, Theorem 3.1.27]. Indeed, let YXY_{\bullet}\to X be an hypercover of XX by totally disconnected perfectoid spaces, we want to show that the natural map

𝒟a(𝔹I,b+(R,R+))lim[n]Δ𝒟a(𝔹I,b+(Yn))\mathscr{D}^{a}(\mathbb{B}^{+}_{I,b}(R,R^{+})_{\square})\to\varprojlim_{[n]\in\Delta}\mathscr{D}^{a}(\mathbb{B}^{+}_{I,b}(Y_{n})_{\square})

is an equivalence. Concretely, this amounts to show the following:

  • i.

    For M𝒟a(𝔹I,b+(R,R+))M\in\mathscr{D}^{a}(\mathbb{B}^{+}_{I,b}(R,R^{+})_{\square}) the natural map

    Mlim[n]Δ(M𝔹I,b+(R,R+)L𝔹I,b+(Yn))M\to\varprojlim_{[n]\in\Delta}(M\otimes^{L}_{\mathbb{B}^{+}_{I,b}(R,R^{+})_{\square}}\mathbb{B}^{+}_{I,b}(Y_{n})_{\square})

    is an equivalence.

  • ii.

    For (Mn)[n]Δ(M_{n})_{[n]\in\Delta} a cocartesian section of lim[n]Δ𝒟a(𝔹I,b+(Yn))\varprojlim_{[n]\in\Delta}\mathscr{D}^{a}(\mathbb{B}^{+}_{I,b}(Y_{n})_{\square}) with totalization MM the natural map

    M𝔹I,b+(R,R+)L𝔹I,b+(Yn)MnM\otimes^{L}_{\mathbb{B}^{+}_{I,b}(R,R^{+})_{\square}}\mathbb{B}^{+}_{I,b}(Y_{n})_{\square}\to M_{n}

    is an equivalence for all nn\in\mathbb{N}.

Suppose first that b=1/rb=1/r, then by Lemma 2.1.4 (3) we have that 𝔹I,r+=𝒪+/ϖ1/r[T]\mathbb{B}^{+}_{I,r}=\mathscr{O}^{+}/\varpi^{1/r}[T] is a polynomial algebra over 𝒪+/ϖ1/r\mathscr{O}^{+}/\varpi^{1/r} where TT is the class of p/[ϖ]1/rp/[\varpi]^{1/r}. By [Man22b, Theorem 3.1.27] we have an equivalence of categories

𝒟a(R+/ϖ1/r)lim[n]Δ𝒟a(𝒪+(Yn)/ϖ1/r).\mathscr{D}^{a}(R^{+}_{\square}/\varpi^{1/r})\xrightarrow{\sim}\varprojlim_{[n]\in\Delta}\mathscr{D}^{a}(\mathscr{O}^{+}(Y_{n})_{\square}/\varpi^{1/r}). (2.4)

Consider the maps of augmented cosimplicial diagrams of analytic rings (𝒪+(Yn)/ϖ1/r)[n]Δ+(𝔹I,b+(Yn))[n]Δ+(\mathscr{O}^{+}(Y_{n})_{\square}/\varpi^{1/r})_{[n]\in\Delta_{+}}\to(\mathbb{B}^{+}_{I,b}(Y_{n})_{\square})_{[n]\in\Delta_{+}} with Y1=XY_{-1}=X. For any map α:[n][m]\alpha:[n]\to[m] consider the commutative square provide by base change

𝒟a((𝒪+(Yn)/ϖ1/r){\mathscr{D}^{a}((\mathscr{O}^{+}(Y_{n})_{\square}/\varpi^{1/r})}𝒟a((𝒪+(Ym)/ϖ1/r){\mathscr{D}^{a}((\mathscr{O}^{+}(Y_{m})_{\square}/\varpi^{1/r})}𝒟a(𝔹I,b+(Yn)){\mathscr{D}^{a}(\mathbb{B}^{+}_{I,b}(Y_{n})_{\square})}𝒟a(𝔹I,b+(Ym)).{\mathscr{D}^{a}(\mathbb{B}^{+}_{I,b}(Y_{m})_{\square}).}fα\scriptstyle{f^{*}_{\alpha}}hn\scriptstyle{h_{n}^{*}}hm\scriptstyle{h_{m}^{*}}gα\scriptstyle{g^{*}_{\alpha}}

Let hn,h_{n,*} be the right adjoint of hnh^{*}_{n} given by the forgetful functor. Since any map of discrete Huber pairs is steady [Man22b, Proposition 2.9.7 (ii)], the natural transformations fαhn,hm,gαf_{\alpha}^{*}h_{n,*}\xrightarrow{\sim}h_{m,*}g_{\alpha}^{*} of functors 𝒟a(𝔹I,b+(Yn))𝒟a((𝒪+(Ym)/ϖ1/r)\mathscr{D}^{a}(\mathbb{B}^{+}_{I,b}(Y_{n})_{\square})\to\mathscr{D}^{a}((\mathscr{O}^{+}(Y_{m})_{\square}/\varpi^{1/r}) is an equivalence. Therefore, the forgetful functors hn,h_{n,*} preserve cocartesian sections and induce a functor

(hn,)[n]:lim[n]Δ𝒟a(𝔹I,b+(Yn))lim[n]Δ𝒟a(𝒪+(Yn)/ϖ1/r)𝒟a(R+/ϖ1/r)(h_{n,*})_{[n]}:\varprojlim_{[n]\in\Delta}\mathscr{D}^{a}(\mathbb{B}^{+}_{I,b}(Y_{n})_{\square})\to\varprojlim_{[n]\in\Delta}\mathscr{D}^{a}(\mathscr{O}^{+}(Y_{n})_{\square}/\varpi^{1/r})\cong\mathscr{D}^{a}(R^{+}/\varpi^{1/r})

which is the right adjoint of the natural base change along (𝒪+(Yn)/ϖ1/r)[n]Δ(𝔹I,b+(Yn))[n]Δ(\mathscr{O}^{+}(Y_{n})_{\square}/\varpi^{1/r})_{[n]\in\Delta}\to(\mathbb{B}^{+}_{I,b}(Y_{n})_{\square})_{[n]\in\Delta}, and that fits in a commutative square

𝒟a(𝔹I,b+(R,R+)){\mathscr{D}^{a}(\mathbb{B}^{+}_{I,b}(R,R^{+})_{\square})}lim[n]Δ𝒟a(𝔹I,b+(Yn)){\varprojlim_{[n]\in\Delta}\mathscr{D}^{a}(\mathbb{B}^{+}_{I,b}(Y_{n})_{\square})}𝒟a(R+/ϖ1/r){\mathscr{D}^{a}(R^{+}/\varpi^{1/r})}lim[n]Δ𝒟a(𝒪+(Yn)/ϖ1/r).{\varprojlim_{[n]\in\Delta}\mathscr{D}^{a}(\mathscr{O}^{+}(Y_{n})_{\square}/\varpi^{1/r}).}h\scriptstyle{h_{*}}(hn,)[n]\scriptstyle{(h_{n,*})_{[n]}}\scriptstyle{\sim}

Therefore, since the functors hn,h_{n,*} are conservative, in order to show (i) or (ii) we can apply (hn,)[n](h_{n,*})_{[n]} where the claim follows from (2.4).

The case of general bb follows from derived Nakayama’s lemma: for either (i) or (ii) above we have to show that a map of solid 𝔹I,b+(R,R+)\mathbb{B}^{+}_{I,b}(R,R^{+})-modules NNN\to N^{\prime} is an equivalence. For this, it suffices to check that it is an equivalence after taking derived quotients by [ϖ]1/r[\varpi]^{1/r} where it was already proven. ∎

Finally, we recall how smooth representation theory appears in terms of solid almost quasi-coherent sheaves.

Proposition 2.2.3.

Let X=Spa(R,R+)X=\operatorname{Spa}(R,R^{+}) be a totally disconnected perfectoid space with pseudo-uniformizer ϖ\varpi. Let Π\Pi be a locally profinite group acting on XX and consider the vv-stack X/ΠX/\Pi. Then the pullback along the map f:XX/Πf:X\to X/\Pi realizes 𝒟a(X/Π,𝔹I,r+)\mathscr{D}^{a}_{\square}(X/\Pi,\mathbb{B}^{+}_{I,r}) as the derived \infty-category of semilinear smooth almost representations Rep𝔹I,b+(R,R+)sm,a(Π)\operatorname{Rep}^{sm,a}_{\mathbb{B}^{+}_{I,b}(R,R^{+})_{\square}}(\Pi) of Π\Pi (denoted as 𝒟sm,a(𝔹I,b+(R,R+),Π)\mathscr{D}^{sm,a}(\mathbb{B}^{+}_{I,b}(R,R^{+})_{\square},\Pi) in [Man22b, Definition 3.4.11]).

Proof.

This follows from the same argument of [Man22b, Lemma 3.4.26]. ∎

Remark 2.2.4.

By construction the \infty-category Rep𝔹I,b+(R,R+)sm,a(Π)\operatorname{Rep}^{sm,a}_{\mathbb{B}^{+}_{I,b}(R,R^{+})_{\square}}(\Pi) is the derived category of its heart
Rep𝔹I,b+(R,R+)sm,a,(Π)\operatorname{Rep}^{sm,a,\heartsuit}_{\mathbb{B}^{+}_{I,b}(R,R^{+})_{\square}}(\Pi). There is an obvious forgetful functor

Rep𝔹I,b+(R,R+)sm,a,(Π)Moda(𝔹I,b+(R,R+)[Π])\operatorname{Rep}^{sm,a,\heartsuit}_{\mathbb{B}^{+}_{I,b}(R,R^{+})_{\square}}(\Pi)\to\operatorname{Mod}^{a}(\mathbb{B}^{+}_{I,b}(R,R^{+})_{\square}[\Pi])

to the category of almost solid modules over the semilinear solid group algebra 𝔹I,b+(R,R+)[Π]\mathbb{B}^{+}_{I,b}(R,R^{+})_{\square}[\Pi]. This gives rise a map of derived \infty-categories

Rep𝔹I,b+(R,R+)sm,a(Π)𝒟a(𝔹I,b+(R,R+)[Π]).\operatorname{Rep}^{sm,a}_{\mathbb{B}^{+}_{I,b}(R,R^{+})_{\square}}(\Pi)\to\mathscr{D}^{a}(\mathbb{B}^{+}_{I,b}(R,R^{+})_{\square}[\Pi]).

This map is not in general fully faithful, see Remark [Man22b, 3.4.18] for a counter example when Π=𝔽p\Pi=\prod_{\mathbb{N}}\mathbb{F}_{p}. However, the fully faithfulness is expected when Π\Pi has a basis of compact open subgroups with uniformly bounded finite pp-cohomological dimension, eg. when Π\Pi is a pp-adic Lie group.

2.3. The décalage functor

In the next section we recall some facts about the décalage functor LηL\eta_{\mathcal{I}} of [BMS18]. For us it will suffice to consider this functor at the level of the homotopy category of the \infty-derived category of modules of an algebra in a topos as in loc. cit.

Let (T,𝒪T)(T,\mathscr{O}_{T}) be a ringed topos. Let K(𝒪T)K(\mathscr{O}_{T}) be the category of complexes of 𝒪T\mathscr{O}_{T}-modules up to homotopy, and D(𝒪T)D(\mathscr{O}_{T}) the derived category of 𝒪T\mathscr{O}_{T}-modules obtained by inverting quasi-isomorphisms in K(𝒪T)K(\mathscr{O}_{T}).

Let 𝒪T\mathcal{I}\subset\mathscr{O}_{T} be an invertible ideal, and let Kfree(𝒪T)K^{\mathcal{I}-\operatorname{\scriptsize free}}(\mathscr{O}_{T}) denote the full subcategory of K(𝒪T)K(\mathscr{O}_{T}) whose objects are \mathcal{I}-torsion free complexes. By [BMS18, Lemma 6.1] the (non \infty!) derived category D(𝒪T)D(\mathscr{O}_{T}) is the localization of Kfree(𝒪T)K^{\mathcal{I}-\operatorname{\scriptsize free}}(\mathscr{O}_{T}) along quasi-isomorphisms.

Definition 2.3.1 ([BMS18, Definition 6.2]).

Let CKfree(𝒪T)C^{\bullet}\in K^{\mathcal{I}-\operatorname{\scriptsize free}}(\mathscr{O}_{T}). Define a new object ηC=(ηC)Kfree(𝒪T)\eta_{\mathcal{I}}C^{\bullet}=(\eta_{\mathcal{I}}C)^{\bullet}\in K^{\mathcal{I}-\operatorname{\scriptsize free}}(\mathscr{O}_{T}) with terms

(ηC)i={xCi|dxCi+1}𝒪Ti,(\eta_{\mathcal{I}}C)^{i}=\{x\in C^{i}|dx\in\mathcal{I}C^{i+1}\}\otimes_{\mathscr{O}_{T}}\mathcal{I}^{\otimes i},

and differentials

dηC,i:(ηC)i(ηIC)i+1d_{\eta_{\mathcal{I}}C,i}:(\eta_{\mathcal{I}}C)^{i}\to(\eta_{I}C)^{i+1}

making the following diagram commute

(ηC)i{(\eta_{\mathcal{I}}C)^{i}}Ci+1i{\mathcal{I}C^{i+1}\otimes\mathcal{I}^{\otimes i}}(ηC)i+1{(\eta_{\mathcal{I}}C)^{i+1}}Ci+1i+1.{C^{i+1}\otimes\mathcal{I}^{\otimes i+1}.}dCii\scriptstyle{d_{C^{i}}\otimes\mathcal{I}^{\otimes i}}d(ηC)i\scriptstyle{d_{(\eta_{\mathcal{I}}C)^{i}}}\scriptstyle{\simeq}

By [BMS18, Corollary 6.5] the operator LηL\eta_{\mathcal{I}} preserves quasi-isomorphisms and extends to a filtered colimit preserving functor

Lη:D(𝒪T)D(𝒪T).L\eta_{\mathcal{I}}:D(\mathscr{O}_{T})\to D(\mathscr{O}_{T}).

Moreover, the following properties hold:

  • For CD(𝒪T)C\in D(\mathscr{O}_{T}) there are natural isomorphisms [BMS18, Lemma 6.4]

    Hi(LηC)Hi(C)/Hi(C)[]𝒪Ti.H^{i}(L\eta_{\mathcal{I}}C)\cong H^{i}(C)/H^{i}(C)[\mathcal{I}]\otimes_{\mathscr{O}_{T}}\mathcal{I}^{i}.
  • LηL\eta_{\mathcal{I}} is lax symmetric monoidal, i.e. for C,DD(𝒪T)C,D\in D(\mathscr{O}_{T}) there is a natural map

    LηC𝒪TLLηDLη(C𝒪TLD)L\eta_{\mathcal{I}}C\otimes^{L}_{\mathscr{O}_{T}}L\eta_{\mathcal{I}}D\to L\eta_{\mathcal{I}}(C\otimes^{L}_{\mathscr{O}_{T}}D)

    functorial in CC and DD, and symmetric in CC and DD [BMS18, Lemma 6.7].

  • Suppose that the topos is replete. Let CD(𝒪T)C\in D(\mathscr{O}_{T}), then the natural maps

    (LηIC)LηI(C)Rlimn(Lη(C𝒪TL𝒪T/n))(L\eta_{I}C)^{\wedge\mathcal{I}}\to L\eta_{I}(C^{\wedge\mathcal{I}})\to R\varprojlim_{n}(L\eta_{\mathcal{I}}(C\otimes^{L}_{\mathscr{O}_{T}}\mathscr{O}_{T}/\mathcal{I}^{n}))

    are equivalences [BMS18, Lemma 6.20]. Here for an object MD(𝒪T)M\in D(\mathscr{O}_{T}) we let M=Rlimn(M𝒪TL𝒪T/n)M^{\wedge\mathcal{I}}=R\varprojlim_{n}(M\otimes^{L}_{\mathscr{O}_{T}}\mathscr{O}_{T}/\mathcal{I}^{n}) be the derived \mathcal{I}-adic completion.

Remark 2.3.2.

The functor LηL\eta_{\mathcal{I}} preserves filtered colimits but is not exact, i.e. it does not preserves cones. For example, we have the short exact sequence

0/2𝒪T/2𝒪T/00\to\mathcal{I}/\mathcal{I}^{2}\to\mathscr{O}_{T}/\mathcal{I}^{2}\to\mathscr{O}_{T}/\mathcal{I}\to 0

but we also have that

Lη(/2)=Lη(𝒪T/)=0L\eta_{\mathcal{I}}(\mathcal{I}/\mathcal{I}^{2})=L\eta_{\mathcal{I}}(\mathscr{O}_{T}/\mathcal{I})=0

and

Lη(𝒪T/2)=𝒪T/.L\eta_{\mathcal{I}}(\mathscr{O}_{T}/\mathcal{I}^{2})=\mathscr{O}_{T}/\mathcal{I}.

We shall need the following behavior of the décalage operator with respect to the passage to the special fiber.

Lemma 2.3.3.

Let CD(𝒪T)C\in D(\mathscr{O}_{T}) and k1k\geq 1, then the natural map

LηC𝒪TL𝒪T/k=LηC𝒪TLLη(𝒪T/k+1)Lη(C𝒪TL𝒪T/k+1)L\eta_{\mathcal{I}}C\otimes^{L}_{\mathscr{O}_{T}}\mathscr{O}_{T}/\mathcal{I}^{k}=L\eta_{\mathcal{I}}C\otimes^{L}_{\mathscr{O}_{T}}L\eta_{\mathcal{I}}(\mathscr{O}_{T}/\mathcal{I}^{k+1})\xrightarrow{\sim}L\eta_{\mathcal{I}}(C\otimes^{L}_{\mathscr{O}_{T}}\mathscr{O}_{T}/\mathcal{I}^{k+1}) (2.5)

is an equivalence.

Proof.

Let CC^{\bullet} be an \mathcal{I}-torsion free complex representing CC. Then, C𝒪TL𝒪T/k+1C\otimes^{L}_{\mathscr{O}_{T}}\mathscr{O}_{T}/\mathcal{I}^{k+1} is represented by the complex DD^{\bullet} with terms

Di=CiCi+1𝒪Tk+1D^{i}=C^{i}\oplus C^{i+1}\otimes_{\mathscr{O}_{T}}\mathcal{I}^{k+1}

and differentials

dDi=(dCiidCi+1ι0(1)idCi+1k+1)d_{D^{i}}=\left(\begin{array}[]{cc}d_{C^{i}}&\operatorname{id}_{C^{i+1}}\otimes\iota\\ 0&(-1)^{i}d_{C^{i+1}}\otimes\mathcal{I}^{k+1}\end{array}\right)

where ι:k+1𝒪T\iota:\mathcal{I}^{k+1}\to\mathscr{O}_{T}. Therefore, Lη(C𝒪TL𝒪T/k+1)L\eta_{\mathcal{I}}(C\otimes^{L}_{\mathscr{O}_{T}}\mathscr{O}_{T}/\mathcal{I}^{k+1}) is represented by the complex D~\widetilde{D}^{\bullet} with

D~i={(a,b)Di|dDi(a,b)Di+1}𝒪Ti.\widetilde{D}^{i}=\{(a,b)\in D^{i}|d_{D^{i}}(a,b)\in\mathcal{I}D^{i+1}\}\otimes_{\mathscr{O}_{T}}\mathcal{I}^{i}.

More explicitly, we have

dDi(a,b)=(dCi(a)+b,(1)ndCi+1(b)).d_{D^{i}}(a,b)=(d_{C^{i}}(a)+b,(-1)^{n}d_{C^{i+1}}(b)).

Since bCi+1k+1Ci+1b\in C^{i+1}\otimes\mathcal{I}^{k+1}\subset\mathcal{I}C^{i+1}, one deduces that

D~i\displaystyle\widetilde{D}^{i} ={(a,b)CiCi+1𝒪Tk+1|dCi(a)Ci+1 and dCi+1(b)Ci+1𝒪Tk+1}i\displaystyle=\{(a,b)\in C^{i}\oplus C^{i+1}\otimes_{\mathscr{O}_{T}}\mathcal{I}^{k+1}|d_{C^{i}}(a)\in\mathcal{I}C^{i+1}\mbox{ and }d_{C^{i+1}}(b)\in\mathcal{I}C^{i+1}\otimes_{\mathscr{O}_{T}}\mathcal{I}^{k+1}\}\otimes\mathcal{I}^{i}
={aCi|dCi(a)Ci+1}i{bCi+1𝒪Tk|dCi+1(b)Ci+1𝒪Tk}𝒪Ti+1\displaystyle=\{a\in C^{i}|d_{C^{i}}(a)\in\mathcal{I}C^{i+1}\}\otimes\mathcal{I}^{i}\oplus\{b\in C^{i+1}\otimes_{\mathscr{O}_{T}}\mathcal{I}^{k}|d_{C^{i+1}}(b)\in\mathcal{I}C^{i+1}\otimes_{\mathscr{O}_{T}}\mathcal{I}^{k}\}\otimes_{\mathscr{O}_{T}}\mathcal{I}^{i+1}
=(ηC)i(ηC)i+1𝒪Tk.\displaystyle=(\eta_{\mathcal{I}}C^{\bullet})^{i}\oplus(\eta_{\mathcal{I}}C^{\bullet})^{i+1}\otimes_{\mathscr{O}_{T}}\mathcal{I}^{k}.

It is straightforward to check that the differentials of D~\widetilde{D}^{\bullet} are those arising from the cone of ηC𝒪TkηC\eta_{\mathcal{I}}C^{\bullet}\otimes_{\mathscr{O}_{T}}\mathcal{I}^{k}\to\eta_{\mathcal{I}}C^{\bullet}, and that the resulting quasi-isomorphism

LηC𝒪TL𝒪T/kLη(C𝒪TL𝒪T/k+1)L\eta_{\mathcal{I}}C\otimes^{L}_{\mathscr{O}_{T}}\mathscr{O}_{T}/\mathcal{I}^{k}\cong L\eta_{\mathcal{I}}(C\otimes^{L}_{\mathscr{O}_{T}}\mathscr{O}_{T}/\mathcal{I}^{k+1})

is the one induced by the map (2.5). ∎

2.4. Solid locally analytic representations

Throughout this paper we will use the theory of solid locally analytic representations of [RJRC22, RJRC23]. In this section we briefly recall some of the main definitions and properties that will be needed later.

Let Solid\mathrm{Solid} be the abelian category of solid abelian groups and let \otimes_{\square} be its solid tensor product. for a ring RSolidR\in\mathrm{Solid} we let SolidR\mathrm{Solid}_{R} be the abelian category of solid RR-modules. We write 𝒟(R)\mathscr{D}(R) for the derived \infty-category of SolidR\mathrm{Solid}_{R}. Let GG be a compact pp-adic Lie group and let p,[G]=limHp[G/H]\mathbb{Z}_{p,{\square}}[G]=\varprojlim_{H}\mathbb{Z}_{p}[G/H] be the free solid p\mathbb{Z}_{p}-algebra generated by GG; it coincides with the Iwasawa algebra of GG with coefficients in p\mathbb{Z}_{p}. We set p,[G]=p,[G][1p]\mathbb{Q}_{p,{\square}}[G]=\mathbb{Z}_{p,{\square}}[G][\frac{1}{p}]. The group GG has a space Cla(G,p)C^{la}(G,\mathbb{Q}_{p}) of locally analytic functions, it can be written as the filtered colimit

Cla(G,p)=limhCh(G,p)C^{la}(G,\mathbb{Q}_{p})=\varinjlim_{h\to\infty}C^{h}(G,\mathbb{Q}_{p})

where Ch(G,p)=𝒪(𝔾(h))C^{h}(G,\mathbb{Q}_{p})=\mathscr{O}(\mathbb{G}^{(h)}) is the affinoid algebra of a decreasing sequence of affinoid groups over p\mathbb{Q}_{p}

G𝔾(h+1)𝔾(h)G\subset\cdots\subset\mathbb{G}^{(h+1)}\subset\mathbb{G}^{(h)}\subset\cdots

with limh𝔾(h)=G\varprojlim_{h}\mathbb{G}^{(h)}=G.

Given C𝒟(p,[G])C\in\mathscr{D}(\mathbb{Q}_{p,\square}[G]) a derived solid GG-representation, its (derived) locally analytic vectors [RJRC23, Definition 3.1.4] is the solid GG-representation

CRGla:=RΓ(G,Cp,LCla(G,p)1)C^{RG-la}:=R\Gamma(G,C\otimes^{L}_{\mathbb{Q}_{p},\square}C^{la}(G,\mathbb{Q}_{p})_{\star_{1}})

where

  • Cla(G,p)1C^{la}(G,\mathbb{Q}_{p})_{\star_{1}} is endowed with the left regular GG-action.

  • The tensor Cp,LCla(G,p)1C\otimes^{L}_{\mathbb{Q}_{p},\square}C^{la}(G,\mathbb{Q}_{p})_{\star_{1}} is endowed with the diagonal GG-action.

  • The GG-action on CRGlaC^{RG-la} arises from the right regular action on Cla(G,p)C^{la}(G,\mathbb{Q}_{p}).

We say that CC is (derived) locally analytic if the natural map

CRGlaCC^{RG-la}\to C

is an equivalence. We let Repp,la(G)𝒟(p,[G])\operatorname{Rep}^{la}_{\mathbb{Q}_{p,\square}}(G)\subset\mathscr{D}(\mathbb{Q}_{p,\square}[G]) be the full subcategory of locally analytic representations. This category satisfies the following properties:

  • Repp,la(G)\operatorname{Rep}^{la}_{\mathbb{Q}_{p,\square}}(G) is stable under colimits and p,\mathbb{Q}_{p,\square}-linear tensor products in 𝒟(p,[G])\mathscr{D}(\mathbb{Q}_{p,\square}[G]) [RJRC23, Proposition 3.2.3].

  • The tt-structure of 𝒟(p,[G])\mathscr{D}(\mathbb{Q}_{p,\square}[G]) induces a tt-structure on Repp,la(G)\operatorname{Rep}^{la}_{\mathbb{Q}_{p,\square}}(G) [RJRC23, Proposition 3.2.5]. Moreover, Repp,la(G)\operatorname{Rep}^{la}_{\mathbb{Q}_{p,\square}}(G) is the derived category of its heart [RJRC23, Proposition 3.2.6].

  • The functor CCRGlaC\mapsto C^{RG-la} is the right adjoint of the inclusion Repp,la(G)𝒟(p,[G])\operatorname{Rep}^{la}_{\mathbb{Q}_{p,\square}}(G)\subset\mathscr{D}(\mathbb{Q}_{p,\square}[G]) [RJRC23, Corollary 3.2.7].

The key lemma that we will use in this paper is the following criterion of locally analyticity:

Lemma 2.4.1.

Let C𝒟0(p,[G])C\in\mathscr{D}_{\geq 0}(\mathbb{Z}_{p,\square}[G]) be a connective derived pp-adically complete solid representation of GG. Suppose that there is an open compact subgroup G0GG_{0}\subset G, and a finite extension 𝒪K/p\mathcal{O}_{K}/\mathbb{Z}_{p} with pseudo-uniformizer π\pi, such that for all gG0g\in G_{0} the map

1g:Vp,L𝒪K/πVp,L𝒪K/π1-g\colon V\otimes^{L}_{\mathbb{Z}_{p},\square}\mathcal{O}_{K}/\pi\to V\otimes^{L}_{\mathbb{Z}_{p},\square}\mathcal{O}_{K}/\pi

is homotopic to zero as 𝒪K/π\mathcal{O}_{K}/\pi-module. Then V[1p]V[\frac{1}{p}] is a locally analytic representation of GG.

Proof.

We can assume without loss of generality that GG is an uniform pro-pp-group. By [RJRC23, Proposition 3.3.2] to show that V[1p]V[\frac{1}{p}] is GG-locally analytic, it suffices to show that for all gGg\in G it is Γg=gp\Gamma_{g}=g^{\mathbb{Z}_{p}}-locally analytic. Then, we can assume that GpG\cong\mathbb{Z}_{p}. The lemma follows from the same argument of [RJRC23, Proposition 3.3.3] applied to Vp𝒪KV\otimes_{\mathbb{Z}_{p}}\mathcal{O}_{K} and π\pi instead of VV and pp respectively. ∎

With this criteria one can show that actions of pp-adic Lie groups on rigid spaces are always locally analytic:

Corollary 2.4.2.

Let KK be a complete non-archimedean field of characteristic zero. Let AA be a Tate algebra of finite type over KK and GG a compact pp-adic Lie group acting continuously on AA. Then AA is a locally analytic representation of GG.

Proof.

Let A0AA_{0}\subset A be a ring of definition of AA, we can suppose without loss of generality that A0A_{0} is stable under the action of GG and that A0A_{0} is topologically generated over 𝒪K\mathcal{O}_{K} by finitely many variables T1,,TnT_{1},\ldots,T_{n}. Thus, since the action of GG on A0/pA_{0}/p is smooth, there is some open subgroup G0GG_{0}\subset G leaving the variables TiT_{i} fixed. But then G0G_{0} acts trivially on A0/pA_{0}/p and by Lemma 2.4.1 AA is a locally analytic representation of GG. ∎

2.5. Equivariant sheaves over flag varieties

Let 𝐆\mathbf{G} be a reductive group over p\mathbb{Q}_{p} and let μ\mu be a conjugacy class of cocharacters of GG with field of definition EE. We denote by FL𝐆,μ,E\operatorname{FL}_{\mathbf{G},\mu,E} the flag variety over EE parametrizing decreasing μ\mu-filtrations on 𝐆\mathbf{G}-representations seen as an algebraic variety, we let 𝐆,μ,E\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu,E} denote its analytification as an adic space over Spa(E,𝒪E)\operatorname{Spa}(E,\mathcal{O}_{E}) as in [Hub96]. Note that FL𝐆,μ1,E\operatorname{FL}_{\mathbf{G},\mu^{-1},E} is also the flag variety parametrizing increasing μ\mu-filtrations.

Let C/EC/E be a complete algebraically closed non-archimedean field an let us write FL𝐆,μ\operatorname{FL}_{\mathbf{G},\mu} and 𝐆,μ\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu} for the base change of the flag varieties to CC. We fix a cocharacter μ:𝔾m𝐆C\mu:\mathbb{G}_{m}\to\mathbf{G}_{C} so that FL𝐆,μ𝐆C/𝐏μ\operatorname{FL}_{\mathbf{G},\mu}\cong\mathbf{G}_{C}/\mathbf{P}_{\mu} where 𝐏μ𝐆C\mathbf{P}_{\mu}\subset\mathbf{G}_{C} is the parabolic subgroup parametrizing decreasing μ\mu-filtrations. We let 𝐍μ𝐏μ\mathbf{N}_{\mu}\subset\mathbf{P}_{\mu} be its unipotent radical and let 𝐌μ\mathbf{M}_{\mu} be the Levi subgroup, i.e. the centralizer of μ\mu in 𝐆C\mathbf{G}_{C}. We have a semi-direct product decomposition 𝐏μ=𝐍μ𝐌μ\mathbf{P}_{\mu}=\mathbf{N}_{\mu}\rtimes\mathbf{M}_{\mu}.

Set =SpecC*=\operatorname{Spec}C. We have an isomorphism of Artin stacks

[1]:/𝐏μ𝐆C\(𝐆C/𝐏μ)=𝐆C\FL𝐆,μ.[1]:*/\mathbf{P}_{\mu}\xrightarrow{\sim}\mathbf{G}_{C}\backslash(\mathbf{G}_{C}/\mathbf{P}_{\mu})=\mathbf{G}_{C}\backslash\operatorname{FL}_{\mathbf{G},\mu}.

Therefore, pullback along [1][1] gives rise an equivalence of quasi-coherent sheaves on the stacks. The previous translates in the classical equivalence of representation categories:

[1]:𝐆QCoh(FL𝐆,μ)RepCalg𝐏μ[1]^{*}:\mathbf{G}-\operatorname{QCoh}(\operatorname{FL}_{\mathbf{G},\mu})\xrightarrow{\sim}\operatorname{Rep}^{\operatorname{\scriptsize alg}}_{C}\mathbf{P}_{\mu} (2.6)

from 𝐆\mathbf{G}-equivariant quasi-coherent sheaves on FL𝐆,μ\operatorname{FL}_{\mathbf{G},\mu} and algebraic CC-linear representations of 𝐏μ\mathbf{P}_{\mu}. We write 𝒲𝐆,μ\mathcal{W}_{\mathbf{G},\mu} for the inverse of (2.6).

Next, we introduce some notation appearing in the localization theory of Beilinson-Bernstein [BB81]. Let 𝔤=Lie𝐆\mathfrak{g}=\operatorname{Lie}\mathbf{G} be the Lie algebra of 𝐆\mathbf{G} over p\mathbb{Q}_{p} and let 𝔤C\mathfrak{g}_{C} be its base change to CC. Let 𝔭μ,𝔫μ\mathfrak{p}_{\mu},\mathfrak{n}_{\mu} and 𝔪μ\mathfrak{m}_{\mu} be the Lie algebras of 𝐏μ\mathbf{P}_{\mu}, 𝐍μ\mathbf{N}_{\mu} and 𝐌μ\mathbf{M}_{\mu} respectively. We let 𝔤μ0=𝒪FL𝐆,μp𝔤\mathfrak{g}^{0}_{\mu}=\mathscr{O}_{\operatorname{FL}_{\mathbf{G},\mu}}\otimes_{\mathbb{Q}_{p}}\mathfrak{g}, and let 𝔭μ0,𝔫μ0\mathfrak{p}_{\mu}^{0},\mathfrak{n}^{0}_{\mu} and 𝔪μ0\mathfrak{m}^{0}_{\mu} be the 𝐆\mathbf{G}-equivariant sheaves over FL𝐆,μ\operatorname{FL}_{\mathbf{G},\mu} corresponding to the adjoint action of 𝐏μ\mathbf{P}_{\mu} via (2.6). Note that we have inclusions of 𝐆\mathbf{G}-equivariant sheaves 𝔫μ0𝔭μ0𝔤μ0\mathfrak{n}^{0}_{\mu}\subset\mathfrak{p}^{0}_{\mu}\subset\mathfrak{g}^{0}_{\mu} and an isomorphism 𝔭μ0/𝔫μ0=𝔪μ0\mathfrak{p}^{0}_{\mu}/\mathfrak{n}^{0}_{\mu}=\mathfrak{m}^{0}_{\mu}. The action of 𝐆\mathbf{G} on FL𝐆,μ\operatorname{FL}_{\mathbf{G},\mu} can be differentiated to a 𝐆\mathbf{G}-equivariant 𝒪FL𝐆,μ\mathscr{O}_{\operatorname{FL}_{\mathbf{G},\mu}}-linear map

α:𝔤μ0𝒯FL𝐆,μ\alpha:\mathfrak{g}^{0}_{\mu}\to\mathcal{T}_{\operatorname{FL}_{\mathbf{G},\mu}} (2.7)

where 𝒯FL𝐆,μ\mathcal{T}_{\operatorname{FL}_{\mathbf{G},\mu}} is the tangent space of FL𝐆,μ\operatorname{FL}_{\mathbf{G},\mu}. The map (2.7) induces an isomorphism

α¯:𝔤μ0/𝔭μ0𝒯FL𝐆,μ.\overline{\alpha}:\mathfrak{g}^{0}_{\mu}/\mathfrak{p}^{0}_{\mu}\xrightarrow{\sim}\mathcal{T}_{\operatorname{FL}_{\mathbf{G},\mu}}. (2.8)

Similarly, let π𝐌μ:FL𝐆,μ+FL𝐆,μ\pi_{\mathbf{M}_{\mu}}:\operatorname{FL}^{+}_{\mathbf{G},\mu}\to\operatorname{FL}_{\mathbf{G},\mu} be the natural 𝐌μ\mathbf{M}_{\mu}-torsor over FL𝐆,μ\operatorname{FL}_{\mathbf{G},\mu} given by FL𝐆,μ+=𝐆E/𝐍μ\operatorname{FL}^{+}_{\mathbf{G},\mu}=\mathbf{G}_{E}/\mathbf{N}_{\mu}. The action of 𝐆\mathbf{G} induces a 𝐆×𝐌μ\mathbf{G}\times\mathbf{M}_{\mu}-equivariant 𝒪FL𝐆,μ+\mathscr{O}_{\operatorname{FL}^{+}_{\mathbf{G},\mu}}-linear map

α+:𝒪FL𝐆,μ+p𝔤𝒯FL𝐆,μ+\alpha^{+}:\mathscr{O}_{\operatorname{FL}^{+}_{\mathbf{G},\mu}}\otimes_{\mathbb{Q}_{p}}\mathfrak{g}\to\mathcal{T}_{\operatorname{FL}^{+}_{\mathbf{G},\mu}}

with 𝒯FL𝐆,μ+\mathcal{T}_{\operatorname{FL}^{+}_{\mathbf{G},\mu}} the tangent space of FL𝐆,μ+\operatorname{FL}^{+}_{\mathbf{G},\mu}. Taking pushforward along π𝐌μ\pi_{\mathbf{M}_{\mu}} and 𝐌μ\mathbf{M}_{\mu}-invariants, we get a 𝐆\mathbf{G}-equivariant map

α+:𝔤μ0(π𝐌μ,(𝒯FL𝐆,μ+))𝐌μ\alpha^{+}:\mathfrak{g}^{0}_{\mu}\to(\pi_{\mathbf{M}_{\mu},*}(\mathcal{T}_{\operatorname{FL}^{+}_{\mathbf{G,\mu}}}))^{\mathbf{M}_{\mu}}

that induces an isomorphism

α¯+:𝔤μ0/𝔫μ0(π𝐌μ,(𝒯FL𝐆,μ+))𝐌μ.\overline{\alpha}^{+}:\mathfrak{g}^{0}_{\mu}/\mathfrak{n}^{0}_{\mu}\xrightarrow{\sim}(\pi_{\mathbf{M}_{\mu},*}(\mathcal{T}_{\operatorname{FL}^{+}_{\mathbf{G,\mu}}}))^{\mathbf{M}_{\mu}}.
Remark 2.5.1.
  1. (1)

    In order to construct the equivalence (2.6) it suffices to consider a base change to F/EF/E such that the conjugacy class μ\mu admits a representative. Then the groups 𝐏μ,𝐍μ\mathbf{P}_{\mu},\mathbf{N}_{\mu} and 𝐌μ\mathbf{M}_{\mu} are defined over FF.

  2. (2)

    The Lie algebroids 𝔤μ0\mathfrak{g}^{0}_{\mu}, 𝔭μ\mathfrak{p}_{\mu}, 𝔫μ\mathfrak{n}_{\mu} and 𝔪μ0\mathfrak{m}^{0}_{\mu} as well as the anchor map (2.7) admit natural descent to EE. Indeed, the descent of the Lie algebroid 𝔤μ0\mathfrak{g}^{0}_{\mu} is nothing but 𝔤μ,E0=𝒪FL𝐆,μ,Ep𝔤\mathfrak{g}^{0}_{\mu,E}=\mathscr{O}_{\operatorname{FL}_{\mathbf{G},\mu,E}}\otimes_{\mathbb{Q}_{p}}\mathfrak{g}. One has a sub Lie algebroid 𝔤Eder,0𝔤μ,E0\mathfrak{g}^{\mathrm{der},0}_{E}\subset\mathfrak{g}^{0}_{\mu,E} induced by the derived Lie algebra 𝔤der𝔤\mathfrak{g}^{\mathrm{der}}\subset\mathfrak{g}. Since 𝐆\mathbf{G} acts on 𝐆,μ,E\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu,E} one has an anchor map by taking derivations

    𝔤Eder,0𝔤μ,E0𝒯FL𝐆,μ,E\mathfrak{g}^{\mathrm{der},0}_{E}\subset\mathfrak{g}^{0}_{\mu,E}\to\mathcal{T}_{\operatorname{FL}_{\mathbf{G},\mu,E}}

    with kernels 𝔭μ,Eder,0\mathfrak{p}^{\mathrm{der},0}_{\mu,E} and 𝔭μ,E0\mathfrak{p}^{0}_{\mu,E} respectively. One can then define 𝔫μ,E0\mathfrak{n}^{0}_{\mu,E} as the unipotent radical of 𝔭Eder,0𝔤Eder,0\mathfrak{p}^{\mathrm{der},0}_{E}\subset\mathfrak{g}^{\mathrm{der},0}_{E} and 𝔪μ,E0=𝔭μ,E0/𝔫μ,E0\mathfrak{m}^{0}_{\mu,E}=\mathfrak{p}^{0}_{\mu,E}/\mathfrak{n}^{0}_{\mu,E} (the reason to take the derived Lie algebra 𝔤der\mathfrak{g}^{\mathrm{der}} is that 𝔤\mathfrak{g} cannot distinguish the Lie algebra of an unipotent group and a torus).

We finish by introducing some notation that will be relevant in Section 5. Given the cocharacter μ\mu of 𝐆C\mathbf{G}_{C} we also have an opposite parabolic subgroup parametrizing increasing μ\mu-filtrations. It is equivalently obtained as 𝐏μ1\mathbf{P}_{\mu^{-1}}. Then, we have the following subgroups of 𝐆C\mathbf{G}_{C}: 𝐍μ1𝐏μ1\mathbf{N}_{\mu^{-1}}\subset\mathbf{P}_{\mu^{-1}} with Levi quotient 𝐌μ1\mathbf{M}_{\mu^{-1}}. Note that 𝐌μ=𝐌μ1\mathbf{M}_{\mu}=\mathbf{M}_{\mu^{-1}} as the centralizers of μ\mu and μ1\mu^{-1} are the same, if the Levi subgroup is clear from the context we will write 𝐌\mathbf{M} instead. We have another flag variety FL𝐆,μ1\operatorname{FL}_{\mathbf{G},\mu^{-1}} and the inverse of the equivalence (2.6) is written as 𝒲𝐆,μ1\mathcal{W}_{\mathbf{G},\mu^{-1}}. To stress the difference between the Lie algebroids we shall write 𝔤μ10:=𝒪FL𝐆,μ1p𝔤\mathfrak{g}^{0}_{\mu^{-1}}:=\mathscr{O}_{\operatorname{FL}_{\mathbf{G},\mu^{-1}}}\otimes_{\mathbb{Q}_{p}}\mathfrak{g}. We also have Lie algebroids over FL𝐆,μ1\operatorname{FL}_{\mathbf{G},\mu^{-1}} given by 𝔫μ10𝔭μ10𝔤μ10\mathfrak{n}^{0}_{\mu^{-1}}\subset\mathfrak{p}_{\mu^{-1}}^{0}\subset\mathfrak{g}^{0}_{\mu^{-1}} and 𝔪μ10=𝔭μ10/𝔫μ10\mathfrak{m}^{0}_{\mu^{-1}}=\mathfrak{p}^{0}_{\mu^{-1}}/\mathfrak{n}^{0}_{\mu^{-1}}.

3. Local Shimura varieties

In this section we introduce local Shimura varieties following [SW20]. We first recall some facts about torsors on the Fargues-Fontaine curve, cf. [FS24, §III.4 and 5]. Then, we recall the definition of moduli spaces of shtukas of one leg from [SW20, Lecture XXIII] as well as the construction of the Grothendieck-Messing and Hodge-Tate period maps. Finally, we specialize the set up to local Shimura varieties and deduce a pp-adic Riemann-Hilbert correspondence for automorphic proétale local systems. This last result is a direct consequence of the theory developed in [SW20], and we only reformulate it in the version that is more convenient for this paper.

Throughout this section we use the notation of [Sch22]. Let Spdp\operatorname{Spd}\mathbb{Z}_{p} denote the vv-sheaf parametrizing untilts SS^{\sharp} of objects SPerfS\in\operatorname{Perf}, we let SpdpSpdp\operatorname{Spd}\mathbb{Q}_{p}\subset\operatorname{Spd}\mathbb{Z}_{p} be the open subspace parametrizing untilts in characteristic zero. Given an analytic adic space XX over p\mathbb{Z}_{p} we let XX^{\lozenge} denote its diamond over Spdp\operatorname{Spd}\mathbb{Z}_{p}. We let k=𝔽¯pk=\overline{\mathbb{F}}_{p} be an algebraic closure of 𝔽p\mathbb{F}_{p} and let Perfk\operatorname{Perf}_{k} be the category of perfectoid spaces over kk. We let ˘p=W(k)[1p]\breve{\mathbb{Q}}_{p}=W(k)[\frac{1}{p}] be the completion of the maximal unramified extension of p\mathbb{Q}_{p}. For E/pE/\mathbb{Q}_{p} a finite extension we write E˘=E˘p\breve{E}=E\breve{\mathbb{Q}}_{p}. We let σ\sigma denote the Frobenius automorphism of kk and ˘p\breve{\mathbb{Q}}_{p}.

3.1. 𝐆\mathbf{G}-torsors over Fargues-Fontaine curves

In this section we recall some facts about torsors over the Fargues-Fontaine curve that we will need throughout the paper. Let 𝐆\mathbf{G} be a reductive group over p\mathbb{Q}_{p} and let B(𝐆)B(\mathbf{G}) be the Kottwitz set of Frobenius-conjugacy classes of elements in 𝐆(˘p)\mathbf{G}(\breve{\mathbb{Q}}_{p}) [Kot97]. Given bB(𝐆)b\in B(\mathbf{G}) and SPerfkS\in\operatorname{Perf}_{k} a perfectoid space we let b\mathcal{E}_{b} denote the 𝐆\mathbf{G}-torsor on 𝒳S\mathcal{X}_{S} obtained via descent from the trivial torsor 𝐆×𝒴(0,),SFF\mathbf{G}\times\mathcal{Y}^{\operatorname{\scriptsize FF}}_{(0,\infty),S} with Frobenius b×φb\times\varphi (in the definition of torsor we take the Tannakian point of view of [SW20, Appendix to Lecture XIX]). Let (br)r(\mathcal{E}_{b}^{\geq r})_{r\in\mathbb{Q}} be the Harder-Narasimhan filtration of b\mathcal{E}_{b}. We take the following definition from [FS24, §5.1].

Definition 3.1.1.

Let bB(𝐆)b\in B(\mathbf{G}). The automorphism group of b\mathcal{E}_{b} is the vv-sheaf on groups

G~b=Aut¯𝐆(b):(SPerfk)Aut𝐆×𝒳SFF(b|𝒳SFF).\widetilde{G}_{b}=\underline{\operatorname{Aut}}_{\mathbf{G}}(\mathcal{E}_{b}):(S\in\operatorname{Perf}_{k})\mapsto\operatorname{Aut}_{\mathbf{G}\times\mathcal{X}^{\operatorname{\scriptsize FF}}_{S}}(\mathcal{E}_{b}|_{\mathcal{X}^{\operatorname{\scriptsize FF}}_{S}}).

Let 𝐆b\mathbf{G}_{b} be the reductive group over p\mathbb{Q}_{p} mapping a ring RR to

𝐆b(R)={g𝐆(Rp˘p)|gb=bσ(g)}.\mathbf{G}_{b}(R)=\{g\in\mathbf{G}(R\otimes_{\mathbb{Q}_{p}}\breve{\mathbb{Q}}_{p})|gb=b\sigma(g)\}.

By [Kot97, §3.3] Gb=𝐆b(p)G_{b}=\mathbf{G}_{b}(\mathbb{Q}_{p}) is the automorphism group associated to the 𝐆\mathbf{G}-isocrystal attached to bb. We have the following structure theorem for the group G~b\widetilde{G}_{b}.

Proposition 3.1.2 ([FS24, Proposition III.5.1]).

One has

G~b=G~b>0Gb\widetilde{G}_{b}=\widetilde{G}_{b}^{>0}\rtimes G_{b}

where G~b>0\widetilde{G}_{b}^{>0} is the subgroup of unipotent automorphisms with respect to the Harder-Narasimhan filtration of b\mathcal{E}_{b}. In particular, if bb is basic, G~b=Gb\widetilde{G}_{b}=G_{b} is the p\mathbb{Q}_{p}-valued points of a pure inner form of 𝐆\mathbf{G}.

In order to construct the period maps we need to introduce the BdR+B^{+}_{\operatorname{dR}}-affine Grassmannian.

Definition 3.1.3 ([SW20, Definition 19.1.1]).

Let SPerfk/SpdES\in\operatorname{Perf}_{k}/\operatorname{Spd}E be an affinoid perfectoid with untilt SS^{\sharp} over E˘\breve{E}. Given HH an algebraic variety over EE we shall write L+HL^{+}H for the vv-sheafification of the presheaf

SH(𝔹dR+(𝒪(S))).S\mapsto H(\mathbb{B}_{\operatorname{dR}}^{+}(\mathscr{O}(S^{\sharp}))).

Similarly we define LHLH to be the vv-sheafification of

SH(𝔹dR(𝒪(S))).S\mapsto H(\mathbb{B}_{\operatorname{dR}}(\mathscr{O}(S^{\sharp}))).

The BdR+B^{+}_{\operatorname{dR}}-affine Grassmannian of 𝐆\mathbf{G} is the vv-sheaf over Spdp\operatorname{Spd}\mathbb{Q}_{p} given by the quotient of groups

Gr𝐆=L𝐆/L+𝐆.\operatorname{Gr}_{\mathbf{G}}=L\mathbf{G}/L^{+}\mathbf{G}.

We shall write Gr𝐆,E˘\operatorname{Gr}_{\mathbf{G},\breve{E}} for the base change of Gr𝐆\operatorname{Gr}_{\mathbf{G}} from Spdp\operatorname{Spd}\mathbb{Q}_{p} to SpdE\operatorname{Spd}E.

Let SPerfkS\in\operatorname{Perf}_{k} be a perfectoid and let SS^{\sharp} be an untilt over ˘p\breve{\mathbb{Q}}_{p}. By [FS24, Proposition II.1.18] the map

ι:S𝒳SFF\iota:S^{\sharp}\to\mathcal{X}^{\operatorname{\scriptsize FF}}_{S}

is an effective Cartier divisor. The pullback of b\mathcal{E}_{b} to the completion 𝒳SFF,,ι\mathcal{X}_{S}^{\operatorname{\scriptsize FF},\wedge,\iota} of 𝒳SFF\mathcal{X}^{\operatorname{\scriptsize FF}}_{S} at ι\iota is a trivial 𝐆\mathbf{G}-torsor. The automorphism group of the trivial 𝐆\mathbf{G}-torsor over 𝒳SFF,,ι\mathcal{X}_{S}^{\operatorname{\scriptsize FF},\wedge,\iota} is then equal to L+𝐆(S)L^{+}\mathbf{G}(S). Thus, pullback along the formal completion gives rise to a group homomorphism of vv-sheaves

G~b×SpdpL+𝐆.\widetilde{G}_{b}\times\operatorname{Spd}\mathbb{Q}_{p}\to L^{+}\mathbf{G}. (3.1)

3.2. Moduli space of shtukas of one leg

In the following section we recall the definition of moduli space of shtukas of one leg and the construction of the Grothendieck-Messing and Hodge-Tate period maps. We shall follow [SW20, §23.3].

Let (𝐆,b,μ)(\mathbf{G},b,\mu) be a local shtuka datum of one leg, namely, a triple consisting on a reductive group 𝐆\mathbf{G} over p\mathbb{Q}_{p}, an element bB(𝐆)b\in B(\mathbf{G}), and a conjugacy class of cocharacters μ:𝔾m𝐆¯p\mu:\mathbb{G}_{m}\to\mathbf{G}_{\overline{\mathbb{Q}}_{p}}. Let EE be the field of definition of μ\mu. Recall the equivalent definition of the moduli space of shtukas of one leg from [SW20, Proposition 23.3.1].

Definition 3.2.1.

Let K𝐆(p)K\subset\mathbf{G}(\mathbb{Q}_{p}) be a compact open subgroup. The moduli space Sht𝐆,b,μ,K\operatorname{Sht}_{\mathbf{G},b,\mu,K} of stukas associated to (𝐆,b,μ)(\mathbf{G},b,\mu) at level KK is the presheaf on Perfk\operatorname{Perf}_{k} mapping SPerfkS\in\operatorname{Perf}_{k} to the isomorphism classes of quadruples (S,,α,)(S^{\sharp},\mathcal{E},\alpha,\mathbb{P}) where

  • SS^{\sharp} is an until of SS over E˘\breve{E},

  • \mathcal{E} is a 𝐆\mathbf{G}-torsor on 𝒳SFF\mathcal{X}^{\operatorname{\scriptsize FF}}_{S}, which is trivial at every geometric point of SS,

  • α\alpha is an isomorphism of 𝐆\mathbf{G}-torsors

    α:|𝒳SFF\Sb|𝒳SFF\S,\alpha:\mathcal{E}|_{\mathcal{X}^{\operatorname{\scriptsize FF}}_{S}\backslash S^{\sharp}}\xrightarrow{\sim}\mathcal{E}_{b}|_{\mathcal{X}^{\operatorname{\scriptsize FF}}_{S}\backslash S^{\sharp}},

    which is meromorphic at SS^{\sharp} and bounded by μ\mu, and finally

  • \mathbb{P} is a KK-lattice in the proétale 𝐆(p)\mathbf{G}(\mathbb{Q}_{p})-torsor corresponding to \mathcal{E} via [SW20, Theorem 22.5.2].

By [SW20, Theorem 23.1.4] the spaces Sht𝐆,b,μ,K\operatorname{Sht}_{\mathbf{G},b,\mu,K} are diamonds living over SpdE˘\operatorname{Spd}\breve{E}. We also define the moduli space of shtukas at infinite level.

Definition 3.2.2.

Let Sht𝐆,b,μ,=limK𝐆(p)Sht𝐆,b,μ,K\operatorname{Sht}_{\mathbf{G},b,\mu,\infty}=\varprojlim_{K\subset\mathbf{G}(\mathbb{Q}_{p})}\operatorname{Sht}_{\mathbf{G},b,\mu,K} be the infinite level moduli space of shtukas. By construction, Sht𝐆,b,μ,\operatorname{Sht}_{\mathbf{G},b,\mu,\infty} is the presheaf on Perfk\operatorname{Perf}_{k} parametrizing tuples (S,α)(S^{\sharp},\alpha) where

  • SS^{\sharp} is an untilt of SS over E˘\breve{E}.

  • α\alpha is an isomorphism of 𝐆\mathbf{G}-torsors

    α:1|𝒳SFF\Sb|𝒳SFF\S\alpha:\mathcal{E}_{1}|_{\mathcal{X}^{\operatorname{\scriptsize FF}}_{S}\backslash S^{\sharp}}\xrightarrow{\sim}\mathcal{E}_{b}|_{\mathcal{X}^{\operatorname{\scriptsize FF}}_{S}\backslash S^{\sharp}}

    which is meromorphic at SS^{\sharp} and bounded by μ\mu.

Let SPerfkS\in\operatorname{Perf}_{k} be a perfectoid space and let (S,α)(S^{\sharp},\alpha) be an SS-point of Sht𝐆,b,μ,\operatorname{Sht}_{\mathbf{G},b,\mu,\infty}. Let ι:S𝒳SFF\iota:S^{\sharp}\to\mathcal{X}^{\operatorname{\scriptsize FF}}_{S} be the closed Cartier divisor defined by the untilt. The pullbacks of 1\mathcal{E}_{1} and b\mathcal{E}_{b} to the formal completion 𝒳SFF,,ι\mathcal{X}_{S}^{\operatorname{\scriptsize FF},\wedge,\iota} at ι\iota are trivial 𝐆\mathbf{G}-torsors. Therefore, the modification α\alpha is defined by an element gαL𝐆(S)=𝔹dR(S)g_{\alpha}\in L\mathbf{G}(S)=\mathbb{B}_{\operatorname{dR}}(S^{\sharp}). The automorphisms 𝐆(p)\mathbf{G}(\mathbb{Q}_{p}) of 1\mathcal{E}_{1} act on gαg_{\alpha} by right multiplication while the automorphisms G~b\widetilde{G}_{b} of b\mathcal{E}_{b} act by left multiplication. Therefore, we have two maps to the affine BdRB_{\operatorname{dR}}-grassmannian

Sht𝐆,b,μ,{\operatorname{Sht}_{\mathbf{G},b,\mu,\infty}}L𝐆/L+𝐆=Gr𝐆,E˘{L\mathbf{G}/L^{+}\mathbf{G}=\operatorname{Gr}_{\mathbf{G},\breve{E}}}Gr𝐆,E˘=L+𝐆\L𝐆{\operatorname{Gr}_{\mathbf{G},\breve{E}}=L^{+}\mathbf{G}\backslash L\mathbf{G}}πGM\scriptstyle{\pi_{\operatorname{\scriptsize GM}}}πHT\scriptstyle{\pi_{\operatorname{\scriptsize HT}}} (3.2)

by taking a left or right coset respectively. In particular, since α\alpha is bounded by μ\mu by hypothesis, the diagram (3.2) actually restricts to

Sht𝐆,b,μ,{\operatorname{Sht}_{\mathbf{G},b,\mu,\infty}}Gr𝐆,E˘,μ{\operatorname{Gr}_{\mathbf{G},\breve{E},\leq\mu}}Gr𝐆,E˘,μ1.{\operatorname{Gr}_{\mathbf{G},\breve{E},\leq\mu^{-1}}.}πGM\scriptstyle{\pi_{\operatorname{\scriptsize GM}}}πHT\scriptstyle{\pi_{\operatorname{\scriptsize HT}}} (3.3)

When bb is basic we have a duality for the diagram (3.3), cf. [SW20, Corollary 23.3.2].

Proposition 3.2.3.

Let (𝐆,b,μ)(\mathbf{G},b,\mu) be a local shtuka datum with bb basic. Define a shtuka datum (𝐆ˇ,bˇ,μˇ)(\check{\mathbf{G}},\check{b},\check{\mu}) via 𝐆ˇ=𝐆b\check{\mathbf{G}}=\mathbf{G}_{b}, bˇ=b1𝐆b(˘p)=𝐆(˘p)\check{b}=b^{-1}\in\mathbf{G}_{b}(\breve{\mathbb{Q}}_{p})=\mathbf{G}(\breve{\mathbb{Q}}_{p}) and μˇ=μ1\check{\mu}=\mu^{-1} under the identification 𝐆¯p𝐆ˇ¯p\mathbf{G}_{\overline{\mathbb{Q}}_{p}}\cong\check{\mathbf{G}}_{\overline{\mathbb{Q}}_{p}}. Then there is a natural 𝐆(p)×𝐆ˇ(p)\mathbf{G}(\mathbb{Q}_{p})\times\check{\mathbf{G}}(\mathbb{Q}_{p})-equivariant isomorphism

Sht𝐆,b,μ,Sht𝐆ˇ,bˇ,μˇ,\operatorname{Sht}_{\mathbf{G},b,\mu,\infty}\cong\operatorname{Sht}_{\check{\mathbf{G}},\check{b},\check{\mu},\infty} (3.4)

interchanging the maps πGM\pi_{\operatorname{\scriptsize GM}} and πHT\pi_{\operatorname{\scriptsize HT}} of (3.3).

Proof.

Let SPerfkS\in\operatorname{Perf}_{k} and (S,α)(S^{\sharp},\alpha) an SS-point of Sht𝐆,b,μ,\operatorname{Sht}_{\mathbf{G},b,\mu,\infty}. The equivariant isomorphism is [SW20, Corollary 23.3.2]. It is given by mapping a modification

α:1|𝒳SFF\Sb|𝒳SFF\S\alpha:\mathcal{E}_{1}|_{\mathcal{X}^{\operatorname{\scriptsize FF}}_{S}\backslash S^{\sharp}}\xrightarrow{\sim}\mathcal{E}_{b}|_{\mathcal{X}^{\operatorname{\scriptsize FF}}_{S}\backslash S^{\sharp}}

to the modification of 𝐆b\mathbf{G}_{b}-torsors

αˇ:1|𝒳SFF\Sbˇ|𝒳SFF\S\check{\alpha}:\mathcal{F}_{1}|_{\mathcal{X}^{\operatorname{\scriptsize FF}}_{S}\backslash S^{\sharp}}\xrightarrow{\sim}\mathcal{F}_{\check{b}}|_{\mathcal{X}^{\operatorname{\scriptsize FF}}_{S}\backslash S^{\sharp}}

obtained by mapping a 𝐆\mathbf{G}-torsor \mathcal{E} to the 𝐆b\mathbf{G}_{b}-torsor =Aut𝐆(,b)\mathcal{F}=\operatorname{Aut}_{\mathbf{G}}(\mathcal{E},\mathcal{E}_{b}). Then, since the pullback of the torsors 1\mathcal{E}_{1} and 2\mathcal{E}_{2} to 𝔹dR(S)\mathbb{B}_{\operatorname{dR}}(S^{\sharp}) are trivial, the map αˇ\check{\alpha} seen as an object in 𝐆(𝔹dR(S))𝐆ˇ(𝔹dR(S))\mathbf{G}(\mathbb{B}_{\operatorname{dR}}(S^{\sharp}))\cong\check{\mathbf{G}}(\mathbb{B}_{\operatorname{dR}}(S^{\sharp})) is just the inverse of the map α\alpha, proving that the period morphisms πGM\pi_{\operatorname{\scriptsize GM}} and πHT\pi_{\operatorname{\scriptsize HT}} are exchanged. ∎

3.3. Local Shimura varieties

Recall the definition of a local Shimura datum [SW20, Definition 24.1.1]

Definition 3.3.1.

A local Shimura datum is a triple (𝐆,b,μ)(\mathbf{G},b,\mu) consisting of a reductive group 𝐆\mathbf{G} over p\mathbb{Q}_{p}, a conjugacy class μ\mu of minuscule cocharacters 𝔾m𝐆¯p\mathbb{G}_{m}\to\mathbf{G}_{\overline{\mathbb{Q}}_{p}}, and an element bB(𝐆,μ1)b\in B(\mathbf{G},\mu^{-1}).

Let (𝐆,b,μ)(\mathbf{G},b,\mu) be a local Shimura datum and let EE be the field of definition of μ\mu. We keep the representation theory notation of Section 2.5. Let K𝐆(p)K\subset\mathbf{G}(\mathbb{Q}_{p}) be an open compact subgroup and consider Sht𝐆,b,μ,K\operatorname{Sht}_{\mathbf{G},b,\mu,K} the moduli space of shtukas associated to (𝐆,b,μ)(\mathbf{G},b,\mu) at level KK. By [SW20, Proposition 23.3.3] the period map

πGM:Sht𝐆,b,μ,KGr𝐆,E˘,μ\pi_{\operatorname{\scriptsize GM}}:\operatorname{Sht}_{\mathbf{G},b,\mu,K}\to\operatorname{Gr}_{\mathbf{G},\breve{E},\leq\mu}

is étale. On the other hand, since μ\mu is minuscule, the Bialynicki-Birula map

πμ:Gr𝐆,E˘,μ=Gr𝐆,E˘,μ𝐆,μ,E˘\pi_{\mu}:\operatorname{Gr}_{\mathbf{G},\breve{E},\leq\mu}=\operatorname{Gr}_{\mathbf{G},\breve{E},\mu}\to\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu,\breve{E}}^{\lozenge}

is an isomorphism [SW20, Proposition 19.4.2]. This produces an étale map

πGM:Sht𝐆,b,μ,K𝐆,μ,E˘.\pi_{\operatorname{\scriptsize GM}}:\operatorname{Sht}_{\mathbf{G},b,\mu,K}\to\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu,\breve{E}}^{\lozenge}. (3.5)
Definition 3.3.2.

For K𝐆(p)K\subset\mathbf{G}(\mathbb{Q}_{p}) let 𝐆,b,μ,K\mathcal{M}_{\mathbf{G},b,\mu,K} be the unique smooth rigid space over E˘\breve{E} endowed with an étale map 𝐆,b,μ,K𝐆,μ,E˘\mathcal{M}_{\mathbf{G},b,\mu,K}\to\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu,\breve{E}} such that

Sht𝐆,b,μ,K𝐆,b,μ,K\operatorname{Sht}_{\mathbf{G},b,\mu,K}\cong\mathcal{M}_{\mathbf{G},b,\mu,K}^{\lozenge}

as diamonds over 𝐆,μ,E˘\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu,\breve{E}}^{\lozenge}. We shall write 𝐆,b,μ,=limK𝐆,b,μ,K\mathcal{M}_{\mathbf{G},b,\mu,\infty}^{\lozenge}=\varprojlim_{K}\mathcal{M}_{\mathbf{G},b,\mu,K}^{\lozenge} for the infinite level Shimura variety.

By (3.2) we get a 𝐆(p)×G~b\mathbf{G}(\mathbb{Q}_{p})\times\widetilde{G}_{b}-equivariant diagram of period maps

𝐆,b,μ,{\mathcal{M}_{\mathbf{G},b,\mu,\infty}^{\lozenge}}𝐆,μ,E˘{\operatorname{\mathcal{F}\ell}^{\lozenge}_{\mathbf{G},\mu,\breve{E}}}𝐆,μ1,E˘.{\operatorname{\mathcal{F}\ell}^{\lozenge}_{\mathbf{G},\mu^{-1},\breve{E}}.}πHT\scriptstyle{\pi_{\operatorname{\scriptsize HT}}}πGM\scriptstyle{\pi_{\operatorname{\scriptsize GM}}} (3.6)

In the rest of the section we will translate the diagram (3.6) in terms of pp-adic Hodge theory of Shimura varieties. More precisely, we shall deduce a Riemann-Hilbert correspondence for proétale local systems arising from algebraic 𝐆\mathbf{G}-representations.

Let 𝐆,μ,E˘a𝐆,μ,E˘\operatorname{\mathcal{F}\ell}^{a}_{\mathbf{G},\mu,\breve{E}}\subset\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu,\breve{E}} be the admissible locus of the flag variety. By [SW20, Corollary 23.5.3] the map πGM\pi_{\operatorname{\scriptsize GM}} factors through 𝐆,μ,E˘a\operatorname{\mathcal{F}\ell}^{a}_{\mathbf{G},\mu,\breve{E}} and the map

πGM:𝐆,b,μ,𝐆,μ,E˘a\pi_{\operatorname{\scriptsize GM}}:\mathcal{M}^{\lozenge}_{\mathbf{G},b,\mu,\infty}\to\operatorname{\mathcal{F}\ell}^{a}_{\mathbf{G},\mu,\breve{E}}

is a proétale 𝐆(p)\mathbf{G}(\mathbb{Q}_{p})-torsor.

Definition 3.3.3.

For K𝐆(p)K\subset\mathbf{G}(\mathbb{Q}_{p}) a closed subgroup we denote

𝐆,b,μ,K:=𝐆,b,μ,/K.\mathcal{M}^{\lozenge}_{\mathbf{G},b,\mu,K}:=\mathcal{M}^{\lozenge}_{\mathbf{G},b,\mu,\infty}/K.

Let VV be a pp-adically complete or p\mathbb{Q}_{p}-Banach continuous representation of KK. We let V\mathcal{F}_{V} be the vv-sheaf on 𝐆,b,μ,K\mathcal{M}^{\lozenge}_{\mathbf{G},b,\mu,K} obtained via descent from the constant KK-equivariant sheaf V¯\underline{V} on 𝐆,μ,\mathcal{M}_{\mathbf{G},\mu,\infty}^{\lozenge} with

V¯(S)=Cont(|S|,V)\underline{V}(S)=\operatorname{{Cont}}(|S|,V)

for S𝐆,b,μ,,vS\in\mathcal{M}^{\lozenge}_{\mathbf{G},b,\mu,\infty,v} affinoid perfectoid. For V=``limi′′ViV=``\varinjlim_{i}^{\prime\prime}V_{i} an ind-system of pp-adically complete or Banach representations we define V:=limiVi\mathcal{F}_{V}:=\varinjlim_{i}\mathcal{F}_{V_{i}}.

We let EE^{\prime} be a finite extension over EE over which 𝐆\mathbf{G} is split and consider the pullback of the local Shimura varieties and flag varieties to E˘\breve{E}^{\prime}. Furthermore, we fix a Hodge cocharacter μ:𝔾m,E𝐆E\mu:\mathbb{G}_{m,E^{\prime}}\to\mathbf{G}_{E^{\prime}} which determines parabolic subgroups 𝐏μ\mathbf{P}_{\mu} and 𝐏μ1\mathbf{P}_{\mu^{-1}} of 𝐆E\mathbf{G}_{E^{\prime}}, as well as their unipotent radicals 𝐍μ\mathbf{N}_{\mu} and 𝐍μ1\mathbf{N}_{\mu^{-1}}, and the Levi subgroup 𝐌μ=𝐌μ1=𝐌\mathbf{M}_{\mu}=\mathbf{M}_{\mu^{-1}}=\mathbf{M} 111Taking this finite extension is unnecessary for the forthcoming discussion but it allows us to use the dictionary between representations of the chosen parabolic 𝐏μ\mathbf{P}_{\mu} and 𝐆\mathbf{G}-equivariant quasi-coherent sheaves on the flag variety of Section 2.5. We left to the reader the cocharacter-free formulation of the statements in terms of filtered 𝐆\mathbf{G}-representations..

Let VReppalg(𝐆)V\in\operatorname{Rep}_{\mathbb{Q}_{p}}^{\operatorname{\scriptsize alg}}(\mathbf{G}) be a p\mathbb{Q}_{p}-linear algebraic representation of 𝐆\mathbf{G}. Let VdRV_{\operatorname{dR}} be the 𝐆\mathbf{G}-equivariant flat connection over 𝐆,μ,E˘\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu,\breve{E}^{\prime}} given by

VdR=𝒪𝐆,μ,E˘pVV_{\operatorname{dR}}=\mathscr{O}_{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu},\breve{E}^{\prime}}\otimes_{\mathbb{Q}_{p}}V

with Hodge filtration induced by the 𝐏μ\mathbf{P}_{\mu}-filtration of VpEV\otimes_{\mathbb{Q}_{p}}E^{\prime} via the functor 𝒲𝐆,μ\mathcal{W}_{\mathbf{G},\mu} of (2.6), see Remark 2.5.1 (1). By an abuse of notation we will also write VdRV_{\operatorname{dR}} for the restriction to the admissible locus.

Let V\mathcal{F}_{V} be the vv-sheaf over 𝐆,μ,E˘a\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu,\breve{E}^{\prime}}^{a} associated to the 𝐆(p)\mathbf{G}(\mathbb{Q}_{p})-representation VV via Definition 3.3.3. Let Vproe´tV_{\operatorname{\scriptsize pro\acute{e}t}} be the restriction of V\mathcal{F}_{V} to a sheaf on the proétale site 𝐆,μ,E˘,proe´ta\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu,\breve{E}^{\prime},\operatorname{\scriptsize pro\acute{e}t}}^{a} of [Sch13]. We have the following Riemann-Hilbert correspondence for local Shimura varieties.

Proposition 3.3.4.

The proétale local system Vproe´tV_{\operatorname{\scriptsize pro\acute{e}t}} is de Rham in the sense of [Sch13, Definition 8.3] with associated filtered flat connection VdRV_{\operatorname{dR}} 222Strictly speaking this notion is only defined for lisse p\mathbb{Z}_{p}-local systems and Vproe´tV_{\operatorname{\scriptsize pro\acute{e}t}} over 𝐆,μ,E˘a\operatorname{\mathcal{F}\ell}^{a}_{\mathbf{G},\mu,\breve{E}^{\prime}} is not of this form. However, it becomes lisse after pulling back to any finite level local Shimura variety.. More precisely, we have a G~b\widetilde{G}_{b}-equivariant map of filtered 𝔹dR\mathbb{B}_{\operatorname{dR}}-sheaves on 𝐆,μ,E˘,proe´ta\operatorname{\mathcal{F}\ell}^{a}_{\mathbf{G},\mu,\breve{E}^{\prime},\operatorname{\scriptsize pro\acute{e}t}}

Vproe´tp𝔹dRV¯p𝔹dR,V_{\operatorname{\scriptsize pro\acute{e}t}}\otimes_{\mathbb{Q}_{p}}\mathbb{B}_{\operatorname{dR}}\cong\underline{V}\otimes_{\mathbb{Q}_{p}}\mathbb{B}_{\operatorname{dR}}, (3.7)

where the 𝔹dR+\mathbb{B}_{\operatorname{dR}}^{+}-filtration in the left hand side is the trivial one, and the filtration in the right hand side is given by

Fili(V¯p𝔹dR):=(Fili(VdR𝒪𝐆,μ,E˘a𝒪𝔹dR))=0.\mathrm{Fil}^{i}(\underline{V}\otimes_{\mathbb{Q}_{p}}\mathbb{B}_{\operatorname{dR}}):=(\mathrm{Fil}^{i}(V_{\operatorname{dR}}\otimes_{\mathscr{O}_{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu,\breve{E}^{\prime}}^{a}}}\mathscr{O}\!\mathbb{B}_{\operatorname{dR}}))^{\nabla=0}.

The action of G~b\widetilde{G}_{b} is trivial on Vproe´tV_{\operatorname{\scriptsize pro\acute{e}t}} in the left hand side and it factors through G~bL𝐆\widetilde{G}_{b}\to L\mathbf{G} and the natural action on V¯p𝔹dR\underline{V}\otimes_{\mathbb{Q}_{p}}\mathbb{B}_{\operatorname{dR}} in the right hand side.

Remark 3.3.5.

By [SW20, Corollary 17.1.9], for a smooth rigid variety XX there is no distinction between filtered 𝔹dR+\mathbb{B}^{+}_{\operatorname{dR}}-vector bundles on the proétale site Xproe´tX_{\operatorname{\scriptsize pro\acute{e}t}} of [Sch13] or filtered 𝔹dR+\mathbb{B}_{\operatorname{dR}}^{+}-vector bundles on XvX_{v}. Thus, the equivariant isomorphism (3.7) can also be stated as a G~b\widetilde{G}_{b}-equivariant isomorphism of filtered 𝔹dR\mathbb{B}_{\operatorname{dR}} sheaves on the vv-site

Vp𝔹dRV¯p𝔹dR.\mathcal{F}_{V}\otimes_{\mathbb{Q}_{p}}\mathbb{B}_{\operatorname{dR}}\cong\underline{V}\otimes_{\mathbb{Q}_{p}}\mathbb{B}_{\operatorname{dR}}.
Proof of Proposition 3.3.4.

Let us denote X=𝐆,μ,E˘,proe´taX=\operatorname{\mathcal{F}\ell}^{a}_{\mathbf{G},\mu,\breve{E}^{\prime},\operatorname{\scriptsize pro\acute{e}t}}. Since VdRV_{\operatorname{dR}} has horizontal sections VV, we have an isomorphism

V¯p𝔹dR=(VdR𝒪X𝒪𝔹dR)=0\underline{V}\otimes_{\mathbb{Q}_{p}}\mathbb{B}_{\operatorname{dR}}=(V_{\operatorname{dR}}\otimes_{\mathscr{O}_{X}}\mathscr{O}\!\mathbb{B}_{\operatorname{dR}})^{\nabla=0}

in Xproe´tX_{\operatorname{\scriptsize pro\acute{e}t}}. By [Sch13, Theorem 7.6] 𝕄:=(Fil0(VdR𝒪X𝒪𝔹dR))=0\mathbb{M}^{\prime}:=(\mathrm{Fil}^{0}(V_{\operatorname{dR}}\otimes_{\mathscr{O}_{X}}\mathscr{O}\!\mathbb{B}_{\operatorname{dR}}))^{\nabla=0} is a 𝔹dR+\mathbb{B}_{\operatorname{dR}}^{+}-lattice of V¯p𝔹dR\underline{V}\otimes_{\mathbb{Q}_{p}}\mathbb{B}_{\operatorname{dR}} in the proétale site of XX. By [SW20, Corollary 17.1.9] we can view 𝕄\mathbb{M}^{\prime} as a 𝔹dR+\mathbb{B}_{\operatorname{dR}}^{+}-lattice in the vv-site of XX. Thus, we will view V¯p𝔹dR\underline{V}\otimes_{\mathbb{Q}_{p}}\mathbb{B}_{\operatorname{dR}} as a filtered 𝔹dR+\mathbb{B}_{dR}^{+}-module in the vv-site with Fili(V¯p𝔹dR)=ξi𝕄\mathrm{Fil}^{i}(\underline{V}\otimes_{\mathbb{Q}_{p}}\mathbb{B}_{\operatorname{dR}})=\xi^{i}\mathbb{M}^{\prime} for ξ\xi a local generator of the kernel of θ:𝔹dR+𝒪^\theta:\mathbb{B}_{\operatorname{dR}}^{+}\to\widehat{\mathscr{O}}.

Now let SPerfkS\in\operatorname{Perf}_{k} and take (S,α)(S^{\sharp},\alpha) an SS-point of 𝐆,b,μ,,E˘\mathcal{M}^{\lozenge}_{\mathbf{G},b,\mu,\infty,\breve{E}^{\prime}}. Given VReppalg𝐆V\in\operatorname{Rep}^{\operatorname{\scriptsize alg}}_{\mathbb{Q}_{p}}\mathbf{G} an algebraic representation, let 𝒱1\mathcal{V}_{1} and 𝒱b\mathcal{V}_{b} be the 𝐆\mathbf{G}-equivariant vector bundles over 𝒳SFF\mathcal{X}^{\operatorname{\scriptsize FF}}_{S} defined by the torsors 1\mathcal{E}_{1} and b\mathcal{E}_{b} respectively. By definition of the modification we have a 𝐆(p)×G~b\mathbf{G}(\mathbb{Q}_{p})\times\widetilde{G}_{b}-equivariant isomorphism

(Vp𝔹dR)(S)=𝒱1𝒪𝒳SFF𝔹dR(S)𝒱b𝒪𝒳SFF𝔹dR(S)=V¯p𝔹dR(S)(\mathcal{F}_{V}\otimes_{\mathbb{Q}_{p}}\mathbb{B}_{\operatorname{dR}})(S^{\sharp})=\mathcal{V}_{1}\otimes_{\mathscr{O}_{\mathcal{X}^{\operatorname{\scriptsize FF}}_{S}}}\mathbb{B}_{\operatorname{dR}}(S^{\sharp})\cong\mathcal{V}_{b}\otimes_{\mathscr{O}_{\mathcal{X}^{\operatorname{\scriptsize FF}}_{S}}}\mathbb{B}_{\operatorname{dR}}(S^{\sharp})=\underline{V}\otimes_{\mathbb{Q}_{p}}\mathbb{B}_{\operatorname{dR}}(S^{\sharp})

where 𝐆(p)\mathbf{G}(\mathbb{Q}_{p}) acts trivially on V¯\underline{V} and via the projection 𝐆(p)L𝐆\mathbf{G}(\mathbb{Q}_{p})\to L\mathbf{G} on Vp𝔹dR\mathcal{F}_{V}\otimes_{\mathbb{Q}_{p}}\mathbb{B}_{dR}, and G~b\widetilde{G}_{b} acts trivially on V\mathcal{F}_{V} and via the projection G~bL𝐆\widetilde{G}_{b}\to L\mathbf{G} on V¯p𝔹dR\underline{V}\otimes_{\mathbb{Q}_{p}}\mathbb{B}_{dR}. This produces the desired equivariant isomorphism (3.7). The fact that the isomorphism (3.7) is compatible with the filtration follows from the definition of the Bialynicki-Birula map and [SW20, Proposition 19.4.2]. ∎

Given VReppalg𝐆V\in\operatorname{Rep}^{\operatorname{\scriptsize alg}}_{\mathbb{Q}_{p}}\mathbf{G} let 𝕄=Vp𝔹dR+\mathbb{M}=\mathcal{F}_{V}\otimes_{\mathbb{Q}_{p}}\mathbb{B}^{+}_{\operatorname{dR}} and let 𝕄0=V¯p𝔹dR+\mathbb{M}_{0}=\underline{V}\otimes_{\mathbb{Q}_{p}}\mathbb{B}_{\operatorname{dR}}^{+}. The Hodge-Tate filtration of Vp𝒪^\mathcal{F}_{V}\otimes_{\mathbb{Q}_{p}}\widehat{\mathscr{O}} is given by

Filn(Vp𝒪^)=(𝕄Fili𝕄0)/(Fil1𝕄Fili𝕄0).\mathrm{Fil}_{n}(\mathcal{F}_{V}\otimes_{\mathbb{Q}_{p}}\widehat{\mathscr{O}})=(\mathbb{M}\cap\mathrm{Fil}^{-i}\mathbb{M}_{0})/(\mathrm{Fil}^{1}\mathbb{M}\cap\mathrm{Fil}^{-i}\mathbb{M}_{0}).

Let 𝒲𝐆,μ1\mathcal{W}_{\mathbf{G},\mu^{-1}} be the inverse of the functor (2.6) for the cocharacter μ1\mu^{-1}. As corollary we deduce that the Hodge-Tate period map encodes the Hodge-Tate filtration.

Corollary 3.3.6.

There is a 𝐆(p)×G~b\mathbf{G}(\mathbb{Q}_{p})\times\widetilde{G}_{b}-equivariant isomorphism of 𝒪^\widehat{\mathscr{O}}-modules over 𝐆,b,μ,,E˘\mathcal{M}^{\lozenge}_{\mathbf{G},b,\mu,\infty,\breve{E}^{\prime}}

Filn(Vp𝒪^)πHT(𝒲𝐆,μ1(Filn(VpE)))),\mathrm{Fil}_{n}(\mathcal{F}_{V}\otimes_{\mathbb{Q}_{p}}\widehat{\mathscr{O}})\cong\pi_{\operatorname{\scriptsize HT}}^{*}(\mathcal{W}_{\mathbf{G},\mu^{-1}}(\mathrm{Fil}_{n}(V\otimes_{\mathbb{Q}_{p}}E^{\prime})))),

where Fil(VpE)\mathrm{Fil}_{\bullet}(V\otimes_{\mathbb{Q}_{p}}E^{\prime}) is the (increasing) 𝐏μ1\mathbf{P}_{\mu^{-1}}-filtration of VpEV\otimes_{\mathbb{Q}_{p}}E^{\prime}.

Proof.

This is a consequence of Proposition 3.3.4, the definition of the Bialynicki-Birula map and [SW20, Proposition 19.4.2]

Moreover, taking graded pieces in Corollary 3.3.6 one deduces the isomorphism of 𝐌\mathbf{M}-torsors on the infinite level Shimura variety, cf. [CS17, Theorem 2.1.3].

Corollary 3.3.7.

Let WRepEalg𝐌W\in\operatorname{Rep}^{\operatorname{\scriptsize alg}}_{E^{\prime}}\mathbf{M} be an irreducible algebraic representation of the Levi subgroup. There is a natural 𝐆(p)×G~b\mathbf{G}(\mathbb{Q}_{p})\times\widetilde{G}_{b}-equivariant \otimes-isomorphism of 𝒪^\widehat{\mathscr{O}}-modules over 𝐆,b,μ,,E˘\mathcal{M}^{\lozenge}_{\mathbf{G},b,\mu,\infty,\breve{E}^{\prime}}

πHT(𝒲𝐆,μ1(W))πGM(𝒲𝐆,μ(W))𝒪^𝒪^(μ(W))\pi_{\operatorname{\scriptsize HT}}^{*}(\mathcal{W}_{\mathbf{G},\mu^{-1}}(W))\cong\pi_{\operatorname{\scriptsize GM}}^{*}(\mathcal{W}_{\mathbf{G},\mu}(W))\otimes_{\widehat{\mathscr{O}}}\widehat{\mathscr{O}}(-\mu(W))

where μ(W)\mu(W)\in\mathbb{Z} is the weight of WW with respect to μ\mu. In particular, if 𝐌μ,GM\mathbf{M}_{\mu,\operatorname{\scriptsize GM}} and 𝐌μ1,HT\mathbf{M}_{\mu^{-1},\operatorname{\scriptsize HT}} denote the natural 𝐌\mathbf{M}-torsors living over 𝐆,μ,E˘\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu,\breve{E}^{\prime}} and 𝐆,μ1,E˘\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu^{-1},\breve{E}^{\prime}} respectively (see Section 2.5), we have a 𝐆(p)×G~b\mathbf{G}(\mathbb{Q}_{p})\times\widetilde{G}_{b}-equivariant isomorphism of 𝐌\mathbf{M}-torsors over the ringed site (𝐆,b,μ,,E˘,v,𝒪^)(\mathcal{M}^{\lozenge}_{\mathbf{G},b,\mu,\infty,\breve{E}^{\prime},v},\widehat{\mathscr{O}})

πHT(𝐌μ1,HT)πGM(𝐌μ,GM)×𝔾m,μ𝔾m(1),\pi_{\operatorname{\scriptsize HT}}^{*}(\mathbf{M}_{\mu^{-1},\operatorname{\scriptsize HT}})\cong\pi_{\operatorname{\scriptsize GM}}^{*}(\mathbf{M}_{\mu,\operatorname{\scriptsize GM}})\times^{\mathbb{G}_{m},\mu}\mathbb{G}_{m}(-1),

where 𝔾m\mathbb{G}_{m} injects into the center of 𝐌\mathbf{M} via μ\mu, and 𝔾m(1)\mathbb{G}_{m}(-1) is the 𝔾m\mathbb{G}_{m}-torsor of trivializations of the Tate twist 𝒪^(1)\widehat{\mathscr{O}}(-1).

Proof.

This follows after taking graded pieces of the isomorphisms in Corollary 3.3.6 and [Sch13, Proposition 7.9], see also [RC24b, Theorem 4.2.1]. ∎

4. Geometric Sen operators of local Shimura varieties

In this section we compute the geometric Sen operator of local Shimura varieties, proving the local analogue of [RC24b, Theorem 5.2.5]. We keep the notation of Section 3.3, namely we let (𝐆,b,μ)(\mathbf{G},b,\mu) be a local Shimura datum with reflex field E/pE/\mathbb{Q}_{p}. We let E/EE^{\prime}/E be a finite extension of EE over which 𝐆\mathbf{G} is split and fix a representative of the Hodge-cocharacter μ:𝔾m,E𝐆E\mu:\mathbb{G}_{m,E^{\prime}}\to\mathbf{G}_{E^{\prime}}. For K𝐆(p)K\subset\mathbf{G}(\mathbb{Q}_{p}) compact open subgroup we let 𝐆,b,μ,K\mathcal{M}_{\mathbf{G},b,\mu,K} be the local Shimura variety over E˘\breve{E} at level KK, we denote by 𝒪\mathscr{O}_{\mathcal{M}} its structural sheaf as a rigid space, and let Ω1\Omega_{\mathcal{M}}^{1} be its cotangent bundle.

4.1. The Kodaira-Spencer map

In the next paragraph we make explicit the Kodaira-Spencer isomorphism for local Shimura varieties in terms of representation theory over the flag variety. We follow [RC24b, Proposition 5.1.3].

Let 𝔤\mathfrak{g} be the adjoint representation of 𝐆\mathbf{G} over p\mathbb{Q}_{p}, and let 𝔤dR=𝒪𝐆,μ,Ep𝔤=𝔤μ0,\mathfrak{g}^{\vee}_{\operatorname{dR}}=\mathscr{O}_{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu,E^{\prime}}}\otimes_{\mathbb{Q}_{p}}\mathfrak{g}^{\vee}=\mathfrak{g}^{0,\vee}_{\mu} be its associated 𝐆\mathbf{G}-equivariant vector bundle with flat connection over 𝐆,μ,E\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu,E^{\prime}}. Since μ\mu is minuscule, 𝔤dR\mathfrak{g}_{\operatorname{dR}}^{\vee} has Hodge filtration concentrated in degrees [1,1][-1,1] given by

(𝔤dR/𝔭μ0)𝔭μ0,𝔤dR(\mathfrak{g}_{\operatorname{dR}}/\mathfrak{p}^{0}_{\mu})^{\vee}\subset\mathfrak{p}^{0,\vee}_{\mu}\subset\mathfrak{g}_{\operatorname{dR}}^{\vee}

such that

gri𝔤dR={(𝔤dR/𝔭μ0) if i=1,𝔪μ0, if i=0,𝔫μ0, if i=1.\operatorname{gr}^{i}\mathfrak{g}^{\vee}_{\operatorname{dR}}=\begin{cases}(\mathfrak{g}_{\operatorname{dR}}/\mathfrak{p}^{0}_{\mu})^{\vee}&\mbox{ if }i=1,\\ \mathfrak{m}^{0,\vee}_{\mu}&\mbox{ if }i=0,\\ \mathfrak{n}^{0,\vee}_{\mu}&\mbox{ if }i=-1.\end{cases}

Then, the flat connection :𝔤dR𝔤dR𝒪𝐆,μ,EΩ𝐆,μ,E1\nabla:\mathfrak{g}_{\operatorname{dR}}^{\vee}\to\mathfrak{g}_{\operatorname{dR}}^{\vee}\otimes_{\mathscr{O}_{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu,E^{\prime}}}}\Omega^{1}_{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu,E^{\prime}}} induces a map in gr1\operatorname{gr}^{1}-pieces

(𝔤dR/𝔭μ0)𝔪μ0,𝒪𝐆,μ,EΩ𝐆,μ,E1.(\mathfrak{g}_{\operatorname{dR}}/\mathfrak{p}^{0}_{\mu})^{\vee}\to\mathfrak{m}^{0,\vee}_{\mu}\otimes_{\mathscr{O}_{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu,E^{\prime}}}}\Omega^{1}_{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu,E^{\prime}}}.

Taking adjoints we get a 𝐆\mathbf{G}-equivariant map

KS~:(𝔤dR/𝔭μ0)𝒪FL𝐆,μ,E𝔪μ0Ω𝐆,μ1.\widetilde{\operatorname{KS}}:(\mathfrak{g}_{\operatorname{dR}}/\mathfrak{p}^{0}_{\mu})^{\vee}\otimes_{\mathscr{O}_{\operatorname{FL}_{\mathbf{G},\mu,E^{\prime}}}}\mathfrak{m}^{0}_{\mu}\to\Omega^{1}_{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu}}. (4.1)

Looking at the fiber at [1]𝐆,μ,E[1]\in\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu,E^{\prime}} the map (4.1) is nothing but the natural adjoint action of 𝔪μ\mathfrak{m}_{\mu} on (𝔤E/𝔭μ)(\mathfrak{g}_{E^{\prime}}/\mathfrak{p}_{\mu})^{\vee} with 𝔤E=𝔤pE\mathfrak{g}_{E^{\prime}}=\mathfrak{g}\otimes_{\mathbb{Q}_{p}}E^{\prime}:

(𝔤E/𝔭μ)E𝔪μad(𝔤E/𝔭μ)Ω𝐆,μ,E1|[1].(\mathfrak{g}_{E^{\prime}}/\mathfrak{p}_{\mu})^{\vee}\otimes_{E^{\prime}}\mathfrak{m}_{\mu}\xrightarrow{\operatorname{ad}}(\mathfrak{g}_{E^{\prime}}/\mathfrak{p}_{\mu})^{\vee}\cong\Omega^{1}_{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu,E^{\prime}}}|_{[1]}.

Therefore, the map KS~\widetilde{\operatorname{KS}} induces the 𝐆\mathbf{G}-equivariant Kodaira-Spencer isomorphism over the flag variety

KS:(𝔤dR/𝔭μ0)Ω𝐆,μ,E1.\operatorname{KS}:(\mathfrak{g}_{\operatorname{dR}}/\mathfrak{p}^{0}_{\mu})^{\vee}\xrightarrow{\sim}\Omega^{1}_{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu,E^{\prime}}}. (4.2)

which is the inverse of the dual of the anchor map α¯\overline{\alpha} of (2.8). Note that in particular KS\operatorname{KS} is already defined over EE as the anchor map is so, see Remark 2.5.1 (2).

We deduce the following proposition.

Proposition 4.1.1.

Let K𝐆(p)K\subset\mathbf{G}(\mathbb{Q}_{p}) be an open compact subgroup. The Kodaira-Spencer map of 𝐆,b,μ,K,E˘\mathcal{M}_{\mathbf{G},b,\mu,K,\breve{E}^{\prime}}

KS~:gr1(𝔤dR)𝒪gr0(𝔤dR)Ω1\widetilde{\operatorname{KS}}:\operatorname{gr}^{1}(\mathfrak{g}^{\vee}_{\operatorname{dR}})\otimes_{\mathscr{O}_{\mathcal{M}}}\operatorname{gr}^{0}(\mathfrak{g}_{\operatorname{dR}})\to\Omega^{1}_{\mathcal{M}}

constructed in analogue fashion as (4.1) factors through an isomorphism

KS:gr1(𝔤dR)Ω1\operatorname{KS}:\operatorname{gr}^{1}(\mathfrak{g}^{\vee}_{\operatorname{dR}})\xrightarrow{\sim}\Omega^{1}_{\mathcal{M}}

which is noting but the pullback along 𝐆,b,μ,K,E˘𝐆,μ,E˘\mathcal{M}_{\mathbf{G},b,\mu,K,\breve{E}^{\prime}}\to\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu,\breve{E}^{\prime}} of the Kodaira-Spencer isomorphism (4.2).

Proof.

This follows from the Kodaira-Spencer isomorphism (4.2) and the fact that the map 𝐆,b,μ,K𝐆,μ,E˘\mathcal{M}_{\mathbf{G},b,\mu,K}\to\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu,\breve{E}} is étale, namely, the filtered vector bundle with flat connection 𝔤dR\mathfrak{g}^{\vee}_{\operatorname{dR}} over 𝐆,b,μ,K\mathcal{M}_{\mathbf{G},b,\mu,K} is the pullback of the analogue filtered vector bundle with flat connection over the flag variety. ∎

We finish this section by rewriting the Kodaira-Spencer map in the form that will be used in the paper. Let K𝐆(p)K\subset\mathbf{G}(\mathbb{Q}_{p}) be a compact open subgroup. By Proposition 3.3.4 the local system 𝔤\mathcal{F}_{\mathfrak{g}^{\vee}} on the admissible locus 𝐆,μ,E˘a\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu,\breve{E}^{\prime}}^{a} (see Definition 3.3.3) is de Rham with associated filtered flat connection 𝔤dR\mathfrak{g}_{\operatorname{dR}}^{\vee}. Then, Corollaries 3.3.6 and 3.3.7 give rise 𝐆(p)×G~b\mathbf{G}(\mathbb{Q}_{p})\times\widetilde{G}_{b}-equivariant 𝒪^\widehat{\mathscr{O}}-sheaves on 𝐆,b,μ,,E˘\mathcal{M}_{\mathbf{G},b,\mu,\infty,\breve{E}^{\prime}}

πHT(𝔫μ10,)gr1(𝔤p𝒪^)gr1(𝔤dR)𝒪𝒪^(1)πGM((𝔤dR/𝔭μ))(1).\pi_{\operatorname{\scriptsize HT}}^{*}(\mathfrak{n}^{0,\vee}_{\mu^{-1}})\cong\operatorname{gr}_{1}(\mathcal{F}_{\mathfrak{g}^{\vee}}\otimes_{\mathbb{Q}_{p}}\widehat{\mathscr{O}})\cong\operatorname{gr}^{1}(\mathfrak{g}^{\vee}_{\operatorname{dR}})\otimes_{\mathscr{O}_{\mathcal{M}}}\widehat{\mathscr{O}}(-1)\cong\pi_{\operatorname{\scriptsize GM}}^{*}((\mathfrak{g}_{\operatorname{dR}}/\mathfrak{p}^{\vee}_{\mu})^{\vee})(-1). (4.3)

Composing (4.2) and (4.3) we get the following incarnation of the Kodaira-Spencer map

KS:πHT(𝔫μ10,)Ω1𝒪𝒪^(1)\operatorname{KS}^{\prime}:\pi_{\operatorname{\scriptsize HT}}^{*}(\mathfrak{n}^{0,\vee}_{\mu^{-1}})\xrightarrow{\sim}\Omega^{1}_{\mathcal{M}}\otimes_{\mathscr{O}_{\mathcal{M}}}\widehat{\mathscr{O}}(-1) (4.4)

as 𝐆(p)×G~b\mathbf{G}(\mathbb{Q}_{p})\times\widetilde{G}_{b}-equivariant 𝒪^\widehat{\mathscr{O}}-sheaves on 𝐆,b,μ,,E˘,v\mathcal{M}_{\mathbf{G},b,\mu,\infty,\breve{E}^{\prime},v}.

4.2. Pullbacks of equivariant vector bundles along πHT\pi_{\operatorname{\scriptsize HT}}

In this section we compute the Faltings extension of the local Shimura varieties in terms of the representation theory of the Hodge-Tate flag variety. For this computation we need to introduce some sheaves. We keep the representation theory notation of Section 2.5.

Let 𝒪𝔹dR+\mathscr{O}\!\mathbb{B}_{\operatorname{dR}}^{+} and 𝒪𝔹dR\mathscr{O}\!\mathbb{B}_{\operatorname{dR}} be the big proétale de Rham sheaves of 𝐆,μ,E˘a\operatorname{\mathcal{F}\ell}^{a}_{\mathbf{G},\mu,\breve{E}^{\prime}} as in [Sch13]. Let gr1𝒪𝔹dR+\operatorname{gr}^{1}\mathscr{O}\!\mathbb{B}_{\operatorname{dR}}^{+} be the Faltings extension, it is an 𝒪^\widehat{\mathscr{O}}-vector bundle in the proétale site and so it defines naturally a vv-vector bundle that we denote in the same way. We can write

gr0𝒪𝔹dR=Sym𝒪^(gr1𝒪𝔹dR+(1))/(1e(1))\operatorname{gr}^{0}\mathscr{O}\!\mathbb{B}_{\operatorname{dR}}=\mathrm{Sym}_{\widehat{\mathscr{O}}}(\operatorname{gr}^{1}\mathscr{O}\!\mathbb{B}_{\operatorname{dR}}^{+}(-1))/(1-e(1))

where 11 is the unit in the symmetric algebra, and where e:𝒪^gr1𝒪𝔹dR+(1)e:\widehat{\mathscr{O}}\to\operatorname{gr}^{1}\mathscr{O}\!\mathbb{B}_{\operatorname{dR}}^{+}(-1) is the natural map. Therefore, we can see gr0𝒪𝔹dR\operatorname{gr}^{0}\mathscr{O}\!\mathbb{B}_{\operatorname{dR}} as a vv-sheaf which is a filtered colimit of vv-vector bundles. See also [RJRC22, Remark 2.1.2].

Let 𝐍μ1𝐏μ1\mathbf{N}_{\mu^{-1}}\subset\mathbf{P}_{\mu^{-1}} be the unipotent radical of the opposite to the standard parabolic and let 𝒪(𝐍μ1)\mathscr{O}(\mathbf{N}_{\mu^{-1}}) be its space of algebraic functions. We endow 𝒪(𝐍μ1)\mathscr{O}(\mathbf{N}_{\mu^{-1}}) with the unique action of 𝐏μ=𝐍μ1𝐌μ1\mathbf{P}_{\mu}=\mathbf{N}_{\mu^{-1}}\rtimes\mathbf{M}_{\mu^{-1}} such that

  • The restriction to 𝐍μ1\mathbf{N}_{\mu^{-1}} is the left regular action, i.e

    (n1f)(n2)=f(n11n2)(n_{1}\cdot f)(n_{2})=f(n_{1}^{-1}n_{2})

    for n1,n2𝐍μ1n_{1},n_{2}\in\mathbf{N}_{\mu^{-1}} and f𝒪(𝐍μ1)f\in\mathscr{O}(\mathbf{N}_{\mu^{-1}}).

  • The restriction to 𝐌μ1\mathbf{M}_{\mu^{-1}} is the adjoint action, i.e.

    (mf)(n)=f(m1nm)(m\cdot f)(n)=f(m^{-1}nm)

    for m𝐌μ1m\in\mathbf{M}_{\mu^{-1}}, n𝐍μ1n\in\mathbf{N}_{\mu^{-1}} and f𝒪(𝐍μ1)f\in\mathscr{O}(\mathbf{N}_{\mu^{-1}}).

By [RC24b, Proposition 3.3.1] the algebra 𝒪(𝐍μ1)\mathscr{O}(\mathbf{N}_{\mu^{-1}}) has an increasing 𝐏μ1\mathbf{P}_{\mu^{-1}}-filtration 𝒪(𝐍μ1)n\mathscr{O}(\mathbf{N}_{\mu^{-1}})^{\leq n} with graded pieces grn(𝒪(𝐍μ1))SymEn𝔫μ1\operatorname{gr}_{n}(\mathscr{O}(\mathbf{N}_{\mu^{-1}}))\cong\mathrm{Sym}^{n}_{E^{\prime}}\mathfrak{n}^{\vee}_{\mu^{-1}}.

Recall that for an algebraic 𝐏μ1\mathbf{P}_{\mu^{-1}}-representation WW we let 𝒲𝐆,μ1(W)\mathcal{W}_{\mathbf{G},\mu^{-1}}(W) denote the 𝐆\mathbf{G}-equivariant quasi-coherent sheaf over 𝐆,μ1,E\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu^{-1},E^{\prime}} associated to WW via (2.6). We have the following theorem.

Theorem 4.2.1.

There is a natural 𝐆(p)×G~b\mathbf{G}(\mathbb{Q}_{p})\times\widetilde{G}_{b}-equivariant isomorphism of 𝒪^\widehat{\mathscr{O}}-algebras over 𝐆,b,μ,,E˘,v\mathcal{M}^{\lozenge}_{\mathbf{G},b,\mu,\infty,\breve{E}^{\prime},v}

gr0(𝒪𝔹dR)πHT(𝒲𝐆,μ1(𝒪(𝐍μ1))).\operatorname{gr}^{0}(\mathscr{O}\!\mathbb{B}_{\operatorname{dR}})\cong\pi^{*}_{\operatorname{\scriptsize HT}}(\mathcal{W}_{\mathbf{G},\mu^{-1}}(\mathscr{O}(\mathbf{N}_{\mu^{-1}}))).

More precisely, we have a 𝐆(p)×G~b\mathbf{G}(\mathbb{Q}_{p})\times\widetilde{G}_{b}-equivariant isomorphism of extensions

0{0}𝒪^{\widehat{\mathscr{O}}}πHT(𝒲𝐆,μ1(𝒪(𝐍μ1))){\pi^{*}_{\operatorname{\scriptsize HT}}(\mathcal{W}_{\mathbf{G},\mu^{-1}}(\mathscr{O}(\mathbf{N}_{\mu^{-1}})))}πHT(𝔫μ10,){\pi^{*}_{\operatorname{\scriptsize HT}}(\mathfrak{n}^{0,\vee}_{\mu^{-1}})}0{0}0{0}𝒪^{\widehat{\mathscr{O}}}gr1𝒪𝔹dR+(1){\operatorname{gr}^{1}\mathscr{O}\!\mathbb{B}_{\operatorname{dR}}^{+}(-1)}Ω1𝒪𝒪^(1){\Omega^{1}_{\mathcal{M}}\otimes_{\mathscr{O}_{\mathcal{M}}}\widehat{\mathscr{O}}(-1)}0{0}id\scriptstyle{\operatorname{id}}α\scriptstyle{\alpha}KS\scriptstyle{-\operatorname{KS}^{\prime}}

where KS\operatorname{KS}^{\prime} is the Kodaira-Spencer isomorphism of (4.4).

Proof.

The proof follows exactly the same lines of the proof of [RC24b, Theorem 5.1.4] where the key inputs are the Riemann-Hilbert correspondence of Proposition 3.3.4 and the Kodaira-Spencer isomorphism (4.4). Note that in loc. cit. we denoted 𝒪log,𝒮h=gr0(𝒪𝔹dR)\mathscr{O}\!\mathbb{C}_{\log,{\mathcal{S}h}}=\operatorname{gr}^{0}(\mathscr{O}\!\mathbb{B}_{\operatorname{dR}}), and we have identified 𝔤𝔤\mathfrak{g}\cong\mathfrak{g}^{\vee}, 𝔪μ1𝔪μ1\mathfrak{m}_{\mu^{-1}}\cong\mathfrak{m}_{\mu^{-1}}^{\vee} and 𝔫μ1𝔤E/𝔭μ1\mathfrak{n}_{\mu^{-1}}^{\vee}\cong\mathfrak{g}_{E^{\prime}}/\mathfrak{p}_{\mu^{-1}} via the Killing form of the derived Lie algebra of 𝔤\mathfrak{g} 333Strictly speaking one has to use the Killing form of the derived group 𝔤der\mathfrak{g}^{\mathrm{der}} to obtain the self duality.. ∎

4.3. Computation of the geometric Sen operators

We finish this section with the computation of the geometric Sen operators. By [RC23, Theorem 3.3.4], for any compact open subgroup K𝐆(p)K\subset\mathbf{G}(\mathbb{Q}_{p}), there is a natural geometric Sen operator

θ:𝔤p𝒪^Ω1𝒪𝒪^(1)\theta_{\mathcal{M}}:\mathcal{F}_{\mathfrak{g}^{\vee}}\otimes_{\mathbb{Q}_{p}}\widehat{\mathscr{O}}\to\Omega^{1}_{\mathcal{M}}\otimes_{\mathscr{O}_{\mathcal{M}}}\widehat{\mathscr{O}}(-1) (4.5)

seen as a morphism of 𝒪^\widehat{\mathscr{O}}-vector bundles over 𝐆,b,μ,K,E˘,v\mathcal{M}_{\mathbf{G},b,\mu,K,\breve{E}^{\prime},v}. We have the following theorem.

Theorem 4.3.1.

The geometric Sen operator (4.5) is the descent along the KK-torsor πK:𝐆,b,μ,,E˘𝐆,b,μ,K,E˘\pi_{K}:\mathcal{M}^{\lozenge}_{\mathbf{G},b,\mu,\infty,\breve{E}^{\prime}}\to\mathcal{M}^{\lozenge}_{\mathbf{G},b,\mu,K,\breve{E}^{\prime}} of the 𝐆(p)×G~b\mathbf{G}(\mathbb{Q}_{p})\times\widetilde{G}_{b}-equivariant map of 𝒪^\widehat{\mathscr{O}}-vector bundles obtained by pulling back along πHT\pi_{\operatorname{\scriptsize HT}} the map

𝔤μ10,𝔫μ10,\mathfrak{g}^{0,\vee}_{\mu^{-1}}\to\mathfrak{n}^{0,\vee}_{\mu^{-1}}

over 𝐆,μ1,E˘\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu^{-1},\breve{E}^{\prime}}, where πHT(𝔫μ10,)Ω1𝒪𝒪^(1)\pi_{\operatorname{\scriptsize HT}}^{*}(\mathfrak{n}^{0,\vee}_{\mu^{-1}})\cong\Omega^{1}_{\mathcal{M}}\otimes_{\mathscr{O}_{\mathcal{M}}}\widehat{\mathscr{O}}(-1) is the Kodaira-Spencer isomorphism (4.4).

Proof.

The proof is the same as the one of [RC24b, Theorem 5.2.5] where Theorem 4.2.1 replaces [RC24b, Theorem 5.1.4]. ∎

Corollary 4.3.2.

The geometric Sen operator (4.5) is a 𝐆(p)×G~b\mathbf{G}(\mathbb{Q}_{p})\times\widetilde{G}_{b}-equivariant map of 𝒪^\widehat{\mathscr{O}}-sheaves over 𝐆,b,μ,,E˘\mathcal{M}^{\lozenge}_{\mathbf{G},b,\mu,\infty,\breve{E}^{\prime}}.

Proof.

This follows from the fact that the pullback along πHT\pi_{\operatorname{\scriptsize HT}} of the map 𝔤μ10𝔫μ10,\mathfrak{g}^{0}_{\mu^{-1}}\to\mathfrak{n}^{0,\vee}_{\mu^{-1}} is 𝐆(p)×G~b\mathbf{G}(\mathbb{Q}_{p})\times\widetilde{G}_{b}-equivariant. ∎

We finish this section with the vanishing of higher locally analytic vectors for the sheaf 𝒪^\widehat{\mathscr{O}} and the computation of its arithmetic Sen operator. Let C/EC/E^{\prime} be the pp-adic completion of an algebraic closure. Let K𝐆(p)K\subset\mathbf{G}(\mathbb{Q}_{p}) be a compact open subgroup, and let 𝒱K=Cla(K,p)1\mathcal{V}_{K}=C^{la}(K,\mathbb{Q}_{p})_{\star_{1}} be the left regular locally analytic representation of KK. Consider the vv-sheaf 𝒱K\mathcal{F}_{\mathcal{V}_{K}} over 𝐆,b,μ,K,E˘\mathcal{M}_{\mathbf{G},b,\mu,K,\breve{E}} of Definition 3.3.3 which is a filtered colimit of Banach p\mathbb{Q}_{p}-linear vv-sheaves.

Theorem 4.3.3.

Let U𝐆,b,μ,K,CU\subset\mathcal{M}_{\mathbf{G},b,\mu,K,C} be an open affinoid subspace admitting an étale map to a product of tori 𝕋Cd\mathbb{T}^{d}_{C} that factors as a finite composition of rational localizations and finite étale maps. Let U𝐆,b,μ,,CU_{\infty}\subset\mathcal{M}_{\mathbf{G},b,\mu,\infty,C}^{\lozenge} be the pullback of UU, then

RΓv(U,𝒪^^p𝒱K)=𝒪^(U)GlaR\Gamma_{v}(U,\widehat{\mathscr{O}}\widehat{\otimes}_{\mathbb{Q}_{p}}\mathcal{F}_{\mathcal{V}_{K}})=\widehat{\mathscr{O}}(U_{\infty})^{G-la} (4.6)

sits in degree 0 and is equal to the locally analytic vectors of 𝒪^(U)\widehat{\mathscr{O}}(U_{\infty}). Here the completed tensor product is a filtered colimit of pp-completed tensor products obtained by writing 𝒱K\mathcal{F}_{\mathcal{V}_{K}} as a colimit of Banach sheaves, it coincides with the solid tensor product of [AM24, Section 4.1].

Furthermore, the action of 𝔤μ1,C0=𝒪𝐆,μ1,Cp𝔤\mathfrak{g}^{0}_{\mu^{-1},C}=\mathscr{O}_{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu^{-1},C}}\otimes_{\mathbb{Q}_{p}}\mathfrak{g} on 𝒪^(U)Gla\widehat{\mathscr{O}}(U_{\infty})^{G-la} by derivations kills 𝔫μ1,C0\mathfrak{n}^{0}_{\mu^{-1},C}. In particular we have an horizontal action of 𝔪μ1,C0\mathfrak{m}^{0}_{\mu^{-1},C} on 𝒪^(U)Gla\widehat{\mathscr{O}}(U_{\infty})^{G-la}. Moreover, the space 𝒪^(U)Gla\widehat{\mathscr{O}}(U_{\infty})^{G-la} has an arithmetic Sen operator as in [RC24b, Theorem 6.3.5] given by the opposite of the derivative of the Hodge cocharacter θμ=θμ1𝔪μ1,C0-\theta_{\mu}=\theta_{\mu^{-1}}\in\mathfrak{m}^{0}_{\mu^{-1},C}.

Proof.

The equivalence of (4.6) follows from the same proof of Proposition 6.2.8 (1) in [RC24b]. The vanishing of the action of 𝔫μ10\mathfrak{n}^{0}_{\mu^{-1}} on 𝒪^(U)Gla\widehat{\mathscr{O}}(U_{\infty})^{G-la} is Corollary 6.2.12 of loc. cit.. Finally, the existence and computation of the arithmetic Sen operator is Theorem 6.3.5 of loc. cit.. Note that in [RC24b] the statement of the theorem involves proétale cohomology and not vv-cohomology, these two are naturally the same thanks to the vv-decent results of [AM24, Theorem 5.6]. ∎

5. Locally analytic vectors at infinite level

In this last section we show the main results of this paper. First, in Section 5.1 we study the locally analytic vectors of period sheaves at infinite level local Shimura varieties. In particular, we prove that when bb is basic the locally analytic vectors are independent of the two towers of local Shimura varieties (5.1.9), generalizing a result of Pan for the Lubin-Tate tower [Pan22b, Corollary 5.3.9]. Then, in Section 5.2 we prove that, for bb basic, the colimit of compactly supported de Rham cohomologies as the level goes to 11 are independent of the two towers (Theorem 5.2.2), this result has also been independently obtained by Guido Bosco, Wiesława Nizioł and the first author. Finally, in Section 5.3 we prove that the pp-adic Jacquet-Langlands functor of Scholze [Sch18] for the Lubin-Tate tower is compatible with the passage to locally analytic vectors (Theorem 5.3.6).

5.1. Locally analytic vectors of towers of rigid spaces

Let us fix KK a perfectoid field in characteristic zero with tilt KK^{\flat}, we let ϖK\varpi\in K^{\flat} be a pseudo-uniformizer with |ϖ|K=|p|K|\varpi|_{K^{\flat}}=|p|_{K}. Throughout this section we suppose that KK contains all pp-th power roots of unit. For all bb rational we shall take pbKp^{b}\in K an element with |pb|K=|p|Kb|p^{b}|_{K}=|p|_{K}^{b} whenever it exists (similarly for ϖK\varpi\in K^{\flat}). Let GG and HH be two compact pp-adic Lie groups and XX a smooth qcqs rigid space over KK endowed with an action of HH. Suppose we are given with an HH-equivariant proétale GG-torsor X~X\widetilde{X}\to X^{\lozenge} seen as a diamond over Spd(K)\operatorname{Spd}(K), so that X~\widetilde{X} has a commuting action of G×HG\times H.

Let I=[s,r](0,)I=[s,r]\subset(0,\infty) be a compact interval with rational ends and let 𝔹I\mathbb{B}_{I} be the period sheaf on the vv-site XvX_{v}^{\lozenge} as in Definition 2.1.3. We can write 𝔹I=𝔹I+[1[ϖ]]\mathbb{B}_{I}=\mathbb{B}_{I}^{+}[\frac{1}{[\varpi]}] with 𝔹I+\mathbb{B}_{I}^{+} a [ϖ][\varpi]-adically complete sheaf. Set 𝔹I,b+=𝔹I+/[ϖ]b\mathbb{B}^{+}_{I,b}=\mathbb{B}^{+}_{I}/[\varpi]^{b}. Consider the solid p\mathbb{Q}_{p}-vector space RΓv(X~,𝔹I)R\Gamma_{v}(\widetilde{X},\mathbb{B}_{I}) with solid structure induced from the presentation

RΓv(X~,𝔹I)=RlimbRΓv(X~,𝔹I,b+)[1[ϖ]]R\Gamma_{v}(\widetilde{X},\mathbb{B}_{I})=R\varprojlim_{b}R\Gamma_{v}(\widetilde{X},\mathbb{B}_{I,b}^{+})[\frac{1}{[\varpi]}]

where RΓv(X~,𝔹I,b+)R\Gamma_{v}(\widetilde{X},\mathbb{B}_{I,b}^{+}) is a discrete 𝔹I,b+(K,K+,)\mathbb{B}_{I,b}^{+}(K^{\flat},K^{+,\flat})-complex, equal to the étale cohomology

RΓv(X~,𝔹I,b+)=RΓe´t(X~,𝔹I,b+)R\Gamma_{v}(\widetilde{X},\mathbb{B}_{I,b}^{+})=R\Gamma_{{\rm\acute{e}t}}(\widetilde{X},\mathbb{B}_{I,b}^{+})

by [Sch22, Proposition 14.7]. Equivalently, it is the pushforward along X~SpdK\widetilde{X}\to\operatorname{Spd}K of 𝔹I\mathbb{B}_{I} seen as a solid sheaf as in [AM24, §4].

The action of G×HG\times H on X~\widetilde{X} gives rise to the structure of an almost smooth G×HG\times H representation on RΓv(X~,𝔹I,b+)R\Gamma_{v}(\widetilde{X},\mathbb{B}_{I,b}^{+}) by Proposition 2.2.3 which then can be seen as an almost module over 𝔹[0,r],b+(K,K+,)[G×H]\mathbb{B}^{+}_{[0,r],b}(K^{\flat},K^{+,\flat})_{\square}[G\times H] via Remark 2.2.4. Then, after taking limits and colimits, the solid p\mathbb{Q}_{p}-vector space RΓv(X~,𝔹I)R\Gamma_{v}(\widetilde{X},\mathbb{B}_{I}) has a natural action of G×HG\times H, and so it gives rise an object in the derived category 𝒟(p,[G×H])\mathscr{D}(\mathbb{Q}_{p,\square}[G\times H]) of solid p\mathbb{Q}_{p}-linear G×HG\times H-representations (even an object in the derived \infty-category of semilinear solid representations of G×HG\times H over the Huber pair (𝔹I(K),𝔹I+(K+,))(\mathbb{B}_{I}(K^{\flat}),\mathbb{B}_{I}^{+}(K^{+,\flat})), namely the \infty-category 𝒟((𝔹I(K),𝔹I+(K+,))[G×H])\mathscr{D}((\mathbb{B}_{I}(K^{\flat}),\mathbb{B}_{I}^{+}(K^{+,\flat}))_{\square}[G\times H])). We want to prove the following theorem:

Theorem 5.1.1.

The natural map

RΓv(X~,𝔹I)R(G×H)laRΓv(X~,𝔹I)RGlaR\Gamma_{v}(\widetilde{X},\mathbb{B}_{I})^{R(G\times H)-la}\xrightarrow{\sim}R\Gamma_{v}(\widetilde{X},\mathbb{B}_{I})^{RG-la}

from G×HG\times H-locally analytic vectors to GG-locally analytic vectors is an equivalence.

Proof.

We can assume without loss of generality that both GG and HH are uniform pro-pp-groups. We shall consider almost mathematics with respect to ([ϖ]1/pn)n([\varpi]^{1/p^{n}})_{n}.

The strategy to prove Theorem 5.1.1 is to apply the locally analytic criterion of Lemma 2.4.1 for the group HH for suitable “lattices” of RΓv(X~,𝔹I)RGlaR\Gamma_{v}(\widetilde{X},\mathbb{B}_{I})^{RG-la}. We employ this strategy in different steps. We first make some formal reductions.

Lemma 5.1.2.

Suppose that Theorem 5.1.1 holds for smooth affinoid rigid spaces YY admitting toric coordinates ψ:Y𝕋Kd\psi:Y\to\mathbb{T}^{d}_{K}. Then it holds for general qcqs smooth rigid space XX.

Proof.

We first show that Theorem 5.1.1 holds for a quasi-compact and separated rigid space XX. Let {Xi}i=1k\{X_{i}\}_{i=1}^{k} be an affinoid cover of XX by subspaces admitting toric charts and let {XJ}J{1,,d}\{X_{J}\}_{J\subset\{1,\ldots,d\}} be the poset of finite intersections of the XiX_{i}. Any finite intersection XJX_{J} is then affinoid and admits a toric chart. For all JJ let us write X~J=XJ×XX~\widetilde{X}_{J}=X_{J}\times_{X}\widetilde{X}. Thus, we have that

RΓv(X~,𝔹I)=limJ{1,,d}RΓv(X~J,𝔹I).R\Gamma_{v}(\widetilde{X},\mathbb{B}_{I})=\varprojlim_{J\subset\{1,\ldots,d\}}R\Gamma_{v}(\widetilde{X}_{J},\mathbb{B}_{I}).

The claim follows since the functor of locally analytic vectors commutes with finite limits (being an exact functor of stable \infty-categories). Now, for general XX qcqs, we argue as before by taking a finite affinoid cover {Xi}i=1n\{X_{i}\}_{i=1}^{n}, and noticing that any finite intersection XJX_{J} is a quasi-compact separated rigid space. ∎

From now on we suppose that XX has toric coordinates ψ:X𝕋Kd\psi:X\to\mathbb{T}^{d}_{K}.

5.1.1. Modifying the locally analytic functions

First, we can write the left regular representation Cla(G,p)1=limhVhC^{la}(G,\mathbb{Q}_{p})_{\star_{1}}=\varinjlim_{h\to\infty}V_{h} as a colimit of analytic Banach representations of GG with Vh=Ch(G,p)1V_{h}=C^{h}(G,\mathbb{Q}_{p})_{\star_{1}} a suitable space of hh-analytic functions endowed with a left regular action. We can fix compatible p\mathbb{Z}_{p}-lattices Vh+VhV_{h}^{+}\subset V_{h} and define the pp-adically complete vv-sheaf Vh+\mathcal{F}_{V_{h}^{+}} over XvX_{v} obtained by proétale descent along the GG-torsor X~X\widetilde{X}\to X^{\lozenge}, see Definition 3.3.3. We set Vh=Vh+[1p]\mathcal{F}_{V_{h}}=\mathcal{F}_{V_{h}^{+}}[\frac{1}{p}].

Lemma 5.1.3.

There is a natural G×HG\times H-equivariant isomorphism of solid abelian groups

RΓv(X~,𝔹I)RGla=limhRΓv(X,𝔹I+^pVh+)[1[ϖ]],R\Gamma_{v}(\widetilde{X},\mathbb{B}_{I})^{RG-la}=\varinjlim_{h\to\infty}R\Gamma_{v}(X,\mathbb{B}^{+}_{I}\widehat{\otimes}_{\mathbb{Z}_{p}}\mathcal{F}_{V^{+}_{h}})[\frac{1}{[\varpi]}],

where the completed tensor in the RHS term is a pp-adically complete tensor product.

Proof.

Recall that both pp and [ϖ][\varpi] are pseudo-uniformizers in 𝔹I\mathbb{B}_{I}. Since X~\widetilde{X} is qcqs we can write

RΓv(X~,𝔹I)=RΓv(X~,𝔹I+)[1[ϖ]]R\Gamma_{v}(\widetilde{X},\mathbb{B}_{I})=R\Gamma_{v}(\widetilde{X},\mathbb{B}_{I}^{+})[\frac{1}{[\varpi]}]

Thus, since GG is compact, we get that

RΓv(X~,𝔹I)RGla\displaystyle R\Gamma_{v}(\widetilde{X},\mathbb{B}_{I})^{RG-la} =RΓ(G,RΓv(X~,𝔹I)p,LCla(G,p)1)\displaystyle=R\Gamma(G,R\Gamma_{v}(\widetilde{X},\mathbb{B}_{I})\otimes^{L}_{\mathbb{Q}_{p},\square}C^{la}(G,\mathbb{Q}_{p})_{\star_{1}})
=limhRΓ(G,RΓv(X~,𝔹I+)p,LVh+)[1[ϖ]]\displaystyle=\varinjlim_{h}R\Gamma(G,R\Gamma_{v}(\widetilde{X},\mathbb{B}_{I}^{+})\otimes^{L}_{\mathbb{Z}_{p},\square}V_{h}^{+})[\frac{1}{[\varpi]}]

where the spaces Cla(G,p)1C^{la}(G,\mathbb{Q}_{p})_{\star_{1}} and VhV_{h} have the left regular action of GG, the first equality is solid group cohomology as in [RJRC22, Definition 5.1 (1)], and in the second equality we use that the trivial representation is a compact p,[G]\mathbb{Z}_{p,\square}[G]-module thanks to the Lazard resolution which exists since GG is an uniform pro-pp-group.

Now, since both Vh+V_{h}^{+} and RΓv(X~,𝔹I+)R\Gamma_{v}(\widetilde{X},\mathbb{B}_{I}^{+}) are almost bounded to the right and derived pp-complete, we have by [Man22b, Proposition 2.12.10 (i)] that the solid tensor product RΓv(X~,𝔹I+)p,LVh+R\Gamma_{v}(\widetilde{X},\mathbb{B}_{I}^{+})\otimes^{L}_{\mathbb{Z}_{p},\square}V_{h}^{+} is derived pp-complete and almost equal to

Rlimk,s(RΓv(X~,𝔹I+/[ϖ]k)/psLVh+/ps)=Rlimk,s(RΓv(X~,𝔹I+/[ϖ]k/psLVh+/ps))R\varprojlim_{k,s}(R\Gamma_{v}(\widetilde{X},\mathbb{B}_{I}^{+}/[\varpi]^{k})\otimes^{L}_{\mathbb{Z}/p^{s}}V^{+}_{h}/p^{s})=R\varprojlim_{k,s}(R\Gamma_{v}(\widetilde{X},\mathbb{B}_{I}^{+}/[\varpi]^{k}\otimes^{L}_{\mathbb{Z}/p^{s}}\mathcal{F}_{V^{+}_{h}}/p^{s})) (5.1)

where in the second term we use that XX is qcqs and that V+/psV^{+}/p^{s} is a discrete /ps\mathbb{Z}/p^{s}-module. We get that

RΓv(X~,𝔹I)RGla\displaystyle R\Gamma_{v}(\widetilde{X},\mathbb{B}_{I})^{RG-la} =limh(Rlimk,sRΓ(G,RΓv(X~,𝔹I+/[ϖ]k/psLVh+/ps)))[1[ϖ]]\displaystyle=\varinjlim_{h}\left(R\varprojlim_{k,s}R\Gamma(G,R\Gamma_{v}(\widetilde{X},\mathbb{B}_{I}^{+}/[\varpi]^{k}\otimes^{L}_{\mathbb{Z}/p^{s}}\mathcal{F}_{V^{+}_{h}}/p^{s}))\right)[\frac{1}{[\varpi]}] (5.2)
=limh(Rlimk,sRΓsm(G,RΓv(X~,𝔹I+/[ϖ]k/psLVh+/ps)))[1[ϖ]]\displaystyle=\varinjlim_{h}\left(R\varprojlim_{k,s}R\Gamma^{sm}(G,R\Gamma_{v}(\widetilde{X},\mathbb{B}_{I}^{+}/[\varpi]^{k}\otimes^{L}_{\mathbb{Z}/p^{s}}\mathcal{F}_{V^{+}_{h}}/p^{s}))\right)[\frac{1}{[\varpi]}]
=limh(Rlimk,sRΓv(X,𝔹I+/[ϖ]k/psVh+/ps))[1[ϖ]]\displaystyle=\varinjlim_{h}\left(R\varprojlim_{k,s}R\Gamma_{v}(X,\mathbb{B}^{+}_{I}/[\varpi]^{k}\otimes_{\mathbb{Z}/p^{s}}\mathcal{F}_{V^{+}_{h}}/p^{s})\right)[\frac{1}{[\varpi]}]
=limhRΓv(X,𝔹I+^pVh+)[1[ϖ]]\displaystyle=\varinjlim_{h}R\Gamma_{v}(X,\mathbb{B}^{+}_{I}\widehat{\otimes}_{\mathbb{Z}_{p}}\mathcal{F}_{V^{+}_{h}})[\frac{1}{[\varpi]}]

where in the first equality we use (5.1), the fact group cohomology for GG commutes with limits, and that is also commutes with colimits since GG is uniform pro-pp. The second equality we use Proposition 2.2.3 and [Man22b, Remark 3.4.12] to compare solid and smooth cohomology. In the third equality we also use Proposition 2.2.3 and the fact that vv-cohomology is the pushforward along the map X=X~/G(SpdK)/GSpdKX^{\lozenge}=\widetilde{X}/G\to(\operatorname{Spd}K)/G\to\operatorname{Spd}K. In the last equality we use that cohomology commutes with derived limits of sheaves and that 𝔹I+^pVh+\mathbb{B}^{+}_{I}\widehat{\otimes}_{\mathbb{Z}_{p}}\mathcal{F}_{V^{+}_{h}} is derived ([ϖ],p)([\varpi],p)-complete. This proves the lemma. ∎

The objects RΓv(X,𝔹I+^pVh+)R\Gamma_{v}(X,\mathbb{B}_{I}^{+}\widehat{\otimes}_{\mathbb{Z}_{p}}\mathcal{F}_{V_{h}^{+}}) are ϖ\varpi-adically complete solid G×HG\times H-representations over 𝔹I+(K)\mathbb{B}_{I}^{+}(K), where the HH-action arises from the action on XX and the GG-action is induced by the right regular action on Vh+V_{h}^{+}. To prove Theorem 5.1.1 it suffices to show the following:

Dévisage 1.

The solid HH-representation

RΓv(X,𝔹I+^pVh+)[1[ϖ]]R\Gamma_{v}(X,\mathbb{B}_{I}^{+}\widehat{\otimes}_{\mathbb{Z}_{p}}\mathcal{F}_{V_{h}^{+}})[\frac{1}{[\varpi]}] (5.3)

is locally analytic.

To prove 1, we need to modify a little bit more the lattices RΓv(X,𝔹I+^pVh+)R\Gamma_{v}(X,\mathbb{B}_{I}^{+}\widehat{\otimes}_{\mathbb{Z}_{p}}\mathcal{F}_{V_{h}^{+}}).

5.1.2. Modifying the sheaf of periods

Let us write I=[s,r](0,)I=[s,r]\subset(0,\infty).Consider the [ϖ][\varpi]-adically complete complex

𝔹~I+:=[𝔹[0,r]+TpT[ϖ]1/s𝔹[0,r]+T]\widetilde{\mathbb{B}}_{I}^{+}:=[\mathbb{B}_{[0,r]}^{+}\langle T\rangle\xrightarrow{pT-[\varpi]^{1/s}}\mathbb{B}_{[0,r]}^{+}\langle T\rangle] (5.4)

and take the lattice of (5.3) given by

RΓv(X,𝔹~I+^pLVh+).R\Gamma_{v}(X,\widetilde{\mathbb{B}}_{I}^{+}\widehat{\otimes}^{L}_{\mathbb{Z}_{p}}\mathcal{F}_{V_{h}^{+}}). (5.5)

5.1.3. Modifying the level

For G0GG_{0}\subset G an open compact normal subgroup let XG0=X~/G0X_{G_{0}}=\widetilde{X}/G_{0}. Then

RΓv(X,𝔹~I+^pLVh+)[1[ϖ]]R\Gamma_{v}(X,\widetilde{\mathbb{B}}_{I}^{+}\widehat{\otimes}^{L}_{\mathbb{Z}_{p}}\mathcal{F}_{V_{h}^{+}})[\frac{1}{[\varpi]}]

is just a finite colimit (given by the invariants of G/G0G/G_{0}) of

RΓv(XG0,𝔹~I+^pLVh+)[1[ϖ]].R\Gamma_{v}(X_{G_{0}},\widetilde{\mathbb{B}}_{I}^{+}\widehat{\otimes}^{L}_{\mathbb{Z}_{p}}\mathcal{F}_{V_{h}^{+}})[\frac{1}{[\varpi]}].

Since the category of solid locally analytic representations is stable under colimits by [RJRC23, Proposition 3.2.3], to show 1 it suffices to prove the following statement:

Dévisage 2.

There is an open compact subgroup G0GG_{0}\subset G such that

RΓv(XG0,𝔹~I+^pLVh+)[1[ϖ]]R\Gamma_{v}(X_{G_{0}},\widetilde{\mathbb{B}}_{I}^{+}\widehat{\otimes}^{L}_{\mathbb{Z}_{p}}\mathcal{F}_{V_{h}^{+}})[\frac{1}{[\varpi]}] (5.6)

is a locally analytic representation of HH.

We shall take G0G_{0} such that the G0G_{0}-module Vh+/pbV^{+}_{h}/p^{b} is isomorphic to the trivial representation I/pb\bigoplus_{I}\mathbb{Z}/p^{b} for some fixed bb that we shall choose in Section 5.1.5.

5.1.4. Applying the decalage functor

Finally, it is well known that the proétale cohomology of 𝒪^+\widehat{\mathscr{O}}^{+} has some junk torsion (eg. see [BMS18]), this makes difficult to apply the analyticity criterion of Lemma 2.4.1 to the lattice of 2. A way to solve this problem is to modify the lattice a little bit by applying a décalage functor. First, we need to guarantee that the décalage functor preserves the structure of a solid representation, for this it suffices to see that the category of solid representations of a profinite group Π\Pi can be obtained as the derived category of abelian sheaves on a ringed topos. This is a consequence of the next lemma:

Lemma 5.1.4 ([Man22a, Lemma 10.3]).

Let Π\Pi be a profinite group, and Λ\Lambda a nuclear p\mathbb{Z}_{p}-algebra. Then there is a natural equivalence of \infty-categories

𝒟(/Π,Λ)𝒟(Λ,p)BG=𝒟(Λ[G])\mathscr{D}_{\square}(*/\Pi,\Lambda)\cong\mathscr{D}_{\square}(\Lambda,\mathbb{Z}_{p})^{BG}=\mathscr{D}(\Lambda_{\square}[G])

between solid sheaves on the proétale site of the vv-stack /Π*/\Pi with p\mathbb{Z}_{p}-coefficients, and the category of Λ\Lambda-linear solid representations of Π\Pi.

Proof.

The result in loc. cit. is only stated for p\ell\neq p, let us see that this is actually not necessary. Indeed, the proétale site of /G*/G is independent of the prime pp used in the definition of vv-stacks where /G*/G is considered. So we could have taken /G*/G as an object in vv-stacks for perfectoid spaces in characteristic p\ell\neq p and still get the same conclusion. ∎

By taking Λ=𝔹I+(K)\Lambda=\mathbb{B}_{I}^{+}(K^{\flat}) we can apply the décalage functor Lη[ϖ]εL\eta_{[\varpi]^{\varepsilon}} for any rational ε\varepsilon to Λ\Lambda-linear solid representations of Π\Pi.

For ε>0\varepsilon>0 rational, to be determined in Section 5.1.5, we shall consider the following lattice of (5.6)

Lη[ϖ]εRΓv(XG0,𝔹~I+^pLVh+).L\eta_{[\varpi]^{\varepsilon}}R\Gamma_{v}(X_{G_{0}},\widetilde{\mathbb{B}}_{I}^{+}\widehat{\otimes}^{L}_{\mathbb{Z}_{p}}\mathcal{F}_{V_{h}^{+}}). (5.7)

In this way, to show 2 it suffices to prove the following:

Dévisage 3.

There exits G0G_{0} and ε>0\varepsilon>0 such that the [ϖ][\varpi]-complete lattice

Lη[ϖ]εRΓv(XG0,𝔹~I+^pLVh+)L\eta_{[\varpi]^{\varepsilon}}R\Gamma_{v}(X_{G_{0}},\widetilde{\mathbb{B}}_{I}^{+}\widehat{\otimes}^{L}_{\mathbb{Z}_{p}}\mathcal{F}_{V_{h}^{+}})

satisfies the hypothesis of Lemma 2.4.1 for the action of HH.

5.1.5. Reduction to 𝒪X+/p1/r\mathscr{O}^{+}_{X}/p^{1/r}

We finally perform the last dévisage in the proof of Theorem 5.1.1. Let us write I=[s,r](0,)I=[s,r]\subset(0,\infty) so that |[ϖ]|1/s|p|[ϖ]1/r|[\varpi]|^{1/s}\leq|p|\leq[\varpi]^{1/r}. Thus, we shall make the following choices:

  • i.

    We take G0G_{0} such that the action of G0G_{0} on Vh+p𝔹[0,r]+(K)/[ϖ]1/sV_{h}^{+}\otimes_{\mathbb{Z}_{p}}\mathbb{B}^{+}_{[0,r]}(K^{\flat})/[\varpi]^{1/s} is trivial. In particular, as Vh+V_{h}^{+} is a torsion free pp-adically complete p\mathbb{Z}_{p}-module, the G0G_{0}-representation Vh+p𝔹[0,r]+(K)/[ϖ]1/sV_{h}^{+}\otimes_{\mathbb{Z}_{p}}\mathbb{B}^{+}_{[0,r]}(K^{\flat})/[\varpi]^{1/s} is isomorphic to a direct sum of copies of 𝔹[0,r]+(K)/[ϖ]1/s\mathbb{B}^{+}_{[0,r]}(K^{\flat})/[\varpi]^{1/s}.

  • ii.

    We take ε<1/r\varepsilon<1/r.

Step 1. We first have to guarantee that Lη[ϖ]εRΓv(XG0,𝔹~I+^pLVh+)L\eta_{[\varpi]^{\varepsilon}}R\Gamma_{v}(X_{G_{0}},\widetilde{\mathbb{B}}_{I}^{+}\widehat{\otimes}^{L}_{\mathbb{Z}_{p}}\mathcal{F}_{V_{h}^{+}}) is [ϖ][\varpi]-adically complete and bounded to the right. The first claim follows from [BMS18, Lemma 6.20] and the fact that RΓv(XG0,𝔹~I+^pLVh+)R\Gamma_{v}(X_{G_{0}},\widetilde{\mathbb{B}}_{I}^{+}\widehat{\otimes}^{L}_{\mathbb{Z}_{p}}\mathcal{F}_{V_{h}^{+}}) is ϖ\varpi-adically complete. To see that it is bounded to the right, by [ϖ][\varpi]-adically completeness and since the décalage functor kills [ϖ]ε[\varpi]^{\varepsilon}-torsion, it suffices to see that

RΓv(XG0,𝔹~I+^pLVh+)/𝕃[ϖ]1/rR\Gamma_{v}(X_{G_{0}},\widetilde{\mathbb{B}}_{I}^{+}\widehat{\otimes}^{L}_{\mathbb{Z}_{p}}\mathcal{F}_{V_{h}^{+}})/^{\mathbb{L}}[\varpi]^{1/r}

is almost bounded to the right. By (5.4), the complex 𝔹~I+\widetilde{\mathbb{B}}^{+}_{I} is constructed with terms given by 𝔹[0,r]+T\mathbb{B}_{[0,r]}^{+}\langle T\rangle. Thus, it suffices to show that

RΓv(XG0,(𝔹[0,r]+/[ϖ]1/r)pLVh+)R\Gamma_{v}(X_{G_{0}},(\mathbb{B}_{[0,r]}^{+}/[\varpi]^{1/r})\otimes^{L}_{\mathbb{Z}_{p}}\mathcal{F}_{V_{h}^{+}})

is bounded to the right. By the choice of G0G_{0}, we know that 𝔹[0,r]+/[ϖ]1/r)LpVh+\mathbb{B}_{[0,r]}^{+}/[\varpi]^{1/r})\otimes^{L}_{\mathbb{Z}_{p}}\mathcal{F}_{V_{h}^{+}} is isomorphic to a direct sum of copies of 𝔹[0,r]+/[ϖ]1/r\mathbb{B}_{[0,r]}^{+}/[\varpi]^{1/r} which by Lemma 2.1.4 (2) is almost isomorphic to a polynomial algebra

𝔹[0,r]+/[ϖ]1/ra𝒪X+/ϖ1/r[S].\mathbb{B}_{[0,r]}^{+}/[\varpi]^{1/r}\cong^{a}\mathscr{O}_{X}^{+}/\varpi^{1/r}[S]. (5.8)

Hence, we are reduced to see that

RΓv(XG0,𝒪X+/p1/r)R\Gamma_{v}(X_{G_{0}},\mathscr{O}_{X}^{+}/p^{1/r})

is almost bounded to the right, which is clear as XG0X_{G_{0}} is an affinoid smooth rigid space.

Step 2. Since [ϖ][\varpi] and pp are both pseudo-uniformizers of 𝔹I\mathbb{B}_{I}, if the hypothesis of Lemma 2.4.1 holds for a power [ϖ]δ[\varpi]^{\delta} of [ϖ][\varpi] then, after base change by a sufficiently ramified extension KK of p\mathbb{Q}_{p}, the hypothesis will hold for the pseudo-uniformizer of π\pi of KK. Indeed, we just need to pick π\pi such that |π|=|p|δr|[ϖ]|δ|\pi|=|p|^{\delta r}\leq|[\varpi]|^{\delta}.

Step 3. We will show that there is an open compact subgroup H0HH_{0}\subset H such that for all hH0h\in H_{0} the map 1h1-h on

Lη[ϖ]εRΓv(XG0,𝔹~I+^pLVh+)/𝕃[ϖ]1/rεL\eta_{[\varpi]^{\varepsilon}}R\Gamma_{v}(X_{G_{0}},\widetilde{\mathbb{B}}_{I}^{+}\widehat{\otimes}^{L}_{\mathbb{Z}_{p}}\mathcal{F}_{V_{h}^{+}})/^{\mathbb{L}}[\varpi]^{1/r-\varepsilon} (5.9)

is homotopic to zero as 𝔹[0,r]+(K,K+,)/[ϖ]1/rε\mathbb{B}^{+}_{[0,r]}(K^{\flat},K^{+,\flat})/[\varpi]^{1/r-\varepsilon}-module.

By Steps 1 and 2 and Lemma 2.4.1 we will obtain that the HH-representation (5.6) is locally analytic proving Theorem 5.1.1.

By Lemma 2.3.3 the object (5.9) is equivalent to

Lη[ϖ]εRΓv(XG0,(𝔹~I+/𝕃[ϖ]1/r)pLVh+).L\eta_{[\varpi]^{\varepsilon}}R\Gamma_{v}(X_{G_{0}},(\widetilde{\mathbb{B}}_{I}^{+}/^{\mathbb{L}}[\varpi]^{1/r})\otimes^{L}_{\mathbb{Z}_{p}}\mathcal{F}_{V_{h}^{+}}). (5.10)

By the choice of G0G_{0}, the proétale sheaf (𝔹~I+/𝕃[ϖ]1/r)pLVh+(\widetilde{\mathbb{B}}_{I}^{+}/^{\mathbb{L}}[\varpi]^{1/r})\otimes^{L}_{\mathbb{Z}_{p}}\mathcal{F}_{V_{h}^{+}} is isomorphic to a direct sum of copies of 𝔹~I+/𝕃[ϖ]1/r\widetilde{\mathbb{B}}_{I}^{+}/^{\mathbb{L}}[\varpi]^{1/r}. Therefore, since the décalage functor commutes with direct sums [BMS18, Corollary 6.5], it suffices to show that for all hH0h\in H_{0} the operator 1h1-h is homotopically equivalent to zero when acting on

Lη[ϖ]εRΓv(XG0,(𝔹~I+/𝕃[ϖ]1/r)).L\eta_{[\varpi]^{\varepsilon}}R\Gamma_{v}(X_{G_{0}},(\widetilde{\mathbb{B}}_{I}^{+}/^{\mathbb{L}}[\varpi]^{1/r})).

By the definition of 𝔹~I+\widetilde{\mathbb{B}}^{+}_{I} in (5.4), and since pp and [ϖ]1/r[\varpi]^{1/r} are divisible by [ϖ]1/r[\varpi]^{1/r} in 𝔹[0,r]+\mathbb{B}_{[0,r]}^{+}, we have that

𝔹~I+/𝕃[ϖ]1/r=𝔹[0,r]+/[ϖ]1/r[T][1]𝔹[0,r]+/[ϖ]1/r[T].\widetilde{\mathbb{B}}_{I}^{+}/^{\mathbb{L}}[\varpi]^{1/r}=\mathbb{B}^{+}_{[0,r]}/[\varpi]^{1/r}[T][1]\oplus\mathbb{B}^{+}_{[0,r]}/[\varpi]^{1/r}[T].

Finally, by (5.8) and since Lη[ϖ]εL\eta_{[\varpi]^{\varepsilon}} preserves shifts and direct sums, we are reduced to show the following dévisage:

Dévisage 4.

There is an open compact subgroup H0HH_{0}\subset H such that for all hH0h\in H_{0} the operator 1h1-h on

LηpεRΓv(XG0,𝒪X+/p1/r)L\eta_{p^{\varepsilon}}R\Gamma_{v}(X_{G_{0}},\mathscr{O}_{X}^{+}/p^{1/r})

is homotopic to zero as 𝒪K/p1/r\mathcal{O}_{K}/p^{1/r}-module.

5.1.6. Final computation

We now prove 4. We can assume without loss of generality that G0=GG_{0}=G and so XG0=XX_{G_{0}}=X. By Lemma 5.1.2 we can also assume that XX has toric coordinates Ψ:X𝕋Kd\Psi:X\to\mathbb{T}^{d}_{K}. Let 𝕋K,d\mathbb{T}^{d}_{K,\infty} be the perfectoid torus and let Γ=p(1)d\Gamma=\mathbb{Z}_{p}(1)^{d} the Galois group of 𝕋K,d\mathbb{T}^{d}_{K,\infty} over 𝕋Kd\mathbb{T}^{d}_{K}. For nn\in\mathbb{N} we let 𝕋K,nd=𝕋K,d/Γpn\mathbb{T}^{d}_{K,n}=\mathbb{T}^{d}_{K,\infty}/\Gamma^{p^{n}} and Xn=X×𝕋Kd𝕋K,ndX_{n}=X\times_{\mathbb{T}^{d}_{K}}\mathbb{T}^{d}_{K,n}, similarly we let X=X×𝕋Kd𝕋K,dX_{\infty}=X\times_{\mathbb{T}^{d}_{K}}\mathbb{T}^{d}_{K,\infty}. By [RC23, Proposition 3.2.3] the pair (𝒪(X),Γ)(\mathscr{O}(X_{\infty}),\Gamma) is a strongly decomposable Sen theory in the sense of [RC23, Definition 2.2.6]. In particular, we have Sen traces Trn:𝒪(X)𝒪(Xn)\mathrm{Tr}_{n}:\mathscr{O}(X_{\infty})\to\mathscr{O}(X_{n}) with kernel CnC_{n}. Thus, by letting Cn+=Cn𝒪(Xn)+C_{n}^{+}=C_{n}\cap\mathscr{O}(X_{n})^{+}, there is some n>>0n>>0 such that the cokernel of the map

Cn+𝒪+(Xn)𝒪+(X)C_{n}^{+}\oplus\mathscr{O}^{+}(X_{n})\to\mathscr{O}^{+}(X_{\infty})

as well as the group cohomology RΓ(Γ,Cn+)R\Gamma(\Gamma,C_{n}^{+}) are killed by pεp^{\varepsilon}. Since we have an almost equivalence

RΓ(Γ,𝒪+(X)/p1/r)aRΓv(X,𝒪X+/p1/r),R\Gamma(\Gamma,\mathscr{O}^{+}(X_{\infty})/p^{1/r})\cong^{a}R\Gamma_{v}(X,\mathscr{O}_{X}^{+}/p^{1/r}),

there exists some n>>0n>>0 depending on ε\varepsilon such that we have an equivalence

Lηpε(R(Γ,𝒪+(Xn)/p1/r))LηpεRΓv(X,𝒪X+/p1/r)L\eta_{p^{\varepsilon}}(R(\Gamma,\mathscr{O}^{+}(X_{n})/p^{1/r}))\cong L\eta_{p^{\varepsilon}}R\Gamma_{v}(X,\mathscr{O}_{X}^{+}/p^{1/r}) (5.11)

of 𝒪K/p1/r\mathcal{O}_{K}/p^{1/r}-complexes. Let us now justify that (5.11) can be promoted to an equivalence of smooth H0H_{0}-representations for some H0HH_{0}\subset H small enough. Indeed, by [Sch18, Lemma 2.3] there is an open subgroup H0HH_{0}\subset H such that the action of H0H_{0} can be lifted from XX to an action σ\sigma on XnX_{n}. Moreover, since for all γΓ/Γpn\gamma\in\Gamma/\Gamma^{p_{n}} the conjugation γ1σγ\gamma^{-1}\circ\sigma\circ\gamma is another lift of the action of H0H_{0} to XnX_{n}, by refining H0H_{0} and using [Sch18, Lemma 2.3] again we can suppose that both the actions of H0H_{0} and Γ/Γpn\Gamma/\Gamma^{p^{n}} on XnX_{n} commute. This shows that the map (5.11) can be upgraded to a map of smooth H0H_{0}-representations as wanted.

But then, by Corollary 2.4.2, we can shrink H0H_{0} so that it acts trivially on 𝒪+(Xn)/p1/r\mathscr{O}^{+}(X_{n})/p^{1/r}. Thus, for hH0h\in H_{0} the action of 1h1-h on the left hand side of (5.11) is homotopic to zero as 𝒪K/p1/r\mathcal{O}_{K}/p^{1/r}-module finishing the proof of Theorem 5.1.1.

Remark 5.1.5.

Theorem 5.1.1 will hold for a much larger class of vv-sheaves or complexes \mathscr{F} following essentially the same proof. For example it holds under the following conditions which hold for 𝔹I\mathbb{B}_{I} and 𝒪^\widehat{\mathscr{O}}-vector bundles:

  • (a)

    The complex \mathscr{F} is of the form =[1p]\mathscr{F}=\mathscr{F}^{\circ}[\frac{1}{p}] with \mathscr{F}^{\circ} a connective derived pp-complete sheaf such that /𝕃p\mathscr{F}^{\circ}/^{\mathbb{L}}p arises from the étale site of XX (in particular \mathscr{F} is a solid sheaf as in [AM24, §4]).

  • (b)

    There is some b>0b>0 such that /𝕃pb\mathscr{F}^{\circ}/^{\mathbb{L}}p^{b} is a retract of a vv-sheaf of the form 𝒪^+/pbpLV\widehat{\mathscr{O}}^{+}/p^{b}\otimes_{\mathbb{Z}_{p}}^{L}V with VV a complex of p\mathbb{Z}_{p}-modules.

Indeed, one has to prove 3 for the cohomology LηpεRΓv(XG0,+)L\eta_{p^{\varepsilon}}R\Gamma_{v}(X_{G_{0}},\mathscr{F}^{+}). But since XG0X_{G_{0}} is qcqs and the décalage operator sends retracts to retracts, the criterion of Lemma 2.4.1 holds for this lattice if it does for

LηpεRΓv(XG0,𝒪^+/pbpLV)=Lηpε[RΓv(XG0,𝒪^+/pb)pLV]L\eta_{p^{\varepsilon}}R\Gamma_{v}(X_{G_{0}},\widehat{\mathscr{O}}^{+}/p^{b}\otimes_{\mathbb{Z}_{p}}^{L}V)=L\eta_{p^{\varepsilon}}[R\Gamma_{v}(X_{G_{0}},\widehat{\mathscr{O}}^{+}/p^{b})\otimes_{\mathbb{Z}_{p}}^{L}V]

Using the symmetric monoidality of the décalage operator [BMS18, Proposition 6.8] one is reduced to 4.

An immediate corollary of Theorem 5.1.1 is that the cohomology groups of qcqs smooth rigid spaces endowed with an action of a pp-adic Lie group is locally analytic.

Corollary 5.1.6.

Let XX be a qcqs smooth rigid space over a perfectoid field KK admitting all pp-th power roots of unit. Suppose that XX is endowed with a continuous action of a pp-adic Lie group HH. Then for any compact interval I(0,)I\subset(0,\infty) the vv-cohomology

RΓv(X,𝔹I)R\Gamma_{v}(X,\mathbb{B}_{I})

is a solid locally analytic HH-representation.

Proof.

This is a particular case of Theorem 5.1.1 where G=1G=1. ∎

5.1.7. Conclusion for local Shimura varieties

In this paragraph we apply Theorem 5.1.1 to local Shimura varieties. Recall that (𝐆,b,μ)(\mathbf{G},b,\mu) denotes a local Shimura datum, and for K𝐆(p)K\subset\mathbf{G}(\mathbb{Q}_{p}) a compact open subgroup we have the local Shimura variety 𝐆,b,μ,K\mathcal{M}_{\mathbf{G},b,\mu,K} of level KK. We also let 𝐆,b,μ,\mathcal{M}^{\lozenge}_{\mathbf{G},b,\mu,\infty} be the infinite level local Shimura variety seen as a diamond over SpdE˘\operatorname{Spd}\breve{E}. We fix C/E˘C/\breve{E} be a complete algebraically closed extension, and consider the base change of local Shimura varieties to CC.

Definition 5.1.7.

We let 𝒪^\widehat{\mathscr{O}}_{\mathcal{M}} be the restriction of the vv-sheaf 𝒪^\widehat{\mathscr{O}} to the topological space |𝐆,b,μ,,C||\mathcal{M}_{\mathbf{G},b,\mu,\infty,C}| and let 𝒪Gla𝒪^\mathscr{O}^{G-la}_{\mathcal{M}}\subset\widehat{\mathscr{O}}_{\mathcal{M}} be the subsheaf mapping a qcqs open subspace U𝐆,b,μ,,pU_{\infty}\subset\mathcal{M}^{\lozenge}_{\mathbf{G},b,\mu,\infty,\mathbb{C}_{p}} to the space of G=𝐆(p)G=\mathbf{G}(\mathbb{Q}_{p})-locally analytic vectors

𝒪Gla(U~)=𝒪^(U)KUla\mathscr{O}^{G-la}_{\mathcal{M}}(\widetilde{U}_{\infty})=\widehat{\mathscr{O}}_{\mathcal{M}}(U_{\infty})^{K_{U_{\infty}}-la}

where KU𝐆(p)K_{U_{\infty}}\subset\mathbf{G}(\mathbb{Q}_{p}) is the stabilizer of UU_{\infty}. If GG is clear from the context we write 𝒪la\mathscr{O}^{la}_{\mathcal{M}} instead of 𝒪Gla\mathscr{O}^{G-la}_{\mathcal{M}}.

Remark 5.1.8.

The fact that 𝒪la\mathscr{O}^{la}_{\mathcal{M}} is a sheaf follows from [RC24b, Lemma 6.2.2].

We obtain a generalization of a theorem of Lue Pan for the Lubin-Tate space, see [Pan22b, Corollary 5.3.9].

Corollary 5.1.9.

For any pp-adic Lie group HG~bH\subset\widetilde{G}_{b} and any qcqs open subspace U𝐆,b,μ,,CU_{\infty}\subset\mathcal{M}^{\lozenge}_{\mathbf{G},b,\mu,\infty,C} the natural map

𝒪Gla(U)RHla𝒪Gla(U)\mathscr{O}^{G-la}_{\mathcal{M}}(U_{\infty})^{RH-la}\xrightarrow{\sim}\mathscr{O}^{G-la}_{\mathcal{M}}(U_{\infty})

is an equivalence. In particular, if bb is basic we have an equality of subsheaves of 𝒪^\widehat{\mathscr{O}}_{\mathcal{M}}

𝒪Gla=𝒪Gbla.\mathscr{O}^{G-la}_{\mathcal{M}}=\mathscr{O}^{G_{b}-la}_{\mathcal{M}}.

More generally, for bb basic and I(0,)I\subset(0,\infty) a compact interval, we have an equivalence of derived solid locally analytic representations of G×GbG\times G_{b}

RΓv(U,𝔹I)RGblaRΓv(U,𝔹I)RG×GblaRΓv(U,𝔹I)RGla.R\Gamma_{v}(U_{\infty},\mathbb{B}_{I})^{RG_{b}-la}\xrightarrow{\sim}R\Gamma_{v}(U_{\infty},\mathbb{B}_{I})^{RG\times G_{b}-la}\xleftarrow{\sim}R\Gamma_{v}(U_{\infty},\mathbb{B}_{I})^{RG-la}.
Proof.

The first claim follows from Theorem 5.1.1 and Remark 5.1.5, and the fact that 𝒪^\widehat{\mathscr{O}}_{\mathcal{M}} has no higher locally analytic vectors by Theorem 4.3.3. The claim when bb is basic follows also from Theorem 5.1.1 and the fact that for KbGbK_{b}\subset G_{b} a compact open subgroup, the quotient 𝐆,b,μ,/KbGˇ,bˇ,μˇ,Kb\mathcal{M}^{\lozenge}_{\mathbf{G},b,\mu,\infty}/K_{b}\cong\mathcal{M}^{\lozenge}_{\check{G},\check{b},\check{\mu},K_{b}} is the diamond attached to a local Shimura variety of level KbK_{b} for the dual Shimura datum (𝐆ˇ,bˇ,μˇ)(\check{\mathbf{G}},\check{b},\check{\mu}). ∎

In the following we shall write 𝐆,μ\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu} and 𝐆,μ1\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu^{-1}} for the base change to (C,𝒪C)(C,\mathcal{O}_{C}) of the flag varieties 𝐆,μ,E\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu,E} and 𝐆,μ1,E\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu^{-1},E} respectively. For bb basic, we write 𝒪la\mathscr{O}^{la}_{\mathcal{M}} for 𝒪Gla\mathscr{O}^{G-la}_{\mathcal{M}}, by 5.1.9 there is no ambiguity in the locally analytic vectors for the group GG or GbG_{b}. Let 𝔤μ10=Lie𝐆(p)p𝒪𝐆,μ1\mathfrak{g}^{0}_{\mu^{-1}}=\operatorname{Lie}\mathbf{G}(\mathbb{Q}_{p})\otimes_{\mathbb{Q}_{p}}\mathscr{O}_{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu^{-1}}} and 𝔤μ0=LieGbp𝒪𝐆,μ\mathfrak{g}_{\mu}^{0}=\operatorname{Lie}G_{b}\otimes_{\mathbb{Q}_{p}}\mathscr{O}_{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu}} be the Lie algebroids over the Hodge-Tate and Grothendieck-Messing flag varieties respectively. Let 𝔫μ0𝔭μ0𝔤μ0\mathfrak{n}^{0}_{\mu}\subset\mathfrak{p}^{0}_{\mu}\subset\mathfrak{g}_{\mu}^{0} be the natural filtration on 𝐆,μ\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu} with Levi quotient 𝔪μ10\mathfrak{m}_{\mu^{-1}}^{0} (resp. for 𝔫μ10𝔭μ10𝔤μ10\mathfrak{n}^{0}_{\mu^{-1}}\subset\mathfrak{p}^{0}_{\mu^{-1}}\subset\mathfrak{g}^{0}_{\mu^{-1}} over 𝐆,μ1\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu^{-1}} with Levi quotient 𝔪μ0\mathfrak{m}^{0}_{\mu}). They arise from the 𝐏μ\mathbf{P}_{\mu}-equivariant filtration 𝔫μ𝔭μ𝔤C\mathfrak{n}_{\mu}\subset\mathfrak{p}_{\mu}\subset\mathfrak{g}_{C} and Levi quotient 𝔪μ:=𝔭μ/𝔫μ\mathfrak{m}_{\mu}:=\mathfrak{p}_{\mu}/\mathfrak{n}_{\mu} (resp. for the 𝐏μ1\mathbf{P}_{\mu^{-1}}-equivariant filtration 𝔫μ1𝔭μ1𝔤C\mathfrak{n}_{\mu^{-1}}\subset\mathfrak{p}_{\mu^{-1}}\subset\mathfrak{g}_{C} and Levi quotient 𝔪μ1=𝔭μ1/𝔫μ1\mathfrak{m}_{\mu^{-1}}=\mathfrak{p}_{\mu^{-1}}/\mathfrak{n}_{\mu^{-1}}). We identify the pullback of 𝔪μ0\mathfrak{m}^{0}_{\mu} and 𝔪μ10\mathfrak{m}^{0}_{\mu^{-1}} to 𝒪la\mathscr{O}^{la}_{\mathcal{M}} via Corollary 3.3.7 (after taking locally analytic vectors) and denote it 𝔪0,la\mathfrak{m}^{0,la}. The following is the generalization of [Pan22b, Corollary 5.3.13].

Theorem 5.1.10.

The actions of 𝔫μ0\mathfrak{n}_{\mu}^{0} and 𝔫μ10\mathfrak{n}_{\mu^{-1}}^{0} on 𝒪la\mathscr{O}^{la}_{\mathcal{M}} vanish. Furthermore, the actions of 𝔪μ0\mathfrak{m}^{0}_{\mu} and 𝔪μ10\mathfrak{m}^{0}_{\mu^{-1}} on 𝒪la\mathscr{O}^{la}_{\mathcal{M}} by derivations are identified via the pullback

𝔪μ10𝒪𝐆,μ1𝒪la=𝔪0,la=𝒪la𝒪𝐆,μ𝔪μ0.\mathfrak{m}^{0}_{\mu^{-1}}\otimes_{\mathscr{O}_{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu^{-1}}}}\mathscr{O}^{la}_{\mathcal{M}}=\mathfrak{m}^{0,la}=\mathscr{O}^{la}_{\mathcal{M}}\otimes_{\mathscr{O}_{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu}}}\mathfrak{m}^{0}_{\mu}.

In particular, the central character of the actions of 𝔪μ0\mathfrak{m}^{0}_{\mu} and 𝔪μ10\mathfrak{m}^{0}_{\mu^{-1}} on 𝒪la\mathscr{O}^{la}_{\mathcal{M}} agree under the natural isomorphism of the center of the enveloping algebras 𝒵(𝔪μ)C𝒵(𝔪μ1)C\mathcal{Z}(\mathfrak{m}_{\mu})_{C}\cong\mathcal{Z}(\mathfrak{m}_{\mu^{-1}})_{C}.

Proof.

The vanishing for the action of the geometric Sen operators follows from Theorem 4.3.3. We now prove the relation between the horizontal actions. In the following we forget about the action of the Galois group of EE and fix a trivialization of the Tate twist p(1)p\mathbb{Z}_{p}(1)\cong\mathbb{Z}_{p} obtained by fixing a sequence of pp-th power roots of unit (ζpn)n(\zeta_{p^{n}})_{n}. In the following all completed tensor products are solid.

Let ,Cla\mathcal{M}^{la}_{\infty,C} be the ringed space whose underlying topological space is |𝐆,b,μ,,C||\mathcal{M}_{\mathbf{G},b,\mu,\infty,C}| and sheaf of functions given by the algebra 𝒪la\mathscr{O}^{la}_{\mathcal{M}}. We have locally analytic Hodge-Tate period maps

,Cla{\mathcal{M}^{la}_{\infty,C}}𝐆,μ{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu}}𝐆,μ1.{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu^{-1}}.}πHTla\scriptstyle{\pi_{\operatorname{\scriptsize HT}}^{la}}πGMla\scriptstyle{\pi_{\operatorname{\scriptsize GM}}^{la}}

Let WW be a representation of the Levi 𝐌(=𝐌μ=𝐌μ1)\mathbf{M}(=\mathbf{M}_{\mu}=\mathbf{M}_{\mu^{-1}}), taking locally analytic vectors in Corollary 3.3.7 we get 𝐆(p)×Gb\mathbf{G}(\mathbb{Q}_{p})\times G_{b}-equivariant isomorphisms of vector bundles over ,Cla\mathcal{M}^{la}_{\infty,C}

𝒲𝐆,μ1(W)𝒪𝐆,μ1𝒪la=𝒪la𝒪𝐆,μ𝒲𝐆,μ(W).\mathcal{W}_{\mathbf{G},\mu^{-1}}(W)\otimes_{\mathscr{O}_{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu^{-1}}}}\mathscr{O}^{la}_{\mathcal{M}}=\mathscr{O}^{la}_{\mathcal{M}}\otimes_{\mathscr{O}_{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu}}}\mathcal{W}_{\mathbf{G},\mu}(W). (5.12)

Let 𝐌μ,GM𝐆,μ\mathbf{M}_{\mu,\operatorname{\scriptsize GM}}\to\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu} and 𝐌μ1,HT𝐆,μ1\mathbf{M}_{\mu^{-1},\operatorname{\scriptsize HT}}\to\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu^{-1}} be the natural 𝐌\mathbf{M}-torsors, the equation (5.12) gives rise to a natural isomorphism of 𝐌\mathbf{M}-torsors over ,Cla\mathcal{M}^{la}_{\infty,C}

πHTla,(𝐌μ1,HT)πGMla,(𝐌μ,GM).\pi_{\operatorname{\scriptsize HT}}^{la,*}(\mathbf{M}_{\mu^{-1},\operatorname{\scriptsize HT}})\cong\pi_{\operatorname{\scriptsize GM}}^{la,*}(\mathbf{M}_{\mu,\operatorname{\scriptsize GM}}).

Thus, if 𝐌μ,GMan\mathbf{M}^{an}_{\mu,\operatorname{\scriptsize GM}} and 𝐌μ,HTan\mathbf{M}_{\mu,\operatorname{\scriptsize HT}}^{an} denote the analytification of the algebraic torsors over the flag varieties, the period maps refined to a mixed period map

πGM,HTla:,Cla𝐌μ,GMan×𝐌an𝐌μ1,HTan.\pi^{la}_{\operatorname{\scriptsize GM},\operatorname{\scriptsize HT}}:\mathcal{M}^{la}_{\infty,C}\to\mathbf{M}^{an}_{\mu,\operatorname{\scriptsize GM}}\times^{\mathbf{M}^{an}}\mathbf{M}^{an}_{\mu^{-1},\operatorname{\scriptsize HT}}.

Note that the Lie algebra 𝔤×𝔤b\mathfrak{g}\times\mathfrak{g}_{b} acts on 𝐌μ,GMan×𝐌an𝐌μ1,HTan\mathbf{M}^{an}_{\mu,\operatorname{\scriptsize GM}}\times^{\mathbf{M}^{an}}\mathbf{M}^{an}_{\mu^{-1},\operatorname{\scriptsize HT}} by derivations, and by construction both horizontal actions 𝔪μ0\mathfrak{m}^{0}_{\mu} and 𝔪μ10\mathfrak{m}_{\mu^{-1}}^{0} are identified after pullback (similarly for the infinitesimal actions of 𝒵(𝔪μ)C\mathcal{Z}(\mathfrak{m}_{\mu})_{C} and 𝒵(𝔪μ1)C\mathcal{Z}(\mathfrak{m}_{\mu^{-1}})_{C}). Therefore, in order to show the theorem it suffices to show that the map of rings

πGM,HTla,1(𝒪𝐌μ,GMan×𝐌an𝐌μ1,HTan)𝒪la\pi^{la,-1}_{\operatorname{\scriptsize GM},\operatorname{\scriptsize HT}}(\mathscr{O}_{\mathbf{M}^{an}_{\mu,\operatorname{\scriptsize GM}}\times^{\mathbf{M}^{an}}\mathbf{M}^{an}_{\mu^{-1},\operatorname{\scriptsize HT}}})\to\mathscr{O}^{la}_{\mathcal{M}}

is dense in a suitable sense. To make this precise, for any compact open subgroup Kp𝐆(p)K_{p}\subset\mathbf{G}(\mathbb{Q}_{p}) consider the KpK_{p}-equivariant sheaf over 𝐆,μ1\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu^{-1}} given by Cla(Kp,𝒪𝐆,μ1)C^{la}(K_{p},\mathscr{O}_{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu^{-1}}}) and consider the colimit

Cla(𝔤μ10,𝒪𝐆,μ1):=limKpCla(Kp,𝒪𝐆,μ1).C^{la}(\mathfrak{g}^{0}_{\mu^{-1}},\mathscr{O}_{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu^{-1}}}):=\varinjlim_{K_{p}}C^{la}(K_{p},\mathscr{O}_{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu^{-1}}}).

Define

Cla(𝔤μ10/𝔫μ10,𝒪𝐆,μ1)=Cla(𝔤μ10,𝒪𝐆,μ1)𝔫μ1,10=0C^{la}(\mathfrak{g}^{0}_{\mu^{-1}}/\mathfrak{n}^{0}_{\mu^{-1}},\mathscr{O}_{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu^{-1}}})=C^{la}(\mathfrak{g}^{0}_{\mu^{-1}},\mathscr{O}_{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu^{-1}}})^{\mathfrak{n}^{0}_{\mu^{-1},\star_{1}}=0}

to be the invariant subspace of 𝔫μ10\mathfrak{n}^{0}_{\mu^{-1}}-horizontal sections for the left regular action. Let 𝒪𝐆(p)sm𝒪la\mathscr{O}^{\mathbf{G}(\mathbb{Q}_{p})-sm}_{\mathcal{M}}\subset\mathscr{O}^{la}_{\mathcal{M}} be the subalgebra of 𝐆(p)\mathbf{G}(\mathbb{Q}_{p})-smooth sections, equal to the colimit of the structural sheaves of the finite level local Shimura varieties 𝐆,b,μ,Kp,C\mathcal{M}_{\mathbf{G},b,\mu,K_{p},C}. By the proof of [RC24b, Proposition 6.2.8] (more precisely, Lemma 6.2.9), the pullback to 𝐆,b,μ,,C,v\mathcal{M}_{\mathbf{G},b,\mu,\infty,C,v}

Cla(𝔤μ10,𝒪^)=Cla(𝔤μ10,𝒪)^𝒪𝐆,μ1𝒪^C^{la}(\mathfrak{g}^{0}_{\mu^{-1}},\widehat{\mathscr{O}})=C^{la}(\mathfrak{g}^{0}_{\mu^{-1}},\mathscr{O}_{\operatorname{\mathcal{F}\ell}})\widehat{\otimes}_{\mathscr{O}_{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu^{-1}}}}\widehat{\mathscr{O}}

is a filtered colimit of ON Banach 𝒪^\widehat{\mathscr{O}}-modules which are relatively locally analytic in the sense of [RC23, Definition 1.0.1] (resp. for the pullback of Cla(𝔤μ10/𝔫μ10,𝒪^)C^{la}(\mathfrak{g}^{0}_{\mu^{-1}}/\mathfrak{n}^{0}_{\mu^{-1}},\widehat{\mathscr{O}})). Moreover, the 𝐆(p)\mathbf{G}(\mathbb{Q}_{p})-smooth vectors of Cla(𝔤μ10,𝒪^)C^{la}(\mathfrak{g}^{0}_{\mu^{-1}},\widehat{\mathscr{O}}) (more precisely, of its restriction Cla(𝔤μ10,𝒪^)C^{la}(\mathfrak{g}^{0}_{\mu^{-1}},\widehat{\mathscr{O}}_{\mathcal{M}}) to a sheaf on the topological space |𝐆,b,μ,,C||\mathcal{M}_{\mathbf{G},b,\mu,\infty,C}|) are well defined (by writing the module as a colimit of KpK_{p}-representations as Kp1K_{p}\to 1), and by construction they consist on the algebra 𝒪la\mathscr{O}^{la}_{\mathcal{M}}. By [RC23, Theorem 3.3.2 (2)] and the computation of the geometric Sen operators in Theorem 4.3.1, we have that

𝒪la\displaystyle\mathscr{O}^{la}_{\mathcal{M}} =Cla(𝔤μ10,𝒪^)𝐆(p)sm\displaystyle=C^{la}(\mathfrak{g}^{0}_{\mu^{-1}},\widehat{\mathscr{O}}_{\mathcal{M}})^{\mathbf{G}(\mathbb{Q}_{p})-sm}
=Cla(𝔤μ10,𝒪^)𝐆(p)sm,𝔫μ1,10=0\displaystyle=C^{la}(\mathfrak{g}^{0}_{\mu^{-1}},\widehat{\mathscr{O}}_{\mathcal{M}})^{\mathbf{G}(\mathbb{Q}_{p})-sm,\mathfrak{n}^{0}_{\mu^{-1},\star_{1}}=0}
=Cla(𝔤μ10/𝔫μ10,𝒪^)𝐆(p)sm.\displaystyle=C^{la}(\mathfrak{g}^{0}_{\mu^{-1}}/\mathfrak{n}^{0}_{\mu^{-1}},\widehat{\mathscr{O}}_{\mathcal{M}})^{\mathbf{G}(\mathbb{Q}_{p})-sm}.

But now Cla(𝔤μ10/𝔫μ10,𝒪^)C^{la}(\mathfrak{g}^{0}_{\mu^{-1}}/\mathfrak{n}^{0}_{\mu^{-1}},\widehat{\mathscr{O}}_{\mathcal{M}}) has trivial geometric Sen action, then [RC23, Theorem 3.3.2 (3)] implies that the orbit map (equivariant for the right regular action in the RHS)

𝒪laCla(𝔤μ10/𝔫μ10,𝒪^)\mathscr{O}^{la}_{\mathcal{M}}\to C^{la}(\mathfrak{g}^{0}_{\mu^{-1}}/\mathfrak{n}^{0}_{\mu^{-1}},\widehat{\mathscr{O}}_{\mathcal{M}})

extends to a 𝔤\mathfrak{g}-equivariant and 𝒪^\widehat{\mathscr{O}}_{\mathcal{M}}-linear isomorphism

𝒪la^𝒪𝐆(p)sm𝒪^Cla(𝔤μ10/𝔫μ10,𝒪^)\mathscr{O}^{la}_{\mathcal{M}}\widehat{\otimes}_{\mathscr{O}^{\mathbf{G}(\mathbb{Q}_{p})-sm}_{\mathcal{M}}}\widehat{\mathscr{O}}_{\mathcal{M}}\xrightarrow{\sim}C^{la}(\mathfrak{g}^{0}_{\mu^{-1}}/\mathfrak{n}^{0}_{\mu^{-1}},\widehat{\mathscr{O}}_{\mathcal{M}}) (5.13)

where the action of 𝔤\mathfrak{g} on the right hand side term is via the right regular action.

Let us write X=𝐌μ,GMan×𝐌an𝐌μ1,HTanX=\mathbf{M}^{an}_{\mu,\operatorname{\scriptsize GM}}\times^{\mathbf{M}^{an}}\mathbf{M}^{an}_{\mu^{-1},\operatorname{\scriptsize HT}}. Over XX we also have the Lie algebroid 𝔤μ1,X0/𝔫μ1,X0\mathfrak{g}^{0}_{\mu^{-1},X}/\mathfrak{n}^{0}_{\mu^{-1},X} which is nothing but the relative tangent space of XX over 𝐆,μ\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu} and we have an isomorphism

Cla(𝔤μ1,X0/𝔫X0,𝒪X)^𝒪X𝒪^=Cla(𝔤μ10/𝔫μ10,𝒪^)C^{la}(\mathfrak{g}^{0}_{\mu^{-1},X}/\mathfrak{n}^{0}_{X},\mathscr{O}_{X})\widehat{\otimes}_{\mathscr{O}_{X}}\widehat{\mathscr{O}}_{\mathcal{M}}=C^{la}(\mathfrak{g}^{0}_{\mu^{-1}}/\mathfrak{n}^{0}_{\mu^{-1}},\widehat{\mathscr{O}}_{\mathcal{M}})

equivariant for infinitesimal action of 𝐆(p)\mathbf{G}(\mathbb{Q}_{p}) for the left and right regular actions, and the action on the coefficients (by writing this sheaf as colimit of Banach sheaves where the actions integrate to compact open subgroups). Thus, we have a commutative diagram

𝒪la{\mathscr{O}^{la}_{\mathcal{M}}}Cla(𝔤μ10/𝔫μ10,𝒪^){C^{la}(\mathfrak{g}^{0}_{\mu^{-1}}/\mathfrak{n}^{0}_{\mu^{-1}},\widehat{\mathscr{O}}_{\mathcal{M}})}πGM,HTla,1(𝒪X){\pi^{la,-1}_{\operatorname{\scriptsize GM},\operatorname{\scriptsize HT}}(\mathscr{O}_{X})}πGM,HTla,1(Cla(𝔤μ1,X0/𝔫μ1,X0,𝒪X)){\pi^{la,-1}_{\operatorname{\scriptsize GM},\operatorname{\scriptsize HT}}(C^{la}(\mathfrak{g}^{0}_{\mu^{-1},X}/\mathfrak{n}^{0}_{\mu^{-1},X},\mathscr{O}_{X}))}

where we the horizontal maps are the orbit maps and the vertical maps are the natural inclusions. This diagram together with (5.13) show that both horizontal Levi actions of 𝔪μ0\mathfrak{m}^{0}_{\mu} and 𝔪μ10\mathfrak{m}^{0}_{\mu^{-1}} agree on 𝒪la\mathscr{O}^{la}_{\mathcal{M}} are they do over 𝒪X\mathscr{O}_{X} and they transform in the natural left regular action of

𝔪μ10𝒪𝐆,μ1𝒪X=𝔪X0=𝒪X𝒪𝐆,μ𝔪μ0\mathfrak{m}^{0}_{\mu^{-1}}\otimes_{\mathscr{O}_{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu^{-1}}}}\mathscr{O}_{X}=\mathfrak{m}^{0}_{X}=\mathscr{O}_{X}\otimes_{\mathscr{O}_{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu}}}\mathfrak{m}^{0}_{\mu}

on Cla(𝔤μ1,X0/𝔫μ1,X0,𝒪X)C^{la}(\mathfrak{g}^{0}_{\mu^{-1},X}/\mathfrak{n}^{0}_{\mu^{-1},X},\mathscr{O}_{X}). This ends the proof of the theorem. ∎

5.2. De Rham cohomology of the two towers

In this section we show that the sheaf 𝒪la\mathscr{O}^{la}_{\mathcal{M}} of Definition 5.1.7 produces an isomorphism between the de Rham cohomology (with compact supports) of the two towers for a duality of local Shimura varieties. Similar results have been obtained independently by Bosco-Dospinescu-Niziol. In order to state the theorem, wee keep the notation prior Theorem 5.1.10. Let X=𝐌μ,GMan×𝐌an𝐌μ1,HTanX=\mathbf{M}^{an}_{\mu,\operatorname{\scriptsize GM}}\times^{\mathbf{M}^{an}}\mathbf{M}^{an}_{\mu^{-1},\operatorname{\scriptsize HT}} and consider the mixed Lie algebroid over 𝒪la\mathscr{O}^{la}_{\mathcal{M}}

𝒯la=𝒯X𝒪X𝒪la\mathcal{T}^{la}=\mathcal{T}_{X}\otimes_{\mathscr{O}_{X}}\mathscr{O}^{la}_{\mathcal{M}}

obtained as the pullback of the tangent space of XX via the map πGM,HTla\pi^{la}_{\operatorname{\scriptsize GM},\operatorname{\scriptsize HT}}. By Theorem 5.1.10 this Lie algebroid acts by derivations on 𝒪la\mathscr{O}^{la}_{\mathcal{M}}, compatible with the derivations on XX. Indeed, let us write by 𝔫μ10,la𝔭μ10,la𝔤μ10,la\mathfrak{n}_{\mu^{-1}}^{0,la}\subset\mathfrak{p}^{0,la}_{\mu^{-1}}\subset\mathfrak{g}^{0,la}_{\mu^{-1}} the base change of 𝔫μ10𝔭μ10𝔤μ10\mathfrak{n}^{0}_{\mu^{-1}}\subset\mathfrak{p}^{0}_{\mu^{-1}}\subset\mathfrak{g}^{0}_{\mu^{-1}} from 𝒪𝐆,μ1\mathscr{O}_{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu^{-1}}} to 𝒪la\mathscr{O}^{la}_{\mathcal{M}}. Similarly, let 𝔫μ0,la𝔭μ0,la𝔤μ0,la\mathfrak{n}_{\mu}^{0,la}\subset\mathfrak{p}^{0,la}_{\mu}\subset\mathfrak{g}^{0,la}_{\mu} be the base change of 𝔫μ0𝔭μ0𝔤μ0\mathfrak{n}^{0}_{\mu}\subset\mathfrak{p}^{0}_{\mu}\subset\mathfrak{g}^{0}_{\mu} from 𝒪𝐆,μ\mathscr{O}_{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu}} to 𝒪la\mathscr{O}^{la}_{\mathcal{M}}. Theorem 5.1.10 also provides an isomorphism

𝔪μ10𝒪𝐆,μ1𝒪la𝔪0,la𝒪la𝒪𝐆,μ𝔪μ0.\mathfrak{m}^{0}_{\mu^{-1}}\otimes_{\mathscr{O}_{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu^{-1}}}}\mathscr{O}^{la}_{\mathcal{M}}\cong\mathfrak{m}^{0,la}\cong\mathscr{O}^{la}_{\mathcal{M}}\otimes_{\mathscr{O}_{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu}}}\mathfrak{m}^{0}_{\mu}.

Then, 𝒯la\mathcal{T}^{la} is the quotient of 𝔤μ0,la𝔤μ10,la\mathfrak{g}^{0,la}_{\mu}\oplus\mathfrak{g}_{\mu^{-1}}^{0,la} by the Lie algebroid 𝔭~0,la\widetilde{\mathfrak{p}}^{0,la} sitting in the cartesian square

𝔭~0,la{\widetilde{\mathfrak{p}}^{0,la}}𝔪0,la{\mathfrak{m}^{0,la}}𝔭μ0,la𝔭μ10,la{\mathfrak{p}^{0,la}_{\mu}\oplus\mathfrak{p}^{0,la}_{\mu^{-1}}}𝔪μ0,la𝔪μ10,la{\mathfrak{m}^{0,la}_{\mu}\oplus\mathfrak{m}^{0,la}_{\mu^{-1}}}(ι,ι)\scriptstyle{(\iota{,}-\iota)}

where (ι,ι)(\iota,-\iota) is the anti-diagonal map. Since 𝔭~0,la\widetilde{\mathfrak{p}}^{0,la} acts trivially on 𝒪la\mathscr{O}^{la}_{\mathcal{M}}, the action by derivations of 𝔤μ0,la𝔤μ10,la\mathfrak{g}^{0,la}_{\mu}\oplus\mathfrak{g}_{\mu^{-1}}^{0,la} on 𝒪la\mathscr{O}^{la}_{\mathcal{M}} descends to 𝒯la\mathcal{T}^{la}.

Remark 5.2.1.

With some additional effort one can prove that 𝒪la\mathscr{O}^{la}_{\mathcal{M}} is formally smooth over CC and that its tangent space is given by 𝒯la\mathcal{T}^{la} but we will not need this fact for the applications in this paper.

We have the following theorem

Theorem 5.2.2.

There are natural 𝐆(p)×Gb\mathbf{G}(\mathbb{Q}_{p})\times G_{b}-equivariant isomorphisms of de Rham complexes over the topological space |𝐆,b,μ,,C||\mathcal{M}_{\mathbf{G},b,\mu,\infty,C}|

DR(𝒪𝐆(p)sm)RΓ(𝒯la,𝒪la)DR(𝒪Gbsm),DR(\mathscr{O}^{\mathbf{G}(\mathbb{Q}_{p})-sm}_{\mathcal{M}})\cong R\Gamma(\mathcal{T}^{la},\mathscr{O}^{la}_{\mathcal{M}})\cong DR(\mathscr{O}^{G_{b}-sm}_{\mathcal{M}}), (5.14)

where:

  1. (1)

    DR(𝒪𝐆(p)sm)DR(\mathscr{O}^{\mathbf{G}(\mathbb{Q}_{p})-sm}_{\mathcal{M}}) is the de Rham complex of the colimit of structural sheaves of the finite level local Shimura varieties 𝐆,b,μ,Kp\mathcal{M}_{\mathbf{G},b,\mu,K_{p}} with Kp𝐆(p)K_{p}\subset\mathbf{G}(\mathbb{Q}_{p}).

  2. (2)

    DR(𝒪Gbsm)DR(\mathscr{O}^{G_{b}-sm}_{\mathcal{M}}) is the de Rham complex of the colimit of structural sheaves of the finite level local dual Shimura varieties 𝐆ˇ,bˇ,μˇ,Kb,p\mathcal{M}_{\check{\mathbf{G}},\check{b},\check{\mu},K_{b,p}} with Kb,p𝐆ˇ(p)=GbK_{b,p}\subset\check{\mathbf{G}}(\mathbb{Q}_{p})=G_{b}.

  3. (3)

    RΓ(𝒯la,𝒪la)R\Gamma(\mathcal{T}^{la},\mathscr{O}^{la}_{\mathcal{M}}) is the de Rham cohomology of 𝒪la\mathscr{O}^{la}_{\mathcal{M}} with respect to the action of the Lie algebroid 𝒯la\mathcal{T}^{la} acting by derivations.

In particular, we have a natural 𝐆(p)×Gb\mathbf{G}(\mathbb{Q}_{p})\times G_{b}-equivariant isomorphism of de Rham cohomologies with compact supports

limKpHdR,ci(𝐆,b,μ,Kp,C)limKb,pHdR,ci(𝐆ˇ,bˇ,μˇ,Kb,p,C).\varinjlim_{K_{p}}H^{i}_{dR,c}(\mathcal{M}_{\mathbf{G},b,\mu,K_{p},C})\cong\varinjlim_{K_{b,p}}H^{i}_{dR,c}(\mathcal{M}_{\check{\mathbf{G}},\check{b},\check{\mu},K_{b,p},C}).
Remark 5.2.3.

One can use the theory of the analytic de Rham stack of [RC24a] to prove Theorem 5.2.2. Indeed, as it was explained by Scholze to the second author, the formation of the analytic de Rham stack descends to (a suitable notion of) diamonds and, at least for what cohomology concerns, commutes with cofiltered limits of qcqs maps. Thus, Theorem 5.2.2 should be thought as an evidence to the fact that one has an equivalence of analytic de Rham stacks

limKp𝐆,b,μ,KpdR𝐆,b,μ,dRlimKb,p𝐆ˇ,bˇ,μˇ,Kb,pdR.\varprojlim_{K_{p}}\mathcal{M}_{\mathbf{G},b,\mu,K_{p}}^{dR}\cong\mathcal{M}_{\mathbf{G},b,\mu,\infty}^{dR}\cong\varprojlim_{K_{b,p}}\mathcal{M}_{\check{\mathbf{G}},\check{b},\check{\mu},K_{b,p}}^{dR}.

After taking quotients by the smooth groups 𝐆(p)sm\mathbf{G}(\mathbb{Q}_{p})^{sm} and 𝐆ˇ(p)sm\check{\mathbf{G}}(\mathbb{Q}_{p})^{sm} such an equivalence would also prove that one has an equivalence of analytic stacks

𝐆(p),μa,dR/𝐆ˇ(p)sm=𝐆(p),μ1a,dR/𝐆(p)sm\operatorname{\mathcal{F}\ell}^{a,dR}_{\mathbf{G}(\mathbb{Q}_{p}),\mu}/\check{\mathbf{G}}(\mathbb{Q}_{p})^{sm}=\operatorname{\mathcal{F}\ell}^{a,dR}_{\mathbf{G}(\mathbb{Q}_{p}),\mu^{-1}}/\mathbf{G}(\mathbb{Q}_{p})^{sm}

between the analytic de Rham stacks of the quotients of the admissible locus of the flag varieties. This gives rise to a “Jacquet-Langlands equivalence” of equivariant analytic DD-modules. We shall not prove this fact in this paper, instead we will give a first shadow of this compatibility of analytic de Rham stacks in the locally analytic Jacquet-Langlands functor for the Lubin-Tate tower in Section 5.3.

Proof of Theorem 5.2.2.

Let 𝒯μ\mathcal{T}_{\mu} and 𝒯μ1\mathcal{T}_{\mu^{-1}} be the tangent spaces of 𝐆,μ\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu} and 𝐆,μ1\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu^{-1}} respectively. We have identifications 𝒯μ=𝔤μ0/𝔭μ0\mathcal{T}_{\mu}=\mathfrak{g}^{0}_{\mu}/\mathfrak{p}^{0}_{\mu} and 𝒯μ1=𝔤μ10/𝔭μ10\mathcal{T}_{\mu^{-1}}=\mathfrak{g}^{0}_{\mu^{-1}}/\mathfrak{p}^{0}_{\mu^{-1}} via the anchor map (2.7). By construction of the Lie algebroid 𝒯la\mathcal{T}^{la}, we have a short exact sequence

0𝔪0,la𝒯la𝔤μ0,la/𝔭μ0,la𝔤μ10,la/𝔭μ10,la0.0\to\mathfrak{m}^{0,la}\to\mathcal{T}^{la}\to\mathfrak{g}^{0,la}_{\mu}/\mathfrak{p}^{0,la}_{\mu}\oplus\mathfrak{g}^{0,la}_{\mu^{-1}}/\mathfrak{p}^{0,la}_{\mu^{-1}}\to 0.

The pullback along the inclusion of 𝔤μ0,la/𝔭μ0,la\mathfrak{g}^{0,la}_{\mu}/\mathfrak{p}^{0,la}_{\mu} in the direct sum corresponds to the Lie algebroid 𝔤μ0,la/𝔫μ0,la\mathfrak{g}^{0,la}_{\mu}/\mathfrak{n}^{0,la}_{\mu} (similarly the pullback for the inclusion of 𝔤μ10,la/𝔭μ10,la\mathfrak{g}^{0,la}_{\mu^{-1}}/\mathfrak{p}^{0,la}_{\mu^{-1}} is 𝔤μ10,la/𝔫μ10,la\mathfrak{g}^{0,la}_{\mu^{-1}}/\mathfrak{n}^{0,la}_{\mu^{-1}}). Thus, we can write the 𝒯la\mathcal{T}^{la}-de Rham complex as the composite

RΓ(𝔤μ10,la/𝔭μ10,la,RΓ(𝔤μ0,la/𝔫μ0,la,𝒪la))RΓ(𝒯la,𝒪la)RΓ(𝔤μ0,la/𝔭μ0,la,RΓ(𝔤μ10,la/𝔫μ10,la,𝒪la)).R\Gamma(\mathfrak{g}^{0,la}_{\mu^{-1}}/\mathfrak{p}^{0,la}_{\mu^{-1}},R\Gamma(\mathfrak{g}^{0,la}_{\mu}/\mathfrak{n}^{0,la}_{\mu},\mathscr{O}^{la}_{\mathcal{M}}))\cong R\Gamma(\mathcal{T}^{la},\mathscr{O}^{la}_{\mathcal{M}})\cong R\Gamma(\mathfrak{g}^{0,la}_{\mu}/\mathfrak{p}^{0,la}_{\mu},R\Gamma(\mathfrak{g}^{0,la}_{\mu^{-1}}/\mathfrak{n}^{0,la}_{\mu^{-1}},\mathscr{O}^{la}_{\mathcal{M}})).

Therefore, in order to prove the quasi-isomorphisms (5.14) it suffices to show the two following facts:

  1. (1)

    The natural map 𝒪𝐆(p)smRΓ(𝔤μ10/𝔫μ10,𝒪la)\mathscr{O}_{\mathcal{M}}^{\mathbf{G}(\mathbb{Q}_{p})-sm}\to R\Gamma(\mathfrak{g}^{0}_{\mu^{-1}}/\mathfrak{n}_{\mu^{-1}}^{0},\mathscr{O}^{la}_{\mathcal{M}}) is an equivalence.

  2. (2)

    The natural map 𝒪𝐆ˇ(p)smRΓ(𝔤μ0/𝔫μ0,𝒪la)\mathscr{O}_{\mathcal{M}}^{\check{\mathbf{G}}(\mathbb{Q}_{p})-sm}\to R\Gamma(\mathfrak{g}^{0}_{\mu}/\mathfrak{n}_{\mu}^{0},\mathscr{O}^{la}_{\mathcal{M}}) is an equivalence.

These claims are symmetric with respect to the period maps, so it suffices to prove the first.

Let us write \mathcal{M}_{\infty} for the infinite level Shimura variety and Kp=/Kp\mathcal{M}_{K_{p}}=\mathcal{M}_{\infty}/K_{p} for its quotient by an open compact subgroup. Let gr0(𝒪𝔹dR)\operatorname{gr}^{0}(\mathscr{O}\!\mathbb{B}_{\operatorname{dR}}) be the Hodge-Tate proétale sheaf of Kp\mathcal{M}_{K_{p}} that appeared in Theorem 4.2.1, let 𝐍μ1𝐏μ1\mathbf{N}_{\mu^{-1}}\subset\mathbf{P}_{\mu^{-1}} be the unipotent radical and let 𝒪(𝐍μ1)\mathscr{O}(\mathbf{N}_{\mu^{-1}}) be the space of algebraic functions of endowed with the natural action of 𝐏μ1\mathbf{P}_{\mu^{-1}} as in Section 4.2. By Theorem 4.2.1 we have that

gr0(𝒪𝔹dR)=πHT𝒲𝐆,μ1(𝒪(𝐍μ1)).\operatorname{gr}^{0}(\mathscr{O}\!\mathbb{B}_{\operatorname{dR}})=\pi_{\operatorname{\scriptsize HT}}^{*}\mathcal{W}_{\mathbf{G},\mu^{-1}}(\mathscr{O}(\mathbf{N}_{\mu^{-1}})). (5.15)

Let νKp:Kp,C,proe´tKp,C,an\nu_{K_{p}}:\mathcal{M}_{K_{p},C,\operatorname{\scriptsize pro\acute{e}t}}\to\mathcal{M}_{K_{p},C,\operatorname{\scriptsize an}} be the projection of sites, then by [Sch13, Proposition 6.16] one has that

RνKp,gr0(𝒪𝔹dR,)=𝒪Kp,C.R\nu_{K_{p},*}\operatorname{gr}^{0}(\mathscr{O}\!\mathbb{B}_{\operatorname{dR},{\mathcal{M}}})=\mathscr{O}_{\mathcal{M}_{K_{p},C}}.

On the other hand, by taking locally analytic vectors for the action of 𝐆(p)\mathbf{G}(\mathbb{Q}_{p}), by the vanishing of higher locally analytic vectors of 𝒪^\widehat{\mathscr{O}}_{\mathcal{M}} of Theorem 4.3.3, and the group cohomology comparisons of [RJRC23, Theorem 6.3.4], one deduces that for VKpV\subset\mathcal{M}_{K_{p}} open affinoid and V=×KpVV_{\infty}=\mathcal{M}_{\infty}\times_{\mathcal{M}_{K_{p}}}V one has

𝒪Kp,C(V)\displaystyle\mathscr{O}_{\mathcal{M}_{K_{p},C}}(V) =RΓproe´t(V,gr0(𝒪𝔹dR,))\displaystyle=R\Gamma_{\operatorname{\scriptsize pro\acute{e}t}}(V,\operatorname{gr}^{0}(\mathscr{O}\!\mathbb{B}_{\operatorname{dR},{\mathcal{M}}}))
=RΓ(Kp,RΓproe´t(V,gr0(𝒪𝔹dR,)))\displaystyle=R\Gamma(K_{p},R\Gamma_{\operatorname{\scriptsize pro\acute{e}t}}(V_{\infty},\operatorname{gr}^{0}(\mathscr{O}\!\mathbb{B}_{\operatorname{dR},{\mathcal{M}}})))
=RΓ(Kp,RΓproe´t(V,gr0(𝒪𝔹dR,))RKpla)\displaystyle=R\Gamma(K_{p},R\Gamma_{\operatorname{\scriptsize pro\acute{e}t}}(V_{\infty},\operatorname{gr}^{0}(\mathscr{O}\!\mathbb{B}_{\operatorname{dR},{\mathcal{M}}}))^{RK_{p}-la})
=RΓ(Kp,RΓproe´t(V,𝒪^)RKpla𝒪𝐆,μ1𝒲𝐆,μ1(𝒪(𝐍μ1)))\displaystyle=R\Gamma(K_{p},R\Gamma_{\operatorname{\scriptsize pro\acute{e}t}}(V_{\infty},\widehat{\mathscr{O}}_{\mathcal{M}})^{RK_{p}-la}\otimes_{\mathscr{O}_{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu^{-1}}}}\mathcal{W}_{\mathbf{G},\mu^{-1}}(\mathscr{O}(\mathbf{N}_{\mu^{-1}})))
=RΓ(Kp,𝒪la(V)𝒪𝐆,μ1𝒲𝐆,μ1(𝒪(𝐍μ1)))\displaystyle=R\Gamma(K_{p},\mathscr{O}^{la}_{\mathcal{M}}(V_{\infty})\otimes_{\mathscr{O}_{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu^{-1}}}}\mathcal{W}_{\mathbf{G},\mu^{-1}}(\mathscr{O}(\mathbf{N}_{\mu^{-1}})))
=RΓsm(Kp,RΓ(𝔤,𝒪la(V)𝒪𝐆,μ1𝒲𝐆,μ1(𝒪(𝐍μ1))))\displaystyle=R\Gamma^{sm}(K_{p},R\Gamma(\mathfrak{g},\mathscr{O}^{la}_{\mathcal{M}}(V_{\infty})\otimes_{\mathscr{O}_{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu^{-1}}}}\mathcal{W}_{\mathbf{G},\mu^{-1}}(\mathscr{O}(\mathbf{N}_{\mu^{-1}}))))

where the second equality is decent along the KpK_{p}-torsor VVV_{\infty}\to V. The third equality is the comparison between solid and locally analytic group cohomology of [RJRC23, Theorem 6.3.4]. The fourth equality follows from the projection formula of locally analytic vectors [RJRC23, Corollary 3.1.15 (3)] and the isomorphism (5.15). The fifth equality follows from the vanishing of higher locally analytic vectors of Theorem 4.3.3. Finally, the sixth equality is the Lie algebra/smooth vs locally analytic cohomology comparison of Theorem [RJRC23, Theorem 6.3.4].

Taking colimits as Kp1K_{p}\to 1, we deduce that

𝒪𝐆(p)sm=limKpRνKp,gr0(𝒪𝔹dR)=RΓ(𝔤,𝒪la𝒪𝐆,μ1𝒲𝐆,μ1(𝒪(𝐍μ1)))).\mathscr{O}^{\mathbf{G}(\mathbb{Q}_{p})-sm}_{\mathcal{M}}=\varinjlim_{K_{p}}R\nu_{K_{p},*}\operatorname{gr}^{0}(\mathscr{O}\!\mathbb{B}_{\operatorname{dR}})=R\Gamma(\mathfrak{g},\mathscr{O}^{la}_{\mathcal{M}}\otimes_{\mathscr{O}_{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu^{-1}}}}\mathcal{W}_{\mathbf{G},\mu^{-1}}(\mathscr{O}(\mathbf{N}_{\mu^{-1}})))). (5.16)

But since 𝐍μ1\mathbf{N}_{\mu^{-1}} is an affine space we know that

RΓ(𝔫μ1,𝒪(𝐍μ1)C)=C,R\Gamma(\mathfrak{n}_{\mu^{-1}},\mathscr{O}(\mathbf{N}_{\mu^{-1}})_{C})=C,

taking the associated equivariant vector bundles over the flag variety and taking pullbacks along πHT\pi_{\operatorname{\scriptsize HT}} one deduces that

RΓ(𝔫μ10,la,𝒪la𝒪𝐆,μ1𝒲𝐆,μ1(𝒪(𝐍μ1))))=𝒪la.R\Gamma(\mathfrak{n}^{0,la}_{\mu^{-1}},\mathscr{O}^{la}_{\mathcal{M}}\otimes_{\mathscr{O}_{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu^{-1}}}}\mathcal{W}_{\mathbf{G},\mu^{-1}}(\mathscr{O}(\mathbf{N}_{\mu^{-1}}))))=\mathscr{O}^{la}_{\mathcal{M}}.

Combining this with (5.16), and by computing 𝔤μ10\mathfrak{g}^{0}_{\mu^{-1}}-lie algebra cohomology in two steps, one gets that

𝒪𝐆(p)sm\displaystyle\mathscr{O}^{\mathbf{G}(\mathbb{Q}_{p})-sm}_{\mathcal{M}} =RΓ(𝔤μ10,la/𝔫μ10,RΓ(𝔫μ10,la,𝒪la𝒪𝐆,μ1𝒲𝐆,μ1(𝒪(𝐍μ1)))\displaystyle=R\Gamma(\mathfrak{g}^{0,la}_{\mu^{-1}}/\mathfrak{n}^{0}_{\mu^{-1}},R\Gamma(\mathfrak{n}^{0,la}_{\mu^{-1}},\mathscr{O}^{la}_{\mathcal{M}}\otimes_{\mathscr{O}_{\operatorname{\mathcal{F}\ell}_{\mathbf{G},\mu^{-1}}}}\mathcal{W}_{\mathbf{G},\mu^{-1}}(\mathscr{O}(\mathbf{N}_{\mu^{-1}})))
=RΓ(𝔤μ10,la/𝔫μ10,la,𝒪la)\displaystyle=R\Gamma(\mathfrak{g}^{0,la}_{\mu^{-1}}/\mathfrak{n}^{0,la}_{\mu^{-1}},\mathscr{O}^{la}_{\mathcal{M}})

proving what we wanted.

The claim about the cohomology comparisons for the de Rham cohomology with compact supports follows for example by using the definition of compactly supported de Rham cohomology arising from the six functor formalism of analytic DD-modules of [RC24a]. One can also argue by using the adhoc definition of [GK00]. Indeed, the compactly supported cohomology of the de Rham complex of loc. cit. is nothing but the compactly supported cohomology of the de Rham complex seen as a sheaf on the underlying Berkovich space of 𝐆,b,μ,,C\mathcal{M}_{\mathbf{G},b,\mu,\infty,C}. To see that this cohomology with compact supports is well defined one can argue as follows: the map 𝐆,b,μ,,C𝐆,b,μ,Kp,C\mathcal{M}_{\mathbf{G},b,\mu,\infty,C}\to\mathcal{M}_{\mathbf{G},b,\mu,K_{p},C} gives rise to a KpK_{p}-torsor of Berkovich spaces

𝐆,b,μ,,CB𝐆,b,μ,Kp,CB.\mathcal{M}_{\mathbf{G},b,\mu,\infty,C}^{B}\to\mathcal{M}_{\mathbf{G},b,\mu,K_{p},C}^{B}. (5.17)

The space 𝐆,b,μ,Kp,CB\mathcal{M}_{\mathbf{G},b,\mu,K_{p},C}^{B} is a locally finite dimensional Hausdorff space (being the Berkovich space of a rigid space) and [HM24, Theorem 4.8.9 (i)] implies that 𝐆,b,μ,Kp,CB\mathcal{M}_{\mathbf{G},b,\mu,K_{p},C}^{B} has a well define functor of cohomology with compact supports for sheaves over p\mathbb{Q}_{p} (in the language of loc. cit. it is p\mathbb{Q}_{p}-fine). Since (5.17) is represented in profinite sets, 𝐆,b,μ,,CB\mathcal{M}_{\mathbf{G},b,\mu,\infty,C}^{B} is also a p\mathbb{Q}_{p}-fine map (this follows from [HM24, Theorem 3.4.11 (ii)] since any maps between profinite sets is p\mathbb{Q}_{p}-fine by construction, see Section 3.5.16 in loc. cit.), i.e. it has a well defined functor of cohomology with compact supports. ∎

5.3. The Jacquet-Langlands functor for admissible locally analytic representations

In this last section we recall the definition of the Jacquet-Langlands functor of [Sch18] for admissible Banach representations. We then proof that this functor is compatible with the passage to locally analytic vectors.

5.3.1. Scholze’s Jacquet-Langlands functor

Let n1n\geq 1 be an integer and F/pF/\mathbb{Q}_{p} a finite extension with ring of integers 𝒪F\mathcal{O}\subset F and ϖ𝒪\varpi\in\mathcal{O} a uniformizer. Let 𝔽=𝔽q\mathbb{F}=\mathbb{F}_{q} be the residue field of 𝒪\mathcal{O}. Consider the group 𝐆𝐋n,F\mathbf{GL}_{n,F}, μ\mu the cocharacter given by (1,0,,0)(1,0,\ldots,0) with n1n-1 occurrences of 0, and bb corresponds to a formal 𝒪\mathcal{O}-module 𝕏b\mathbb{X}_{b} over 𝔽¯\overline{\mathbb{F}} of dimension 11 and FF-height nn. Let DD be the division algebra over FF of invariant 1/n1/n, we have G~b=D×\widetilde{G}_{b}=D^{\times}. Finally, we fix p/F\mathbb{C}_{p}/F the pp-completion of an algebraic closure of FF.

Definition 5.3.1.

We let Def𝕏\operatorname{\scriptsize Def}_{\mathbb{X}} be the functor on formal schemes over 𝒪˘\breve{\mathcal{O}} sending SS to the set of isomorphism classes of pairs (X,ρ)(X,\rho), where X/SX/S is a formal FF-module, and ρ:X×SS¯𝕏×𝔽¯S¯\rho:X\times_{S}\overline{S}\xrightarrow{\sim}\mathbb{X}\times_{\overline{\mathbb{F}}}\overline{S} is a quasi-isogeny of formal 𝒪F\mathcal{O}_{F}-modules, where S¯=S×Spf𝒪˘Spec𝔽¯\overline{S}=S\times_{\operatorname{Spf}\breve{\mathcal{O}}}\operatorname{Spec}\overline{\mathbb{F}}.

By [RZ96] the functor Def𝕏\operatorname{\scriptsize Def}_{\mathbb{X}} is representable by a formal scheme 𝔐𝕏\mathfrak{M}_{\mathbb{X}} over Spf𝒪˘\operatorname{Spf}\breve{\mathcal{O}}, which is formally smooth and locally formally of finite type. We let 𝕏\mathcal{M}_{\mathbb{X}} denote the generic fiber of 𝔐𝕏\mathfrak{M}_{\mathbb{X}} as a rigid space.

Theorem 5.3.2 ([SW20, Corollary 24.3.5]).

There is a natural equivalence of diamonds 𝕏𝐆𝐋n,F,b,μ,K\mathcal{M}^{\lozenge}_{\mathbb{X}}\cong\mathcal{M}^{\lozenge}_{\mathbf{GL}_{n,F},b,\mu,K} with K=𝐆𝐋n(𝒪)K=\mathbf{GL}_{n}(\mathcal{O}).

Let =𝐆𝐋n,F,b,μ,\mathcal{M}^{\lozenge}_{\infty}=\mathcal{M}^{\lozenge}_{\mathbf{GL}_{n,F},b,\mu,\infty}. In this situation the 𝐆𝐋n(F)×D×\mathbf{GL}_{n}(F)\times D^{\times}-equivariant period maps (3.6) restrict to a diagram

{\mathcal{M}_{\infty}^{\lozenge}}F˘n1{\mathbb{P}^{n-1}_{\breve{F}}}ΩF˘{\Omega_{\breve{F}}}πHT\scriptstyle{\pi_{\operatorname{\scriptsize HT}}}πGM\scriptstyle{\pi_{\operatorname{\scriptsize GM}}}

where

  • πGM\pi_{\operatorname{\scriptsize GM}} is a proétale 𝐆𝐋n(F)\mathbf{GL}_{n}(F)-torsor and HH acts on F˘n1\mathbb{P}^{n-1}_{\breve{F}} via the natural inclusion of the map H×𝐆𝐋n(F˘)H^{\times}\subset\mathbf{GL}_{n}(\breve{F}).

  • πHT\pi_{\operatorname{\scriptsize HT}} is a proétale D×D^{\times}-torsor and ΩF˘F˘n1\Omega_{\breve{F}}\subset\mathbb{P}^{n-1}_{\breve{F}} is the 𝐆𝐋n(F)\mathbf{GL}_{n}(F)-stable open Drinfeld space obtained by removing all FF-rational hyperplanes.

Thus, we have an equivalence of vv-stacks

[F˘n1/D×][ΩF˘/𝐆𝐋n(F)].[\mathbb{P}^{n-1}_{\breve{F}}/D^{\times}]\cong[\Omega_{\breve{F}}/\mathbf{GL}_{n}(F)].

The Jacquet-Langlands functor is defined as follows.

Definition 5.3.3.

Let π\pi be a pp-power torsion admissible representation of 𝐆𝐋n(F)\mathbf{GL}_{n}(F) over p\mathbb{Z}_{p} and let π\mathcal{F}_{\pi} be the étale sheaf over F˘n1\mathbb{P}^{n-1}_{\breve{F}} obtained by descent along πGM\pi_{\operatorname{\scriptsize GM}}. The Jacquet-Langlands functor 𝒥\mathcal{JL} is the functor mapping such π\pi to the complex of smooth D×D^{\times}-representations

𝒥(π)=RΓe´t(pn1,π).\mathcal{JL}(\pi)=R\Gamma_{{\rm\acute{e}t}}(\mathbb{P}^{n-1}_{\mathbb{C}_{p}},\mathcal{F}_{\pi}).
Theorem 5.3.4 ([Sch18, Theorem 1.1]).

Let π\pi be a pp-power torsion admissible representation of 𝐆𝐋n(F)\mathbf{GL}_{n}(F) over p\mathbb{Z}_{p}, then 𝒥(π)\mathcal{JL}(\pi) is a complex of admissible representations of D×D^{\times}. In other words, for all ii\in\mathbb{Z} the cohomology

𝒥i(π)=He´ti(pn1,π)\mathcal{JL}^{i}(\pi)=H^{i}_{{\rm\acute{e}t}}(\mathbb{P}^{n-1}_{\mathbb{C}_{p}},\mathcal{F}_{\pi})

is an admissible representation of HH.

For convenience we shall consider the pp-completed analogue of Theorem 5.3.4. Let π\pi be a pp-adically complete admissible representation of 𝐆𝐋n(F)\mathbf{GL}_{n}(F) over p\mathbb{Z}_{p}, we shall write by π\mathcal{F}_{\pi} the pro-étale sheaf over F˘n1\mathbb{P}^{n-1}_{\breve{F}} given by the limit of étale sheaves π=limsπ/ps\mathcal{F}_{\pi}=\varprojlim_{s}\mathcal{F}_{\pi/p^{s}}. Finally, we denote by 𝒥(π)\mathcal{JL}(\pi) the pp-adically complete D×D^{\times}-representation

𝒥(π):=RΓproe´t(pn1,π)=RlimsRΓe´t(pn1,π/ps).\mathcal{JL}(\pi):=R\Gamma_{\operatorname{\scriptsize pro\acute{e}t}}(\mathbb{P}^{n-1}_{\mathbb{C}_{p}},\mathcal{F}_{\pi})=R\varprojlim_{s}R\Gamma_{{\rm\acute{e}t}}(\mathbb{P}^{n-1}_{\mathbb{C}_{p}},\mathcal{F}_{\pi/p^{s}}).
Corollary 5.3.5.

Let π\pi be a pp-adically complete admissible representation of 𝐆𝐋n(F)\mathbf{GL}_{n}(F), then 𝒥(π)\mathcal{JL}(\pi) is a complex of pp-adically complete admissible representations of D×D^{\times}. In other words, the cohomology groups

𝒥i(π)=Hproe´ti(pn1,π)\mathcal{JL}^{i}(\pi)=H^{i}_{\operatorname{\scriptsize pro\acute{e}t}}(\mathbb{P}^{n-1}_{\mathbb{C}_{p}},\mathcal{F}_{\pi})

are pp-adically complete admissible representations of D×D^{\times}. Moreover, we have that

𝒥i(π)=lims𝒥i(π/ps).\mathcal{JL}^{i}(\pi)=\varprojlim_{s}\mathcal{JL}^{i}(\pi/p^{s}).
Proof.

Let KDD×K_{D}\subset D^{\times} be a compact open subgroup which we assume to be an uniform pro-pp-group. Taking p\mathbb{Z}_{p}-duals the complex 𝒥(π)=RHom(𝒥(π),p)\mathcal{JL}(\pi)^{\vee}=R\mathrm{Hom}(\mathcal{JL}(\pi),\mathbb{Z}_{p}) is a pp-adically complete module over the Iwasawa algebra p,[KD]\mathbb{Z}_{p,\square}[K_{D}] whose reduction modulo pp is a perfect 𝔽p,[KD]\mathbb{F}_{p,\square}[K_{D}]-complex by Theorem 5.3.4, this implies that 𝒥(π)\mathcal{JL}(\pi)^{\vee} is itself a perfect complex of p,[KD]\mathbb{Z}_{p,\square}[K_{D}]-modules and so 𝒥(π)\mathcal{JL}(\pi) can be represented by a complex of admissible representations of HH. The rest of the statements are classical and left to the reader, see for example [Eme06, Proposition 1.2.12]. ∎

5.3.2. Locally analytic Jacquet-Langlands functor

Next we show that the Jacquet-Langlands functor of Definition 5.3.3 is compatible with locally analytic vectors. Let Π\Pi be an admissible locally analytic representation of 𝐆𝐋n(F)\mathbf{GL}_{n}(F) over p\mathbb{Q}_{p}, we let Π\mathcal{F}_{\Pi} be the proétale sheaf over F˘n1\mathbb{P}^{n-1}_{\breve{F}} whose SS-points for an affinoid perfectoid SF˘n1,S\to\mathbb{P}^{n-1,\lozenge}_{\breve{F}} are given by

(Π)=(C(|×F˘n1,S|,p)^pΠ)𝐆𝐋n(F)\mathcal{F}(\Pi)=(C(|\mathcal{M}_{\infty}\times_{\mathbb{P}^{n-1,\lozenge}_{\breve{F}}}S|,\mathbb{Q}_{p})\widehat{\otimes}_{\mathbb{Q}_{p}}\Pi)^{\mathbf{GL}_{n}(F)}

where

  • For a perfectoid space XX the algebra C(|X|,p)C(|X|,\mathbb{Q}_{p}) is the space of continuous functions from |X||X| to p\mathbb{Q}_{p}.

  • The completed tensor product is a tensor product of LB representations (equivalently a solid tensor product).

  • The group 𝐆𝐋n(F)\mathbf{GL}_{n}(F) acts via the diagonal action.

This is the same as the proétale solid sheaf on F˘n1\mathbb{P}^{n-1}_{\breve{F}} obtained by descent from the constant sheaf on \mathcal{M}_{\infty} via [AM24, Corollary 4.5].

Theorem 5.3.6.

Let π\pi be a pp-adically complete admissible representation and Π=(π[1p])𝐆𝐋n(F)la\Pi=(\pi[\frac{1}{p}])^{\mathbf{GL}_{n}(F)-la} its LB subrepresentation of locally analytic vectors. Then there is a natural equivalence

(𝒥(π)[1p])RD×laRΓproe´t(pn1,Π)(\mathcal{JL}(\pi)[\frac{1}{p}])^{RD^{\times}-la}\cong R\Gamma_{\operatorname{\scriptsize pro\acute{e}t}}(\mathbb{P}^{n-1}_{\mathbb{C}_{p}},\mathcal{F}_{\Pi}) (5.18)

where the left hand side is the complex of derived D×D^{\times}-analytic vectors of the solid D×D^{\times}-representation 𝒥(π)[1p]\mathcal{JL}(\pi)[\frac{1}{p}]. Moreover, for all ii\in\mathbb{Z} we have an isomorphism of locally analytic admissible D×D^{\times}-representations

(𝒥i(π)[1p])D×laHproe´ti(pn1,Π).(\mathcal{JL}^{i}(\pi)[\frac{1}{p}])^{D^{\times}-la}\cong H^{i}_{\operatorname{\scriptsize pro\acute{e}t}}(\mathbb{P}^{n-1}_{\mathbb{C}_{p}},\mathcal{F}_{\Pi}). (5.19)
Proof.

In the following proof we work with the derived \infty-categories of solid sheaves of diamonds as in [AM24, §4].

Step 0. The equivalence in Eq. 5.19 follows from Eq. 5.18. Indeed, the object 𝒥(π)[1p]\mathcal{JL}(\pi)[\frac{1}{p}] is a complex with cohomologies given by admissible Banach representations of D×D^{\times}. By [RJRC22, Proposition 4.48] (see also [RC24b, Proposition 2.3.1]) the higher locally analytic vectors of a Banach admissible representation vanish, then by the spectral sequence of [RC24b, Theorem 1.5] one deduces that

Hi((𝒥(π)[1p])RD×la)=(𝒥i(π)[1p])D×la.H^{i}((\mathcal{JL}(\pi)[\frac{1}{p}])^{RD^{\times}-la})=(\mathcal{JL}^{i}(\pi)[\frac{1}{p}])^{D^{\times}-la}.

Step 1. We first reinterpret the problem using the period sheaves. By [FS24, Proposition II.2.5] we have a short exact sequence of proétale sheaves

0p𝔹[1,p]φ1𝔹[1,1]0.0\to\mathbb{Q}_{p}\to\mathbb{B}_{[1,p]}\xrightarrow{\varphi-1}\mathbb{B}_{[1,1]}\to 0.

Taking solid (eq. pp-complete in this case) tensor products with the sheaf π\mathcal{F}_{\pi} we get a short exact sequence

0π[1p]𝔹[1,p]^pπ𝔹[1,1]^pπ0.0\to\mathcal{F}_{\pi}[\frac{1}{p}]\to\mathbb{B}_{[1,p]}\widehat{\otimes}_{\mathbb{Z}_{p}}\mathcal{F}_{\pi}\to\mathbb{B}_{[1,1]}\widehat{\otimes}_{\mathbb{Z}_{p}}\mathcal{F}_{\pi}\to 0.

Taking proétale cohomology we get an exact triangle

𝒥(π)[1p]RΓproe´t(pn1,𝔹[1,p]^pπ)RΓproe´t(pn1,𝔹[1,1]^pπ)+.\mathcal{JL}(\pi)[\frac{1}{p}]\to R\Gamma_{\operatorname{\scriptsize pro\acute{e}t}}(\mathbb{P}^{n-1}_{\mathbb{C}_{p}},\mathbb{B}_{[1,p]}\widehat{\otimes}_{\mathbb{Q}_{p}}\mathcal{F}_{\pi})\to R\Gamma_{\operatorname{\scriptsize pro\acute{e}t}}(\mathbb{P}^{n-1}_{\mathbb{C}_{p}},\mathbb{B}_{[1,1]}\widehat{\otimes}_{\mathbb{Q}_{p}}\mathcal{F}_{\pi})\xrightarrow{+}.

On the other hand, taking LB-completed tensor products we get a short exact sequence of proétale sheaves

0Π𝔹[1,p]^pΠ𝔹[1,1]^pΠ0.0\to\mathcal{F}_{\Pi}\to\mathbb{B}_{[1,p]}\widehat{\otimes}_{\mathbb{Q}_{p}}\mathcal{F}_{\Pi}\to\mathbb{B}_{[1,1]}\widehat{\otimes}_{\mathbb{Q}_{p}}\mathcal{F}_{\Pi}\to 0.

Therefore, to prove the theorem it suffices to show that for all I(0,)I\subset(0,\infty) compact interval, we have a natural equivalence of representations of D×D^{\times}

RΓproe´t(pn1,𝔹I^pπ)RD×laRΓproe´t(pn1,𝔹I^pΠ).R\Gamma_{\operatorname{\scriptsize pro\acute{e}t}}(\mathbb{P}^{n-1}_{\mathbb{C}_{p}},\mathbb{B}_{I}\widehat{\otimes}_{\mathbb{Z}_{p}}\mathcal{F}_{\pi})^{RD^{\times}-la}\cong R\Gamma_{\operatorname{\scriptsize pro\acute{e}t}}(\mathbb{P}^{n-1}_{\mathbb{C}_{p}},\mathbb{B}_{I}\widehat{\otimes}_{\mathbb{Z}_{p}}\mathcal{F}_{\Pi}). (5.20)

Step 2. We now reduce the proof of (5.20) to affinoid subspaces of pn1\mathbb{P}^{n-1}_{\mathbb{C}_{p}}. Let ν:p,proe´tn1p,ann1\nu:\mathbb{P}^{n-1}_{\mathbb{C}_{p},\operatorname{\scriptsize pro\acute{e}t}}\to\mathbb{P}^{n-1}_{\mathbb{C}_{p},\operatorname{\scriptsize an}} be the projection of sites and let 𝔘={Ui}iI\mathfrak{U}=\{U_{i}\}_{i\in I} be a finite rational open cover of pn1\mathbb{P}^{n-1}_{\mathbb{C}_{p}}. Then, for any proétale sheaf \mathscr{F} over pn1\mathbb{P}^{n-1}_{\mathbb{C}_{p}} we have equivalences of complexes

RΓˇan(𝔘,Rν)RΓproe´t(pn1,).R\check{\Gamma}_{\operatorname{\scriptsize an}}(\mathfrak{U},R\nu_{*}\mathscr{F})\cong R\Gamma_{\operatorname{\scriptsize pro\acute{e}t}}(\mathbb{P}^{n-1}_{\mathbb{C}_{p}},\mathscr{F}).

functorial on \mathscr{F}, where the left hand side is the Čech cohomology given by

RΓˇan(𝔘,Rν)=limVInt(𝔘)RΓan(V,Rν)R\check{\Gamma}_{\operatorname{\scriptsize an}}(\mathfrak{U},R\nu_{*}\mathscr{F})=\varprojlim_{V\in\mathrm{Int}(\mathfrak{U})}R\Gamma_{\operatorname{\scriptsize an}}(V,R\nu_{*}\mathscr{F})

with Int(𝔘)\mathrm{Int}(\mathfrak{U}) the poset of finite intersections of elements in 𝔘\mathfrak{U}.

Therefore, in order to show (5.20) it suffices to prove that for Upn1U\subset\mathbb{P}^{n-1}_{\mathbb{C}_{p}} a rational open subspace we have a natural equivalence

RΓproe´t(U,𝔹I^pπ)RHlaRΓproe´t(U,𝔹I^pΠ).R\Gamma_{\operatorname{\scriptsize pro\acute{e}t}}(U,\mathbb{B}_{I}\widehat{\otimes}_{\mathbb{Z}_{p}}\mathcal{F}_{\pi})^{RH-la}\cong R\Gamma_{\operatorname{\scriptsize pro\acute{e}t}}(U,\mathbb{B}_{I}\widehat{\otimes}_{\mathbb{Z}_{p}}\mathcal{F}_{\Pi}). (5.21)

Step 3. Finally, we prove (5.21). We can assume without loss of generality that UU is a rational subspace admitting a section U𝕏,pU\subset\mathcal{M}_{\mathbb{X},\mathbb{C}_{p}}. Let KD,UD×K_{D,U}\subset D^{\times} be a compact open subgroup stabilizing UU and let U=×𝕏UU_{\infty}=\mathcal{M}_{\infty}\times_{\mathcal{M}_{\mathbb{X}}}U. Then, we have that

RΓproe´t(U,𝔹I^pπ)RD×la\displaystyle R\Gamma_{\operatorname{\scriptsize pro\acute{e}t}}(U,\mathbb{B}_{I}\widehat{\otimes}_{\mathbb{Z}_{p}}\mathcal{F}_{\pi})^{RD^{\times}-la} RΓ(KD,U,RΓ(𝐆𝐋n(𝒪),RΓproe´t(U,𝔹I^pπ)^pLCla(KD,U,p))\displaystyle\cong R\Gamma(K_{D,U},R\Gamma(\mathbf{GL}_{n}(\mathcal{O}),R\Gamma_{\operatorname{\scriptsize pro\acute{e}t}}(U_{\infty},\mathbb{B}_{I}\widehat{\otimes}_{\mathbb{Z}_{p}}\mathcal{F}_{\pi})\widehat{\otimes}^{L}_{\mathbb{Q}_{p}}C^{la}(K_{D,U},\mathbb{Q}_{p}))
RΓ(KD,U×𝐆𝐋n(𝒪),RΓproe´t(U,𝔹I)^pLπ^pLCla(KD,U,p))\displaystyle\cong R\Gamma(K_{D,U}\times\mathbf{GL}_{n}(\mathcal{O}),R\Gamma_{\operatorname{\scriptsize pro\acute{e}t}}(U_{\infty},\mathbb{B}_{I})\widehat{\otimes}_{\mathbb{Z}_{p}}^{L}\pi\widehat{\otimes}_{\mathbb{Q}_{p}}^{L}C^{la}(K_{D,U},\mathbb{Q}_{p}))
RΓ(𝐆𝐋n(𝒪),RΓproe´t(U,𝔹I)RKHla^pLπ)\displaystyle\cong R\Gamma(\mathbf{GL}_{n}(\mathcal{O}),R\Gamma_{\operatorname{\scriptsize pro\acute{e}t}}(U_{\infty},\mathbb{B}_{I})^{RK_{H}-la}\widehat{\otimes}_{\mathbb{Z}_{p}}^{L}\pi)
RΓ(𝐆𝐋n(𝒪),RΓproe´t(U,𝔹I)RGla^pLπ)\displaystyle\cong R\Gamma(\mathbf{GL}_{n}(\mathcal{O}),R\Gamma_{\operatorname{\scriptsize pro\acute{e}t}}(U_{\infty},\mathbb{B}_{I})^{RG-la}\widehat{\otimes}_{\mathbb{Z}_{p}}^{L}\pi)
RΓ(𝐆𝐋n(𝒪),RΓproe´t(U,𝔹I)RGla^pLΠ)\displaystyle\cong R\Gamma(\mathbf{GL}_{n}(\mathcal{O}),R\Gamma_{\operatorname{\scriptsize pro\acute{e}t}}(U_{\infty},\mathbb{B}_{I})^{RG-la}\widehat{\otimes}_{\mathbb{Q}_{p}}^{L}\Pi)
RΓ(𝐆𝐋n(𝒪),RΓproe´t(U,𝔹I)^pLΠ)\displaystyle\cong R\Gamma(\mathbf{GL}_{n}(\mathcal{O}),R\Gamma_{\operatorname{\scriptsize pro\acute{e}t}}(U_{\infty},\mathbb{B}_{I})\widehat{\otimes}_{\mathbb{Q}_{p}}^{L}\Pi)
RΓproe´t(U,Π).\displaystyle\cong R\Gamma_{\operatorname{\scriptsize pro\acute{e}t}}(U,\mathcal{F}_{\Pi}).

In the first equivalence we use descent along the 𝐆𝐋n(𝒪)\mathbf{GL}_{n}(\mathcal{O})-torsor UUU_{\infty}\to U and write explicitly the definition of KD,UK_{D,U}-locally analytic vectors. The second equivalence is clear as UU_{\infty} is qcqs and π\pi is a Banach space, namely this follows from the analogue computations of the equation (5.2) in the proof of Lemma 5.1.3. The third equivalence follows from projection formula of locally analytic vectors [RJRC23, Corollary 3.1.15 (3)] and the fact that π\pi is a trivial KD,UK_{D,U}-representation. The fourth equivalence is 5.1.9. The fifth equivalence follows from the projection formula of locally analytic vectors and the fact that (π)[1p])RGla=Π(\pi)[\frac{1}{p}])^{RG-la}=\Pi as π\pi is an admissible representation. The sixth equivalence is the projection formula again. The last equivalence is descent along the torsor UUU_{\infty}\to U. This finishes the proof of the theorem. ∎

As a corollary we can prove that the Jacquet-Langlands functor for Banach admissible locally analytic representations preserves central characters.

Corollary 5.3.7.

Let π\pi be an admissible Banach representation of 𝐆𝐋n(L)\mathbf{GL}_{n}(L) over p\mathbb{Q}_{p} and suppose that Π=π𝐆𝐋n(L)la\Pi=\pi^{\mathbf{GL}_{n}(L)-la} has central character χ\chi. Then, for all ii\in\mathbb{Z}, the locally analytic D×D^{\times}-representation 𝒥i(π)D×la\mathcal{JL}^{i}(\pi)^{D^{\times}-la} has central character χ\chi under the natural identification 𝒵(LieD×)𝒵(LieG)\mathcal{Z}(\operatorname{Lie}D^{\times})\cong\mathcal{Z}(\operatorname{Lie}G).

Proof.

The statement can be proven after base change to CC. By [Sch18, Theorem 3.2] we have a natural equivalence

𝒥(π)^pp=RΓproe´t(pn1,π^p𝒪^).\mathcal{JL}(\pi)\widehat{\otimes}_{\mathbb{Q}_{p}}\mathbb{C}_{p}=R\Gamma_{\operatorname{\scriptsize pro\acute{e}t}}(\mathbb{P}^{n-1}_{\mathbb{C}_{p}},\mathcal{F}_{\pi}\widehat{\otimes}_{\mathbb{Q}_{p}}\widehat{\mathscr{O}}).

Then, by (5.20) we deduce an D×D^{\times}-equivariant equivalence

(𝒥(π)^pp)RD×laRΓproe´t(pn1,Π),(\mathcal{JL}(\pi)\widehat{\otimes}_{\mathbb{Q}_{p}}\mathbb{C}_{p})^{RD^{\times}-la}\cong R\Gamma_{\operatorname{\scriptsize pro\acute{e}t}}(\mathbb{P}^{n-1}_{\mathbb{C}_{p}},\mathcal{F}_{\Pi}),

thus it suffices to show that the RHS term has central character given by χ\chi. By picking a suitable affinoid cover {Ui}i\{U_{i}\}_{i} of pn1\mathbb{P}^{n-1}_{\mathbb{C}_{p}} as in Steps 2 and 3 of the proof of Theorem 5.3.6, we are reduced to show that for any small enough open affinoid U𝕏U\subset\mathcal{M}_{\mathbb{X}} with stabilizer KD,UD×K_{D,U}\subset D^{\times}, the central character of RΓproe´t(U,Π^p𝒪^)R\Gamma_{\operatorname{\scriptsize pro\acute{e}t}}(U,\mathcal{F}_{\Pi}\widehat{\otimes}_{\mathbb{Q}_{p}}\widehat{\mathscr{O}}) for the action of KD,UK_{D,U} is χ\chi. Let UU_{\infty}\subset\mathcal{M}_{\infty} be the pullback of UU to infinite level, by Step 3 of the proof of Theorem 5.3.6 we have that

RΓproe´t(U,Π^p𝒪^)RΓ(𝐆𝐋n(𝒪),RΓproe´t(U,𝒪^)RKD,Ula^pΠ),R\Gamma_{\operatorname{\scriptsize pro\acute{e}t}}(U,\mathcal{F}_{\Pi}\widehat{\otimes}_{\mathbb{Q}_{p}}\widehat{\mathscr{O}})\cong R\Gamma(\mathbf{GL}_{n}(\mathcal{O}),R\Gamma_{\operatorname{\scriptsize pro\acute{e}t}}(U_{\infty},\widehat{\mathscr{O}})^{RK_{D,U}-la}\widehat{\otimes}_{\mathbb{Q}_{p}}\Pi),

but by taking UU small enough, the vanishing of higher locally analytic vectors of Theorem 4.3.3 implies that

RΓproe´t(U,Π^p𝒪^)RΓ(𝐆𝐋n(𝒪),𝒪la(U)^pΠ).R\Gamma_{\operatorname{\scriptsize pro\acute{e}t}}(U,\mathcal{F}_{\Pi}\widehat{\otimes}_{\mathbb{Q}_{p}}\widehat{\mathscr{O}})\cong R\Gamma(\mathbf{GL}_{n}(\mathcal{O}),\mathscr{O}^{la}_{\mathcal{M}}(U_{\infty})\widehat{\otimes}_{\mathbb{Q}_{p}}\Pi).

The corollary follows from the identification of the central horizontal actions 𝒵(𝔪μ)p𝒵(𝔪μ1)p\mathcal{Z}(\mathfrak{m}_{\mu})_{\mathbb{C}_{p}}\cong\mathcal{Z}(\mathfrak{m}_{\mu^{-1}})_{\mathbb{C}_{p}} on 𝒪la(U)\mathscr{O}^{la}_{\mathcal{M}}(U_{\infty}) of Theorem 4.3.3 and the fact that the central actions of 𝒵(LieD×)𝒵(LieG)\mathcal{Z}(\operatorname{Lie}D^{\times})\cong\mathcal{Z}(\operatorname{Lie}G) factor through the horizontal actions. ∎

References

  • [AM24] Johannes Anschütz and Lucas Mann. Descent for solid quasi-coherent sheaves on perfectoid spaces. https://arxiv.org/abs/2403.01951, 2024.
  • [AMLB] Johannes Anschütz, Lucas Mann, and Arthur-César Le Bras. A 66-functor formalism for solid quasi-coherent sheaves on the Fargues-Fontaine curve. In preparation.
  • [BB81] Alexandre Beĭlinson and Joseph Bernstein. Localisation de 𝔤\mathfrak{g}-modules. C. R. Acad. Sci. Paris Sér. I Math., 292(1):15–18, 1981.
  • [BMS18] Bhargav Bhatt, Matthew Morrow, and Peter Scholze. Integral pp-adic Hodge theory. Publ. Math. Inst. Hautes Études Sci., 128:219–397, 2018.
  • [Bre10] Christophe Breuil. The emerging pp-adic Langlands programme. In Proceedings of the International Congress of Mathematicians. Volume II, pages 203–230. Hindustan Book Agency, New Delhi, 2010.
  • [CDN20] Pierre Colmez, Gabriel Dospinescu, and Wiesł awa Nizioł. Cohomologie pp-adique de la tour de Drinfeld: le cas de la dimension 1. J. Amer. Math. Soc., 33(2):311–362, 2020.
  • [CDN21] Pierre Colmez, Gabriel Dospinescu, and Wiesł awa Nizioł. Integral pp-adic étale cohomology of Drinfeld symmetric spaces. Duke Math. J., 170(3):575–613, 2021.
  • [CDN23] Pierre Colmez, Gabriel Dospinescu, and Wiesł awa Nizioł. Factorisation de la cohomologie étale pp-adique de la tour de Drinfeld. Forum Math. Pi, 11:Paper No. e16, 62, 2023.
  • [CDP14] Pierre Colmez, Gabriel Dospinescu, and Vytautas Paškūnas. The pp-adic local langlands correspondence for GL2(p)\mathrm{GL}_{2}(\mathbb{Q}_{p}). Camb. J. Math., 2(1):1–47, 2014.
  • [CS17] Ana Caraiani and Peter Scholze. On the generic part of the cohomology of compact unitary Shimura varieties. Ann. of Math. (2), 186(3):649–766, 2017.
  • [CS19] Dustin Clausen and Peter Scholze. Lectures on Condensed Mathematics. https://www.math.uni-bonn.de/people/scholze/Condensed.pdf, 2019.
  • [CS20] Dustin Clausen and Peter Scholze. Lectures on Analytic Geometry. https://www.math.uni-bonn.de/people/scholze/Analytic.pdf, 2020.
  • [DLB17] Gabriel Dospinescu and Arthur-César Le Bras. Revêtements du demi-plan de Drinfeld et correspondance de Langlands pp-adique. Ann. of Math. (2), 186(2):321–411, 2017.
  • [Eme06] Matthew Emerton. On the interpolation of systems of eigenvalues attached to automorphic Hecke eigenforms. Invent. Math., 164(1):1–84, 2006.
  • [Fal02] Gerd Faltings. A relation between two moduli spaces studied by V. G. Drinfeld. In Algebraic number theory and algebraic geometry, volume 300 of Contemp. Math., pages 115–129. Amer. Math. Soc., Providence, RI, 2002.
  • [Far08] Laurent Fargues. L’isomorphisme entre les tours de Lubin-Tate et de Drinfeld et applications cohomologiques. In L’isomorphisme entre les tours de Lubin-Tate et de Drinfeld, volume 262 of Progr. Math., pages 1–325. Birkhäuser, Basel, 2008.
  • [FS24] Laurent Fargues and Peter Scholze. Geometrization of the local Langlands correspondence, 2024.
  • [GK00] Elmar Grosse-Klönne. Rigid analytic spaces with overconvergent structure sheaf. J. Reine Angew. Math., 519:73–95, 2000.
  • [HM24] Claudius Heyer and Lucas Mann. 6-Functor Formalisms and Smooth Representations, 2024.
  • [Hub94] Roland Huber. A generalization of formal schemes and rigid analytic varieties. Math. Z., 217(4):513–551, 1994.
  • [Hub96] Roland Huber. Étale cohomology of rigid analytic varieties and adic spaces. Aspects of Mathematics, E30. Friedr. Vieweg & Sohn, Braunschweig, 1996.
  • [KHH+19] Kiran S. Kedlaya, David Hansen, Bastian Haase, Shizhang Li, Ruochuan Liu, Peter Scholze, Annie Carter, Zonglin Jiang, Jake Postema, Daniel Aaron Smith, Claus Mazanti Sorensen, and Xin Tong. Sheaves, stacks, and shtukas. Perfectoid Spaces, 2019.
  • [Kot97] Robert E. Kottwitz. Isocrystals with additional structure. II. Compos. Math., 109(3):255–339, 1997.
  • [Lur09] Jacob Lurie. Higher topos theory, volume 170 of Ann. Math. Stud. Princeton, NJ: Princeton University Press, 2009.
  • [Lur17] Jacob Lurie. Higher algebra. 2017.
  • [Man22a] Lucas Mann. The 6-functor formalism for \mathbb{Z}_{\ell}- and \mathbb{Q}_{\ell}-sheaves on diamonds. https://arxiv.org/abs/2209.08135, 2022.
  • [Man22b] Lucas Mann. A pp-adic 6-Functor Formalism in Rigid-Analytic Geometry. https://arxiv.org/abs/2206.02022, 2022.
  • [Pan22a] Lue Pan. On locally analytic vectors of the completed cohomology of modular curves. Forum of Mathematics, Pi, 10:e7, 2022.
  • [Pan22b] Lue Pan. On locally analytic vectors of the completed cohomology of modular curves II. https://arxiv.org/abs/2209.06366, 2022.
  • [RC23] Juan Esteban Rodríguez Camargo. Geometric Sen theory over rigid analytic spaces, 2023.
  • [RC24a] Juan Esteban Rodríguez Camargo. The analytic de rham stack in rigid geometry. https://arxiv.org/abs/2401.07738, 2024.
  • [RC24b] Juan Esteban Rodríguez Camargo. Locally analytic completed cohomology, 2024.
  • [RJRC22] Joaquín Rodrigues Jacinto and Juan Esteban Rodríguez Camargo. Solid locally analytic representations of pp-adic Lie groups. Represent. Theory, 26:962–1024, 2022.
  • [RJRC23] Joaquín Rodrigues Jacinto and Juan Esteban Rodríguez Camargo. Solid locally analytic representations. https://arxiv.org/abs/2305.03162, 2023.
  • [RV14] Michael Rapoport and Eva Viehmann. Towards a theory of local Shimura varieties. Münster J. Math., 7(1):273–326, 2014.
  • [RZ96] M. Rapoport and Th. Zink. Period spaces for pp-divisible groups, volume 141 of Annals of Mathematics Studies. Princeton University Press, Princeton, NJ, 1996.
  • [Sch13] Peter Scholze. pp-adic Hodge theory for rigid-analytic varieties. Forum Math. Pi, 1:e1, 77, 2013.
  • [Sch18] Peter Scholze. On the pp-adic cohomology of the Lubin-Tate tower. Ann. Sci. Éc. Norm. Supér. (4), 51(4):811–863, 2018. With an appendix by Michael Rapoport.
  • [Sch22] Peter Scholze. Etale cohomology of diamonds. https://arxiv.org/abs/1709.07343v3, 2022.
  • [ST03] Peter Schneider and Jeremy Teitelbaum. Algebras of pp-adic distributions and admissible representations. Invent. Math., 153(1):145–196, 2003.
  • [SW13] Peter Scholze and Jared Weinstein. Moduli of pp-divisible groups. Camb. J. Math., 1(2):145–237, 2013.
  • [SW20] Peter Scholze and Jared Weinstein. Berkeley Lectures on p-adic Geometry: (AMS-207). Princeton University Press, 2020.
  • [Vez19] Alberto Vezzani. A motivic version of the theorem of Fontaine and Wintenberger. Compos. Math., 155(1):38–88, 2019.